Skip to main content
Advertisement
  • Loading metrics

A chemosensory-like histidine kinase is dispensable for chemotaxis in vitro but regulates the virulence of Borrelia burgdorferi through modulating the stability of RpoS

  • Ching Wooen Sze,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Validation, Visualization, Writing – original draft, Writing – review & editing

    Affiliation Department of Oral Craniofacial Molecular Biology, Virginia Commonwealth University, Richmond, Virginia, United States of America

  • Kai Zhang,

    Roles Investigation, Methodology, Validation

    Affiliation Department of Oral Craniofacial Molecular Biology, Virginia Commonwealth University, Richmond, Virginia, United States of America

  • Michael J. Lynch,

    Roles Data curation, Formal analysis, Investigation, Methodology, Validation, Visualization, Writing – original draft, Writing – review & editing

    Affiliation Department of Chemistry and Chemical Biology, Cornell University, Ithaca, New York, United States of America

  • Radha Iyer,

    Roles Formal analysis, Investigation, Methodology

    Affiliation Department of Pathology, Microbiology, and Immunology, New York Medical College, Valhalla, New York, United States of America

  • Brian R. Crane,

    Roles Conceptualization, Funding acquisition, Resources, Supervision, Validation, Writing – review & editing

    Affiliation Department of Chemistry and Chemical Biology, Cornell University, Ithaca, New York, United States of America

  • Ira Schwartz,

    Roles Conceptualization, Investigation, Methodology, Project administration, Resources, Supervision, Writing – original draft, Writing – review & editing

    Affiliation Department of Pathology, Microbiology, and Immunology, New York Medical College, Valhalla, New York, United States of America

  • Chunhao Li

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

    cli5@vcu.edu.

    Affiliations Department of Oral Craniofacial Molecular Biology, Virginia Commonwealth University, Richmond, Virginia, United States of America, Department of Microbiology and Immunology, Virginia Commonwealth University, Richmond, Virginia, United States of America

Abstract

As an enzootic pathogen, the Lyme disease bacterium Borrelia burgdorferi possesses multiple copies of chemotaxis proteins, including two chemotaxis histidine kinases (CHK), CheA1 and CheA2. Our previous study showed that CheA2 is a genuine CHK that is required for chemotaxis; however, the role of CheA1 remains mysterious. This report first compares the structural features that differentiate CheA1 and CheA2 and then provides evidence to show that CheA1 is an atypical CHK that controls the virulence of B. burgdorferi through modulating the stability of RpoS, a key transcriptional regulator of the spirochete. First, microscopic analyses using green-fluorescence-protein (GFP) tags reveal that CheA1 has a unique and dynamic cellular localization. Second, loss-of-function studies indicate that CheA1 is not required for chemotaxis in vitro despite sharing a high sequence and structural similarity to its counterparts from other bacteria. Third, mouse infection studies using needle inoculations show that a deletion mutant of CheA1 (cheA1mut) is able to establish systemic infection in immune-deficient mice but fails to do so in immune-competent mice albeit the mutant can survive at the inoculation site for up to 28 days. Tick and mouse infection studies further demonstrate that CheA1 is dispensable for tick colonization and acquisition but essential for tick transmission. Lastly, mechanistic studies combining immunoblotting, protein turnover, mutagenesis, and RNA-seq analyses reveal that depletion of CheA1 affects RpoS stability, leading to reduced expression of several RpoS-regulated virulence factors (i.e., OspC, BBK32, and DbpA), likely due to dysregulated clpX and lon protease expression. Bulk RNA-seq analysis of infected mouse skin tissues further show that cheA1mut fails to elicit mouse tnf-α, il-10, il-1β, and ccl2 expression, four important cytokines for Lyme disease development and B. burgdorferi transmigration. Collectively, these results reveal a unique role and regulatory mechanism of CheA1 in modulating virulence factor expression and add new insights into understanding the regulatory network of B. burgdorferi.

Author summary

Lyme disease is an infectious disease caused by the bacterium Borrelia burgdorferi which is transmitted to humans through the bite of infected blacklegged ticks. It is the most commonly reported tick-borne illness in North America and Europe. B. burgdorferi is a highly invasive bacterium that can swim and penetrate deep into tissues to cause a wide range of disorders including arthritis, meningitis, and migratory musculoskeletal pain. This invasiveness is driven by the ability of the spirochete to sense and respond to its surrounding initiated by signal transduction coupled with a sophisticated array of sensory proteins located at both cell ends. The genome of B. burgdorferi encodes multiple copies of chemotaxis-related genes and the function of many of these genes remains unknown. In this report, we focus on deciphering the role of one of the chemotaxis histidine kinases, CheA1. Our results reveal a novel regulatory contribution of CheA1 to the infectious life cycle and pathogenesis of B. burgdorferi.

Introduction

Chemotaxis allows motile bacteria to sense surrounding environments and accordingly navigate their swimming behaviors to promote their survival and ability to colonize and invade a host [1]. The bacterial chemotaxis signaling pathway relies on a two-component system (TCS) that contains a chemotaxis histidine kinase CheA and a response regulator CheY. In Escherichia coli, the histidine kinase CheA is a key component of the chemotaxis signaling pathway [14]. Chemotaxis is initiated when the membrane bound methyl-accepting chemotaxis proteins (MCPs) bind a ligand (i.e., attractants) leading to activation of signal transduction from the MCP to the histidine kinase CheA via a linker protein, CheW. Activated CheA through autophosphorylation then transfers its phosphate group to the chemotaxis response regulator CheY, which in turn binds to the flagellar motor switch proteins to modulate flagellar motor rotation allowing cells to swim toward or away from an environment.

Lyme disease (LD), also known as Lyme borreliosis, is the most commonly reported tick-borne illness in the United States and Europe, with more than 476,000 new cases annually in the U.S. [5]. LD is caused by the bacterium Borrelia burgdorferi (also known as Borreliella burgdorferi) which is transmitted to humans through the bite of infected Ixodes ticks. Upon dissemination from the bite site, highly motile spirochetes can invade other tissues and distal parts of the body. Early-stage LD can be treated effectively with antibiotics. Unrecognized LD, however, can have debilitating long-term effects, such as arthritis, meningitis, and migratory musculoskeletal pain, which can appear after months or even years of infection [6,7].

The genome of B. burgdorferi encodes for two cheA (cheA1 and cheA2), three cheY (cheY1, cheY2, and cheY3), and three cheW (cheW1, cheW2, and cheW3) [8,9]. The majority of these genes are located within two gene clusters: cheA2-cheW3-cheX-cheY3, and cheW2-bb0566-cheA1-cheB2-bb0569-cheY2 [911]. The current evidence has revealed that all of the chemotaxis genes in the cheA2 cluster are required for chemotaxis [1113]. In contrast, the genes in the cheW2 cluster that have been studied (with the exception of bb0569, which encodes for an atypical MCP [14]) are not involved in chemotaxis in vitro. For instance, cheW2 and cheY2 mutants have a similar in vitro phenotype as the wild type with no apparent defect in chemotactic function [11,13,15,16]. Our previous study showed that CheA2 was essential for the chemotaxis of B. burgdorferi, e.g., the cheA2 mutant failed to reverse, ran constantly in one direction, and was non-chemotactic to attractants [11]. In addition, in vivo studies showed that CheA2 was essential for B. burgdorferi to establish infection in a mammalian host as well as to accomplish its enzootic life cycle [17].

The role of CheA1 currently remains unknown. CheA1 contains all the characteristic domains of a histidine kinase similar to CheA2 but the P2 domain required for the binding of CheY response regulator is not as conserved as other known CheAs [11]. A recent in vivo study by Xu et al., as well as our group, indicated that cheY2 and cheW2 are dispensable for chemotaxis and motility of B. burgdorferi in vitro but essential for a productive infection within the vertebrate host [15,18]. In order to comprehend the function of the cheW2 operon, and in particular CheA1, in the chemotaxis and pathogenicity of B. burgdorferi, a cheA1 mutant was generated in the virulent B31 A3-68 ∆bbe02 background [19] and its phenotype was characterized in great detail using different in vitro and in vivo approaches. The ability of this mutant to complete the enzootic infection cycle was analyzed and its contribution to RpoS-mediated regulation was explored.

Results

Structural comparison of B. burgdorferi CheA1 and CheA2

To investigate why CheA1 is dispensable for chemotaxis in vitro while CheA2 is required [11], we generated AlphaFold [20] models of the two CheA proteins (BbCheA1 and BbCheA2) and compared them to the solved structure of CheA from Thermotoga maritima (Tm) [21,22] (Fig 1A). The models show high confidence (pLDDT scores >80) for all domains in each monomer, with pLDDT scores < 50 only for the residues that comprise the P1-P2 and P2-P3 linkers. The domain composition of TmCheA is similar to BbCheA1 and BbCheA2 with some notable differences (Fig 1B). BbCheA1 has an analogous domain arrangement compared to TmCheA with the only major difference being an extended P2-P3 linker (73 residues-long compared to 42 residues of TmCheA). In contrast, BbCheA2 differs from BbCheA1 and TmCheA in several distinct ways. First, BbCheA2 has two P2 domains (CheY binding domains, referred to here as P2α and P2β) compared to a single P2 domain in TmCheA and BbCheA1. Superimposition of the P2α and P2β domains reveals a high degree of structural similarity despite a mere 26% identity (Fig 1C). Sequence alignment suggests that both P2 domains bind to CheY, which could potentially act to increase the local phospho-CheY concentration within the chemosensory arrays of B. burgdorferi (S1A Fig), as has been shown for the P2 domain of the canonical E. coli CheA [23]. The second major difference between BbCheA1 and BbCheA2 is that BbCheA2 has an extended P3 domain (Fig 1B). CheA homodimerizes via P3-P3’ interactions, wherein two adjacent P3 domains come together to form a four-helix bundle [21]. In BbCheA2, the 122 residues in P3 domain is significantly longer than the ~53 residues domain found in TmCheA and BbCheA1 due to inserts between residues Ile446-Leu515 (Fig 1A and 1B).

thumbnail
Fig 1. Structural comparison of B. burgdorferi CheA1 and CheA2.

(A) AlphaFold models of B. burgdorferi CheA1 (BbCheA1) and CheA2 (BbCheA2) subunits compared to Thermotoga maritima CheA (TmCheA). Protein structures are colored according to their pLDDT scores [20]. (B) Domain organization of BbCheA1, BbCheA2, and TmCheA. The first P2 domain of BbCheA2 (referred to here as P2α) extends from Glu153 to Lys265 and is connected to the second P2 domain (P2β) via an 18-residue long P2α -P2β linker. P2β is comprised of residues Glu283 to Val387 and is connected to the dimerization domain (P3) via a 32-residue long P2β -P3 linker. Overall, the P1-P2α and P2β -P3 linkers consist of approximately 50 residues, similar to the 43-residue long P2-P3 linker of TmCheA. (C) Superimposition of BbCheA P2α (RMSD: 0.48 Å, cyan, C150-V267) and P2β (gray, F279-V387). (D) Superimposition of phospho-transfer domain (P1), CheY-binding domain (P2), dimerization domain (P3), kinase domain (P4), and regulatory domain (P5) of TmCheA (tan), BbCheA1 (pink), and BbCheA2 (green). RMSD of each alignment is indicated in the lower region of each panel. All structural superimposition and figures were generated in PyMol [105].

https://doi.org/10.1371/journal.ppat.1011752.g001

To gain more insight into the overall differences in the domain architecture of CheA1 and CheA2, we superimposed domains P1-P5 from BbCheA1 (pink) and BbCheA2 (green) to the analogous domains of TmCheA (tan, Fig 1D). Overall, the P1, P4, and P5 domains are quite similar in all three proteins (RMSDs < 1), whereas the analogous P2 and P3 domains vary substantially (RMSD ranging 0.71–10.2 Å) among the three CheAs. For the P2 domain superimposition, the domain cores are similar with most of the structural deviations arising from the loops connecting the β-strands, and the N- and C-termini connecting to the P1-P2 and P2-P3 linkers, respectively. Analysis of the putative CheY binding site for BbCheA1 P2 and BbCheA2 P2 (α and β) reveals that the structures of the P2 domains are generally well-conserved in this region, supporting sequence alignment data that indicates conservation in the CheY:P2 binding interface residues (S1A–S1B Fig). Superimposition of the P3 domains from TmCheA, BbCheA1, and BbCheA2 AlphaFold models reveals the extent by which BbCheA2 P3 domain differs from that of TmCheA and BbCheA1, whose P3 domains align with an RMSD of 0.7 Å. Compared to TmCheA P3, BbCheA2 P3 aligns with an RMSD of 3.7 Å and is approximately 43 Å longer. The extended P3 domain may be important for maintaining the proper protein-protein interactions between CheA subunits within a homodimer and/or between CheA and the chemosensory receptors in the arrays [21,24].

To better contextualize the differences in P2 and P3 predicted for the BbCheAs, we performed multiple-sequence alignment analysis on discrete groups of CheA sequences within the spirochete phylum. We first examined all the CheA sequences in spirochete species that contained only one CheA gene, assuming that this single CheA would be associated with chemotaxis signaling. Within this category, we examined only those members from the order Treponematales. Multiple sequence alignments of these CheA sequences reveal they have a single P2 domain and an extended P3 domain (S2 Fig). These results indicate that while the extra P2 domain may or may not play a role in chemotaxis for other spirochete species, the extended P3 domain is a conserved structural motif for CheAs associated with chemotaxis in spirochetes. We next looked at all CheA1 and CheA2 sequences from species in the Borreliaceae family. Multiple alignment data from these sequences reveal that in species with two distinct CheA isoforms, all the CheA1 proteins encode a single P2 domain and have a short P3 domain similar to BbCheA1 and TmCheA (S3 Fig), and all CheA2 proteins encode for a second P2 domain and an extended P3 domain (S3 Fig). These data suggest that while an extended P3 domain appears to be important for CheA proteins associated with chemotaxis in diverse spirochete species, the addition of a second P2 domain may be a Borrelia spp. and Borreliella spp. specific modification.

Isolation of cheA1 mutant and its complemented strain

To investigate the role of CheA1 in the infectious cycle of B. burgdorferi, we have attempted to inactivate cheA1 in the virulent B31 A3-68 Δbbe02 strain [25]. pGA1kan, a previously constructed vector [11], was linearized and electro-transformed into wild-type competent cells. A previously described PCR analysis was first carried out to screen for clones with the desired targeted mutagenesis [11]. One clone (referred to as cheA1mut) was selected for further analysis. For the complementation, the plasmid CheA1/pBBE22 (Fig 2A) was electro-transformed into the cheA1mut competent cells. The presence of CheA1/pBBE22 in antibiotic-resistant colonies was confirmed by immunoblots using anti-CheA1 antibody and anti-DnaK antibody as a control. As shown in Fig 2B, a band of approximately 80 kDa was detected in the wild type and the cheA1com strain but was absent in cheA1mut, indicating that the cognate gene product was abrogated in the mutant and restored in the complemented strain (Fig 2B). B. burgdorferi B31 strain contains 21 linear and circular plasmids [26] of which some are essential for its infectivity but are easily lost during in vitro cultivation [27]. To check for the plasmid profile of our isolated strains, a previously developed PCR method was used to determine if the obtained positive clones contain the same plasmids as their parental strains [28]. Our results showed that cheA1mut (S4B Fig) and its isogenic complemented strain (S4C Fig) contained the same plasmid profile as the wild-type strain (S4A Fig). Growth analysis indicated that deletion of cheA1 did not affect the fitness of the mutant at 23°C,34°C or 37°C (Fig 2C–2E) indicating that depletion of CheA1 has no impact on B. burgdorferi growth in vitro.

thumbnail
Fig 2. Characterizations of cheA1 mutant (cheA1mut) and its isogenic complemented strain (cheA1com).

(A) Deletion of cheA1 has no impact on B. burgdorferi growth. Construction of CheA1/pBBE22 for the complementation of cheA1mut. bb0566-cheA1 fragment was PCR amplified using primer pair P1/P2 and cloned into pBBE22 [17,99], a shuttle vector of B. burgdorferi, yielding CheA1/pBBE22. (B) Immunoblot analysis of WT, cheA1mut and cheA1com strains. The same amounts of WT, cheA1mut and cheA1com whole-cell lysates were analyzed by SDS-PAGE and then probed with CheA1 and DnaK antibodies, as previously documented [11,97]. Growth analysis of WT, cheA1mut and cheA1com under routine laboratory condition (C) (34°C/pH7.6), (D) tick-like condition (23°C/pH7.6), and (E) host condition (37°C/pH6.8). Growth curve analysis was carried out to determine if CheA1 affects cell growth. 104 or 105 cells/ml of bacteria were inoculated into 10 ml BSK-II media and cultivated under 34°C/pH7.6, 23°C/pH7.6, or 37°C/pH6.8. Cell numbers were enumerated every 1–4 days until cells entered stationary phase. Cell counting was repeated in triplicate with at least two independent samples, and the results are expressed as means ± SEM.

https://doi.org/10.1371/journal.ppat.1011752.g002

The cellular localization of CheA1 is temporally and spatially regulated

As mentioned earlier, the genome of B. burgdorferi encodes for multiple copies of chemotaxis genes. It was speculated that B. burgdorferi may possess two chemotaxis pathways that function in different hosts during the infection cycle [8,11,13,18,29], i.e., CheA2, CheW3, and CheY3 form a pathway that executes chemotaxis in mammalian hosts, whereas CheA1, CheW2, and CheY2 form a distinct pathway that is activated in the tick vector. Consistent with this speculation, our previous work showed that inactivation of cheA2 impaired the ability of the spirochete to establish infection in mice but not in ticks [17]. In order to decipher the function of CheA1 in the enzootic cycle of B. burgdorferi, a GFP reporter construct (S5A Fig) was created and used to complement cheA1mut. The reporter strain was cultivated under laboratory conditions as well as conditions mimicking the tick milieu in vitro and observed microscopically. Surprisingly, we noticed that CheA1-GFP was localized to one cell pole when cultivated under the tick-phase condition but appeared diffused at elevated temperature (Fig 3A). The diffused pattern was not due to degradation of the GFP-fusion protein at elevated temperature as full-length CheA1-GFP fusion protein was detected with no excessive degradation (S5C Fig). This pattern was not observed in a strain carrying only the empty GFP vector (Fig 3A) or in a strain expressing CheA2-GFP fusion protein (S5C–S5D Fig). This result indicates that the polar localization pattern is unique to CheA1 (Fig 3A). In addition, the localization of CheA1 appeared to be connected with cell division. During early log phase, the CheA1-GFP signal was localized to one cell pole. As the cells entered mid log phase, the signal of CheA1-GFP was seen distributed to the mid cell as well as both cell poles (Fig 3B Day 7 onward). As the cells entered late log to stationary phase, CheA1-GFP signal was seen distributed along the cell body in specific puncta as indicated by the white arrows in Fig 3B Day 13–17 (Fig 3B). When the cells were cultivated under elevated temperature, the polar localization of CheA1-GFP disappeared (S6 Fig). At elevated temperature, during the early growth phase, CheA1-GFP signal appeared dispersed along the cell body and only when the cells entered stationary phase single pole localization reappeared as was the case with cells cultivated at 23°C (S6B Fig), suggesting that the localization of CheA1 is regulated spatially and temporally. In contrast to CheA1-GFP, the localization of GFP and CheA2-GFP remained evenly distributed along the cells under all culture conditions and growth phases (S6A and S6C Fig). Collectively, these results strongly suggest that the localization of CheA1 is dynamic throughout the growth of B. burgdorferi and that CheA1 function may be more prominent during the tick phase of the enzootic cycle. In addition, pH (acidic when cell density increases and enters stationary phase [30]) or temperature may have an impact on the localization and function of CheA1.

thumbnail
Fig 3. CheA1 has a unique polar localization when cultivated under tick-like condition.

(A) In vitro localization of CheA1. To localize CheA1, a CheA1-GFP reporter construct (S3 Fig) was used to complement cheA1mut. The obtained cells were cultivated under unfed tick (UF) condition at 23°C/pH 7.6 or routine laboratory culture condition at 34°C/pH 7.6. Cells were monitored for the presence of GFP signal and images were taken at early log phase at ×200 magnification using a Zeiss Axiostar Plus microscope. Scale bars represent 10 μm. A cheA1mut carrying a GFP reporter construct was used as a control to confirm that the polar localization seen in CheA1-GFP complemented strain is due to CheA1 and not an artifact from GFP protein. (B) Time course images of CheA1-GFP strain cultivated under UF condition. 105 cells/ml of CheA1-GFP strain was inoculated into 10 ml fresh BSK-II and cultivated under UF condition. Images were taken every 2 days to monitor the localization of CheA1-GFP signals. As cells entered late log to early stationary phase, increasing numbers of discrete bright puncta were observed evenly distributed along the cell body, coincided with the zones of new peptidoglycan formation that mark the division sites for daughter cells [106].

https://doi.org/10.1371/journal.ppat.1011752.g003

CheA1 has no impact on B. burgdorferi swimming behavior and chemotaxis in vitro

B. burgdorferi has three different swimming modes: run, flex, and reverse [11,12]. Our previous work showed that deletion of cheA1 did not have impacts on the swimming behavior of B. burgdorferi in vitro [11]. All of these studies, however, were conducted using a high-passage avirulent B31A strain. To test whether this is the case in the virulent strain background, the swimming behaviors of the wild-type, cheA1mut, and cheA1com cells were analyzed using a computer-assisted cell tracker as documented previously [11]. Similar to the WT, cheA1mut showed three swimming modes–runs interrupted by reverses/flexes, with a mean swimming velocity of 12.5 ± 0.5 μm/sec compared to 13.7 ± 0.3 and 12.9 ± 0.3 μm/sec of WT and cheA1com, respectively (Fig 4A). In addition, capillary tube assays using N-acetyl-D-glucosamine (NAG) as a chemoattractant [31] (Fig 4B) and swimming plate analysis (Fig 4C–4D) did not reveal significant differences among WT, cheA1mut, and cheA1com, further supporting that CheA1 is dispensable for B. burgdorferi motility and chemotaxis in vitro.

thumbnail
Fig 4. CheA1 is dispensable for spirochete motility and chemotaxis in vitro.

(A) Bacteria tracking analysis. The swimming velocities of WT, cheA1mut, and cheA1com were measured using a computer-based motion tracking system as previously described [11]. At least 25 cells were tracked and the swimming velocities were measured. (B) Capillary tube chemotaxis assay. This experiment was carried out using 100 mM N-acetyl-D-glucosamine (NAG) as an attractant according to a previous report [31]. The results are expressed as fold increases in cell number entering the capillary tubes containing the attractant relative to the number entering control tubes without attractant (buffer alone). Results are expressed as means ± SEM for five tubes. A 2-fold increase is considered the threshold for a positive response. (C) and (D) Swimming plate assays. The assay was carried out on 0.35% agarose containing 1:10-diluted BSK-II medium as described previously [100]. The flgE mutant, a previously constructed nonmotile mutant [103], was included to determine the sizes of the inocula. The data are presented as mean diameters (in millimeters) of rings ± SEM for twelve plates. All assays were repeated at least twice with independent replicates and representative data are shown here. * significant difference (P < 0.05).

https://doi.org/10.1371/journal.ppat.1011752.g004

cheA1mut fails to establish systemic infection in immunocompetent mice by needle inoculation

To determine if CheA1 is essential for the virulence of B. burgdorferi BALB/c mice were infected intradermally with 1 x 105 cells of the WT, cheA1mut, and cheA1com strains (3–7 mice per strain). Tissues from the ear, skin, heart, and joint were harvested three weeks post-infection and transferred to BSK-II medium for re-isolation. The wild-type strain was successfully recovered from all of the infected animals (7/7) while cheA1com was successfully recovered from some tissues of five infected mice (5/5), indicating a partial complementation. Interestingly, cheA1mut was only recovered from the skin at the site of inoculation and not from distal organs (Table 1). To rule out the potential of a spontaneous mutation in our mutant, a new deletion construct was produced to replace the entire cheA1 gene in-frame with a kanamycin (kan) cassette (S4D Fig) and the mouse infection study was repeated. The newly constructed mutant (cheA1IFD) was able to survive at the inoculation site but failed to establish systemic infection in immunocompetent mice (Table 1), similar to cheA1mut. Taken together, these results suggest that CheA1 is not essential for B. burgdorferi colonization of skin, but is required for dissemination from the initial inoculation site and/or to establish a systemic infection in mice.

thumbnail
Table 1. cheA1mut is unable to cause systemic infection in immunocompetent mice.a

https://doi.org/10.1371/journal.ppat.1011752.t001

cheA1mut can persist and proliferate at the site of inoculation

Intradermal infection of the mutant failed to cause systemic infection but the mutant could be re-isolated from the initial inoculation site, suggesting that the mutant is able to survive and persist at the site of injection and evade the host immune system. To determine how long the mutant can persist at the inoculation site, 1 x 105 cells of WT or cheA1mut were injected subcutaneously into BALB/c mice and the inoculation site was marked for bacterial re-isolation at different time points as indicated in the Material and Methods. qRT-PCR analysis showed that the mutant was not only able to persist but replicate at the inoculation site to the same extent as the WT as indicated by the increase in flaB transcript copies for up to 28 days (Fig 5).

thumbnail
Fig 5. cheA1mut is able to persist at the site of inoculation.

To determine if CheA1 is required for B. burgdorferi persistence at the inoculation site, a time point infection study was performed. 18 immunocompetent mice were subcutaneously infected with 105 of WT or cheA1mut and mice were sacrificed at the indicated time points. Skin tissue from the inoculation sites were recovered to assess for borrelial burden using qRT-PCR as described [17]. The data are presented as mean flaB transcript copies over 105 of mouse β-actin ± SEM for each group of mice.

https://doi.org/10.1371/journal.ppat.1011752.g005

cheA1mut is defective in hematogenous dissemination

We reasoned that the failure of the mutant to cause systemic infection could be due to the following three possibilities: 1) the mutant failed to disseminate from the injection site; 2) the mutant successfully disseminated from the injection site into the bloodstream but failed to spread to distal organs due to a defect in tissue tropism; or 3) the mutant successfully disseminated into the bloodstream but was cleared by the host immune system. To test these possibilities, the above animal study was repeated using severe combined immuno-deficiency (SCIDs) mice. Based on the tissue re-isolation results in Table 2, WT and cheA1com were successfully recovered from all of the harvested tissues. In contrast, the cheA1mut mutant was able to disseminate and cause systemic infection in 50% of the injected SCID mice. Consistent with the result obtained using immunocompetent mice, cheA1mut was able to colonize and persist in the skin better than in other tissues (Table 2). qRT-PCR analysis indicated that of those 50% SCID mice with systemic infection, no statistically significant difference in tissue spirochetal burdens was detected between the wild type, cheA1mut, and cheA1com with the exception of the ear and heart tissues (Fig 6). This result strongly suggests that in the absence of an adaptive immune response, cheA1mut can disseminate from the injection site and cause systemic infection albeit to an attenuated degree. In the immunocompetent mice, cheA1mut most likely failed to evade the host adaptive immune response and was cleared from the animal before it was able to colonize and establish infection in distal organs. Furthermore, blood culture showed that 3 out of 6 mice were positive for cheA1mut when injected subcutaneously into SCID mice (Table 2), further substantiating that the mutant is able to disseminate from the inoculation site and spread systemically through the bloodstream. Taken together, the mouse infection studies indicate that in the absence of CheA1, B. burgdorferi is attenuated in both innate and adaptive immune evasion and tissue colonization.

thumbnail
Table 2. cheA1mut is still attenuated in distal organ colonization in SCID mice.b

https://doi.org/10.1371/journal.ppat.1011752.t002

thumbnail
Fig 6. cheA1mut is attenuated in causing systemic infection in immunodeficient mice.

To determine if CheA1 is required for B. burgdorferi immune evasion, needle infection study was repeated using severe combined immunodeficiency (SCIDs) mice. For this study, 105 of WT, cheA1mut and cheA1com strains were subcutaneously inoculated into SCID mice and sacrificed three weeks after infection. Skin tissues around and distal from the injection site were harvested along with tissues from the ear, joint, heart and spleen to assess for borrelial burden using qRT-PCR as described [17]. The data are presented as mean flaB transcript copies over 105 of mouse β-actin ± SEM. *, significant difference (P < 0.05).

https://doi.org/10.1371/journal.ppat.1011752.g006

cheA1mut is attenuated in colonization and adaptive immune evasion

The above study showed that the mutant has reduced infectivity but is still able to disseminate into the bloodstream to cause systemic infection in SCID mice but not in immunocompetent mice. It is possible that in the presence of adaptive immunity, the low number of mutant cells that survived was unable to disseminate and cause systemic infection. To rule out this possibility, a 10-fold higher number (1 x 106 cells) of spirochetes were directly introduced into the bloodstream of BALB/c mice via tail-vein injection to determine if high doses of bacteria can overcome the dissemination and colonization barrier. As a control, a non-motile, non-infectious flaBmut strain was included [32]. 1 x 106 of WT, cheA1mut or flaBmut cells in 100 μl of PBS were injected into the tail-vein of BALB/c mice. Tissues were harvested two weeks post-injection to check for dissemination. All mice injected with the WT strain showed disseminated infection (3/3) whereas the flaBmut control was non-infectious as expected (0/3). For mice injected with the cheA1mut strain, live spirochetes were successfully recovered from the skin of only 1 out of 5 mice infected, indicating that the mutant is attenuated in tissue colonization even at a high inoculum introduced directly into the bloodstream (Table 3). In addition, no mutant cells were recovered from the blood of any mice injected with cheA1mut. This result clearly indicates that CheA1 is essential for the survival of B. burgdorferi in the blood. Consistent with live culture recovery, high levels of flaB transcripts were detected from mice infected with the WT strain whereas cheA1mut-infected mice had less than one copy of flaB transcript per 105 of mouse β-actin transcript overall (most likely due to residual B. burgdorferi that was initially introduced into the bloodstream (Fig 7A)). These findings further substantiate the key role of CheA1 in B. burgdorferi immune evasion and tissue colonization.

thumbnail
Table 3. cheA1mut has defects in dissemination and immune evasion.c

https://doi.org/10.1371/journal.ppat.1011752.t003

thumbnail
Fig 7. cheA1mut has defects in dissemination and serum resistance.

(A) To determine if high dose of inoculum can restore cheA1mut immune evasion and colonization barriers, 106 of WT, cheA1mut and non-motile flaBmut [32] strains were injected into wild type mice via tail-vein and sacrificed two weeks after infection. Tissues from the skin, ear, joint, heart and spleen were harvested to assess for borrelial burden using qRT-PCR as described [17]. The data are presented as mean flaB transcript copies over 105 of mouse β-actin ± SEM. (B) Serum bactericidal activity. 106 cells of WT, cheA1mut and cheA1com were incubated with 20% normal human serum (NHS) or heat-inactivated serum (hiNHS) for 2 hours. The percentage of viable cells in NHS treated cells were determined after normalization to their respective hiNHS-treated groups. The data shown here are means of three independent replicates ± SEM. *, significant difference (P < 0.05).

https://doi.org/10.1371/journal.ppat.1011752.g007

cheA1mut is more susceptible to serum bactericidal activity

B. burgdorferi is able to resist the bactericidal activity of serum [33,34] as it produces several virulence factors that protects the spirochetes from complement mediated killing, such as BBK32, OspC, and CRASP proteins [see recent reviews [3538]]. The attenuated infectivity of cheA1mut in both the SCID mice and tail-vein infection model suggests that CheA1 is important for the immune evasion of B. burgdorferi. To determine if deletion of CheA1 renders the spirochete susceptible to serum bactericidal activity, WT, cheA1mut and cheA1com were treated with normal human serum (NHS) or heat-inactivated NHS (hiNHS) and the percentage of viable cells after treatment was normalized against the input cells. Our data showed that cheA1mut was more susceptible to serum killing with ~25% survival rate as compared to ~60% survival rate of the WT and cheA1com in NHS (Fig 7B). This observation is in congruence with the SCID mice study where only 50% of cheA1mut was able to survive in the blood of infected mice and the complemented strain fully restored the survival rate to WT level (Table 2). The reduced survival of the mutant in the bloodstream can likely be attributed to its susceptibility to serum bactericidal activity.

CheA1 is not required for colonization of Ixodes scapularis ticks

As cheA1mut failed to cause disseminated infection in wild-type mice, we used a microinjection-based infection procedure to artificially infect ticks [39]. For this study, Naïve I. scapularis nymphs were microinjected with equal amounts of WT, cheA1mut, or cheA1com strains. After the infection, ticks were allowed to feed to repletion on naïve C3H mice (5 ticks/ mouse and 3 mice for each strain). The spirochete burdens in engorged ticks were measured by qRT-PCR as described previously [17,40,41]. As shown in Fig 8, the bacterial burden of cheA1mut in fed ticks was not significantly different from that of the WT and cheA1com (Fig 8A). This result indicates that CheA1 is not required for B. burgdorferi colonization and survival in the tick vector.

thumbnail
Fig 8. CheA1 is not required for B. burgdorferi acquisition and survival in tick but necessary for transmission.

(A) Detection of spirochete burdens in microinjected nymphal ticks after feeding. RNA samples were extracted from whole fed ticks (after repletion; 5 to 7 days) and subjected to qRT-PCR analysis. The bacterial burdens in ticks were measured by the number of copies of flaB transcript compared to the number of copies of tick β-actin transcript as previously described [17]. The data are presented as the means of relative levels of flaB transcript ± SEM for each strain (WT, cheA1mut, and cheA1com). (B) Detection of spirochete burdens in mice infected via tick bite. At day 14 after tick feeding, mice were sacrificed and tissues from the skin, heart, joint, and bladder were harvested for qRT-PCR analysis as previously documented [17,107]. No trace of flaB transcript was detected in the mouse tissues fed upon by cheA1mut infected ticks while 1 in 3 mice fed upon by cheA1com showed positive results at the skin and heart tissues. *, significant difference (P < 0.05). (C) Detection of spirochete burdens in naïve nymphal ticks fed on infected mice. C3H mice were artificially infected with WT or cheA1mut strain via needle inoculation. Naïve nymphal ticks were confined to the injection site and allowed to feed to repletion. After 72 hours, fed nymphs were collected and individually tested for the presence of flaB via qPCR. Experiments were repeated twice and data are presented as mean flaB copies per fed nymph from both data sets ± SEM.

https://doi.org/10.1371/journal.ppat.1011752.g008

cheA1mut fails to be transmitted to mice via tick bite

The infectivity of cheA1mut in mice was first evaluated via needle inoculation. However, this method is different from the natural route of mammalian infection through tick bite. As cheA1mut is still able to establish infection in ticks (Fig 8A), we examined whether the mutant can be transmitted to mice when fed upon by infected ticks. Groups of naïve C3H mice were allowed to be parasitized by artificially infected ticks as detailed in the Material and Methods. Seven days after tick feeding, the mice were sacrificed and specimens from the skin, ear, heart, bladder and blood were collected and subjected to qRT-PCR analysis to determine the pathogen level. Consistent with the results obtained via needle inoculation (Table 1), only the tissue specimens from mice fed on by WT- and cheA1com-infected ticks were positive for flaB transcripts via qRT-PCR analysis (Table 4). flaB transcripts were only detected from the skin and heart of one mouse fed upon by cheA1com-infected ticks, consistent with the results observed by needle inoculation of mice in which only a partial complementation was achieved with cheA1com (Table 1). No flaB transcript was detected in the specimens from mice fed upon by cheA1mut-infected ticks (Table 4), indicating that the mutant failed to be transmitted to mice even though it could colonize the tick vector. Collectively, these results suggest that CheA1 is essential for the transmission of B. burgdorferi to mice via tick bite.

cheA1mut can be acquired by naïve tick from infected mice

The tick transmission experiment revealed that cheA1mut was able to survive in ticks via artificial infection but failed to transmit to naïve mammalian hosts (Fig 8A and Table 4). To test the possible role of CheA1 in the acquisition phase, mice were infected with WT or cheA1mut via needle inoculation. Two weeks post-injection, infection was confirmed via culturing of the skin biopsy from the injection site. Nymphs were placed and confined to the injection site and allowed to feed for 72 hours. Fed ticks were then collected for qPCR to determine the spirochete acquisition efficiency. As shown in Fig 8B, the flaB gene was successfully detected in nymphs that parasitized mutant-infected mice and its copy number was comparable to that detected in nymphs that fed on WT-infected mice (Fig 8C). It should be noted that in the majority of cheA1mut-infected ticks, acquisition was greatly attenuated relative to wild type-infected nymphs with 3 out of 12 nymphs showing no detectable flaB gene. This result suggests that CheA1 is not absolutely required for spirochete acquisition by ticks from mice, but may play some role in the process.

CheA1 affects expression of the RpoS-regulon and RpoS stability

B. burgdorferi encodes several proteins that are essential for host tissue binding/interaction and immune evasion in mammalian hosts, e.g., OspC, BBK32, and DbpB/A [4245]. The attenuated phenotype of cheA1mut observed in the mouse infection studies could be the result of altered expression of virulence-associated proteins. To test this possibility, B. burgdorferi cells were cultivated under elevated temperature and reduced pH (37°C/pH 6.8) to mimic the host environment. Total protein profiles of WT, cheA1mut and cheA1com were analyzed on SDS-PAGE (Fig 9A) and a number of B. burgdorferi virulence determinants were measured by immunoblotting analysis and compared to the level of DnaK, a loading control. The protein level of RpoS was significantly reduced in the cheA1mut mutant as compared to the WT and cheA1com strains (Fig 9B). Consistent with this finding, the levels of several RpoS-regulated proteins (i.e., BBK32, OspC, and DbpA) were also significantly reduced in the mutant and restored in the complemented strain. By contrast, the level of P66, a non-RpoS regulated protein [46], was not affected by the deletion of cheA1, suggesting that CheA1 specifically affects RpoS and RpoS-regulon expression (Fig 9B). qRT-PCR analysis revealed that the transcript level of rpoS was slightly reduced in the mutant but not significant while the transcript levels of ospC and dbpA were significantly reduced by 65% and 75%, respectively (Fig 9C). In contrast, no significant change was observed in the mutant for ospA transcript, a RpoD-regulated gene [47]. In sum, deletion of cheA1 leads to reduced expression of RpoS and RpoS-regulated virulence factors, which in turn could explain the attenuated infectivity observed with cheA1mut. Depletion of CheA1 affects RpoS post-transcriptionally as no significant reduction in rpoS transcript was detected in cheA1mut (Fig 9C). To determine how CheA1 affects RpoS protein expression, protein turnover assays were performed using WT and mutant cells cultivated at the condition mimicking the host environment to induce RpoS expression. Compared to the wild type, RpoS turnover was increased in the absence of CheA1. The level of RpoS was relatively stable in the stationary phase of the wild-type culture whereas over 50% turnover of RpoS was observed in cheA1mut as early as one hour after protein synthesis was arrested (Fig 9D). Complementation of CheA1 successfully restored the stability of RpoS (S7 Fig). Thus, the reduced RpoS protein level observed in cheA1mut is due to reduced protein stability in the absence of CheA1.

thumbnail
Fig 9. Depletion of CheA1 negatively affects the level of RpoS, BBK32, OspC, and DbpA.

B. burgdorferi strains (WT, cheA1mut, and cheA1com) were cultivated at 37°C/pH 6.8 to mimic the mammalian host condition and harvested at stationary phase (∼108 cells/ml) for (A,B) immunoblotting analysis, and (C) qRT-PCR analysis. (A) Similar amounts of whole-cell lysates were analyzed by stain-free SDS-PAGE. Gel image showed the total protein of WT, cheA1mut, and cheA1com analyzed on SDS-PAGE. (B) Immunoblot analysis of WT, cheA1mut, and cheA1com samples using specific antibodies against CheA2, CheA1 [11], P66 [101], BBK32 [102], RpoS, OspC, and DbpA [86]. DnaK was used as an internal control, as previously described [97]. (C) Detection of rpoS, ospC, dbpA, and ospA transcript level in WT, cheA1mut, and cheA1com using qRT-PCR. The dnaK transcript was used as an internal control. Experiments were repeated at least twice with independent replicates. Data from two replicates are expressed as mean fold change relative to WT ± SEM. *, significant difference (P < 0.05). (D) RpoS protein turnover in WT and cheA1mut. The stability of RpoS protein was examined in stationary phase WT B31 A3-68 and B31 A3-68 cheA1mut upon protein synthesis arrest using spectinomycin. Samples were harvested at the indicated time points and probed using antibody against RpoS and DnaK (as a loading control). Experiments were repeated at least three times with independent biological replicates. A representative image is shown here.

https://doi.org/10.1371/journal.ppat.1011752.g009

Depletion of CheA1 leads to dysregulated expression of proteases

The homeostasis and stability of proteins are governed by a set of proteases responsible for the regulated degradation of a given protein target [48,49]. The genome of B. burgdorferi encodes several proteases, including Lon-1 [50], Lon-2 [51], HtrA [5255], CtpA [56], and Clp [9]. Lon-1 protease was shown to negatively affect RpoS expression at the transcriptional level [50] while Lon-2 protease positively regulates RpoS expression at the post-transcriptional level [51]. In other bacteria, such as E. coli, the protein stability of RpoS is governed by the ClpXP protease [5759]. Since the stability of RpoS protein was affected in cheA1mut, the expression of these known proteases was examined in the mutant. qRT-PCR data from two independent biological replicates consistently showed that the transcripts of clpX, lon-1, and lon-2 were significantly reduced in cheA1mut as compared to the WT, which was successfully restored in the complemented strain (Fig 10A). Preliminary RNA-seq analysis also revealed a significant down-regulation of clpX, lon-1 and lon-2 expression in cheA1mut (S8 Fig). Our data suggest that depletion of CheA1 leads to dysregulated expression of B. burgdorferi proteases, such as clpX and lon, which subsequently affects the expression and protein stability of RpoS in a negative manner. In congruence with this, deletion of clpX significantly impaired the expression of RpoS along with the various RpoS-regulated proteins (Fig 10B) while deletion of lon-1 led to a high level of RpoS expression as shown by Thompson et al. [50], further supporting the regulatory role of ClpX and Lon proteases on RpoS expression.

thumbnail
Fig 10. Depletion of CheA1 leads to dysregulated expression of proteases.

(A) Detection of clpX, lon-1, and lon-2 transcript level in WT, cheA1mut, and cheA1com using qRT-PCR. The dnaK transcript was used as an internal control. Experiment was repeated with two independent replicates. Data are expressed as mean fold change relative to WT ± SEM. *, significant difference (P < 0.05). (B) Immunoblot analysis of WT, clpXmut, and lon-1mut samples using specific antibodies against P66 [101], BBK32 [102], RpoS, OspC, and DbpA [86]. DnaK was used as an internal control, as previously described [97]. Experiments were repeated twice with independent replicates. Lon-1 mutant (lon-1mut) was kindly provided by the Ouyang’s lab as a positive control to show the up-regulation of RpoS as previously described [50].

https://doi.org/10.1371/journal.ppat.1011752.g010

cheA1mut fails to elicit mouse il-10, tnf-α, ccl2, and il-1β expression

The attenuated phenotype of cheA1mut in establishing systemic infection even in immunodeficient mice implicates a defect in both innate and adaptive immune evasion. Interleukin 10 (IL-10) is a key mediator in innate immune responses during Lyme disease infection [60]. Timely induction of IL-10 expression is critical for the disease progression by dampening the host inflammatory response to favor the survival of B. burgdorferi during early infection. To dissect the cause behind cheA1mut failure to establish systemic infection in mice, RNA-seq analysis was performed using infected mouse skin tissues. Briefly, 100 μl of 1 x 107 WT or cheA1mut were needle inoculated into the dorsal side of BALB/c mice at three separate locations. Skin tissues were harvested from the inoculation sites at 72-hour post-injection (p.i.) for RNA extraction and RNA-seq analysis. Scatter plot of enriched KEGG pathway showed that distinct host genes were differentially induced between Sham vs WT and Sham vs MT (cheA1mut) infected samples at 72 hrs p.i. (Fig 11A). In WT-infected samples, significant changes of numerous immune system related genes were observed, such as genes involved in leukocyte, granulocyte, and neutrophil migration. These genes were unchanged in the mutant infected samples. The transcript levels of several important cytokines and chemokines that have been reported to contribute to Lyme disease progression, such as IL-10, IL-1β, TNF-α, and CCL2 [6168], were upregulated in the WT- but not MT-infected tissues (Fig 11B), which was validated by qRT-PCR (Fig 11C). Consistent with the RNA-seq analysis, qRT-PCR showed that the transcript levels of il-10, il-1β, and ccl-2 were significantly lower in the cheA1mut-infected mouse skin tissues as compared to those infected by WT (Fig 11C). The data obtained from RNA-seq analysis suggests that the inability of cheA1mut to establish systemic infection is partly due to its failure in eliciting timely expression of key host cytokines and chemokines to aid in dissemination and colonization.

thumbnail
Fig 11. RNA-seq analysis of BALB/c mouse skin tissues infected by WT and cheA1mut.

(A) Scatter plot of enriched KEGG pathway of differentially expressed genes between Sham and infected samples at 72hr p.i. Vertical coordinates represent pathway name, and horizontal coordinates represent Rich factor. The size and color of point represent the number of differential genes in the pathway and the range of different Q value, respectively. (B) Volcano plot of differentially expressed genes (DEG) between MT (cheA1mut) vs WT samples at 72hr p.i.. Each dot represents a gene, and its position is determined by the log2 fold change (x-axis) and the -log10 (p-value) (y-axis). Genes with significant upregulation (red) or downregulation (green) are highlighted, while non-significant genes are shown in gray. (C) Detection of il-10, tnf, il-1β, and ccl-2 transcript levels in WT and cheA1mut infected mouse skin tissues using qRT-PCR. The β-actin transcript was used as an internal control. Data are expressed as mean fold change relative to WT ± SEM. *, significant difference (P < 0.05).

https://doi.org/10.1371/journal.ppat.1011752.g011

Discussion

B. burgdorferi lives a complex enzootic life cycle alternating between arthropod vector and mammalian reservoir. As such, highly coordinated expressions of tick- and host-specific transcriptomes are indispensable to ensure the survival and persistence of the spirochete in nature. Two two-component systems (TCS), the Hk2-Rrp2-RpoN-RpoS and HK1-Rrp1 signaling pathways, are largely responsible for these extensive transcriptional adjustments (see reviews [6971]). Among the enzootic cycle stage-specific differentially expressed genes, in addition to multiple crucial surface lipoproteins, many chemotaxis-related genes are also found to have distinct expression profiles when B. burgdorferi alternates between different hosts, which strongly indicates that the composition of the B. burgdorferi chemotaxis machinery differs between ticks and mammals [72]. The genome of B. burgdorferi encodes for orthologs of chemotaxis-related genes, many of which are present in several copies, such as two chemotaxis histidine kinases (cheA1, cheA2) and three chemotaxis response regulators (cheY1, cheY2, cheY3) [9]. Some of these genes have no known function in chemotaxis as mutagenesis studies revealed no apparent contribution to chemotaxis in vitro. For example, a mutant deficient in CheA1 exhibited a normal chemotaxis response whereas loss of CheA2 resulted in a non-chemotactic phenotype in vitro [11,17]. Hence, understanding how these chemotaxis-related proteins are coordinated during the natural enzootic cycle of B. burgdorferi is crucial to fully comprehend the significance of each protein in the modulation of chemotaxis responses in vivo.

Interestingly, even though CheA1 is not required for chemotaxis in vitro, it has a domain arrangement and a local domain architecture more similar to TmCheA than the chemotactic CheA2. AlphaFold models reveal that CheA1 has a single P2 domain, while CheA2 has two P2 domains (α and β, Fig 1A–1B). Sequence alignment, structural superimposition, and previous biochemical data suggest that these domains all bind to CheY-like proteins [13] (S1A–S1B Fig), although whether different P2 domains are specific for one CheY isoform over another (i.e. CheY1-3) is not currently known. In addition to the P2 domain, the other major difference between CheA1 and CheA2 is the length of the P3 domains (Fig 1D). CheA1 has a P3 domain very similar in length to TmCheA, whereas in CheA2 the P3 is over 68 residues and 43 Å longer than CheA1 and TmCheA. Previously, it has been recognized that B. burgdorferi and Treponema denticola have extended P3 domain structures [21,24]. It was hypothesized that these extended P3 domains may stabilize the CheA2 homodimer by increasing the binding interface surface area of the two CheA subunits. Another possibility could be that the longer P3-P3’ four-helix bundle is an adaptation to allow better engagement with the chemosensory arrays of spirochetes, which have distinct architectures owing to the high membrane curvature [21,24]. Although the exact molecular mechanism behind why Borreliaceae CheA2 encodes for a second P2 domain and an extended P3 domain has yet to be determined, it is nonetheless interesting that these two features are conserved across the family of Borreliaceae (S3 Fig).

Motility and chemotaxis contribute significantly to the pathogenesis of B. burgdorferi as loss of motility or chemotaxis resulted in non-infectious spirochetes that failed to colonize tick vectors and/or mammalian hosts [17,73,74]. A previous study from our group established that chemotaxis-signaling mediated by the CheA2 histidine kinase is required for the transmission of spirochetes from ticks to mammals, as well as for establishing an infection in a vertebrate host, but is not essential for its survival in ticks [17]. Based on the organization of the chemotaxis gene clusters [75] as well as mutagenesis studies of genes in the cheA1 [11,1416] and cheA2 [12,13,17,73] operons, we hypothesized that B. burgdorferi may utilize different chemosensory machineries in different environments during its infectious cycle, i.e., CheA2 modulates chemotaxis during transmission/persistence in mammalian hosts while CheA1 constitutes a second chemotaxis pathway to guide the chemotactic response in the tick vector and accordingly mediate B. burgdorferi transmission from infected ticks to mammals. In line with our hypothesis, a recent study by Iyer et al. showed that CheA1 and CheY2 are highly induced in transmitting ticks which indicates that these two proteins may form a functional chemotaxis pathway that is activated when an infected tick takes a blood meal so as to prepare dormant spirochetes for migration and transmission to a new host [72]. Lastly, a recent study by Caimano et al. further substantiated the role of CheA1 during transmission whereby the level of cheA1 is highly induced by RpoS while in fed ticks and mammals [76].

One interesting observation from the current study revealed that CheA1 has a dynamic localization during growth. Importantly, the intense single pole localization of CheA1 when cultivated under condition mimicking the unfed tick milieu versus elevated temperature indicates that CheA1 may play a more prominent role in the arthropod vector as compared to the vertebrate host (Figs 3 and S6). CheA1 may be strategically placed near the cell pole in proximity to the chemoreceptor arrays [16], ready to detect stimuli from the blood meal in order to initiate transmigration out from the tick midgut to the salivary glands for transmission [77,78]. To answer this question, in vivo visualization of cheA1mut cells after the tick blood meal is necessary to reveal whether the mutant’s inability to be transmitted via tick bite (Table 4) is due to failure in migration from the hemolymph to the tick salivary glands. Additionally, the increase in puncta localization of CheA1-GFP as the cells transition from early to late log phase has a striking resemblance to ParB localization. ParB is a chromosome segregation protein that has been shown to recruit the structural maintenance of the chromosome (SMC) complex to the origin of replication in B. burgdorferi [79]. It would be interesting to test if the localization of CheA1 is associated with the DNA partitioning Par proteins perhaps to ensure equal segregation of CheA1 protein among the daughter cells during cell division.

Given that the cheW2 operon is positively regulated by RpoS and BosR [18,76], the failure of cheA1mut during transmission could be partly due to defect in transmigration from the tick gut to the salivary glands during the blood meal, similar to the phenotype of a rpoS mutant in the feeding tick [80]. Most significantly, the possible association of CheA1 with the Par proteins implies that the established pool of CheA1 needs to be distributed equally among the daughter cells. This unique partitioning was also observed in cells cultured at elevated temperature but only during the late stationary phase (S6B Fig). Once these late stationary phase cells were re-inoculated into fresh BSK-II media, CheA1 localization reverted to the diffused pattern as seen during day 1. This protein localization is specific to CheA1 as we did not observe this with CheA2 (S5D and S6C Fig). The need for CheA1 to have spatial and temporal localization is currently unknown. Given that the tick-mouse infection study points toward a role for CheA1 during tick transmission and hematogenous dissemination, it would be interesting to monitor the localization of CheA1 in feeding ticks and in mammals to determine if CheA1 alters its localization in different cellular milieus and whether such localization contributes to its function.

Following initial deposition, either via tick bite or needle inoculation, B. burgdorferi needs to first establish a local infection before disseminating to peripheral organs. Therefore, hematogenous dissemination is a critical pathogenic event for early infection [81]. Once in the bloodstream, the spirochetes must transmigrate via vascular extravasation to spread and colonize distal organs [82,83]. Several findings elucidate the dissemination event in which CheA1 is most likely to be involved. First, the cheA1 mutant is able to survive at the injection site of immune-competent and–deficient mice after needle inoculation, indicating that CheA1 is not required for skin tissue survival. Needle inoculation into SCID mice allowed partial dissemination of the cheA1 mutant to the bloodstream and subsequent colonization of skin regions (i.e., ears) distal from the injection site as well as other organs, but failed to do so in immunocompetent mice. This result indicates that CheA1 is required for the spirochete survival in the bloodstream most likely by aiding the bacteria in immune evasion. In line with this, immunoblot analysis revealed that deletion of CheA1 negatively impacted the expression of several essential surface virulence factors, such as OspC and BBK32 (Fig 9B). OspC has been shown to interact with host immune components, such as the complement component 4b (C4b), and contributes to the survival of B. burgdorferi in the bloodstream [42]. This can explain the finding in SCID mice where only 50% of the mice had systemic infection as the complement system remains intact in the immune-deficient mice and cheA1mut is not as protected in the bloodstream as compared to the wild type (Table 2 and Fig 6). This was corroborated by serum bactericidal assays, as deletion of CheA1 renders the spirochetes more susceptible to serum killing (Fig 7B). B. burgdorferi infection by tail-vein further supported this proposition, as the mutant was cleared from the bloodstream of wild-type mice within 14 days and was severely attenuated in hematogenous dissemination. Tail-vein injection with high doses of spirochetes was not able to rescue the survival and systemic dissemination defect of cheA1 mutant indicating that in addition to attenuated innate and adaptive immune evasion, the mutant may have a defect in tissue colonization or an attenuated ability to extravasate due to dysregulated surface protein expression (Fig 9), such as BBK32 [84], which led to the clearance of spirochetes in the bloodstream by the host immune system. The ability of cheA1mut to transmigrate from the vasculature will need to be examined via intravital imaging in order to determine if the reduced ability of cheA1mut to colonize and disseminate from the bloodstream of mice is partially due to a defect in extravasation in addition to impaired immune evasion. Artificial infection of Ixodes ticks showed no significant variation in the survival between wild type and mutant (Fig 8A). Furthermore, the cheA1 mutant can be acquired from the skin of infected mice suggesting that CheA1 is dispensable for acquisition. In addition, cheA1mut failed to be transmitted via infected tick bites, suggesting that CheA1 is required for tick transmission, consistent with the observed increase in cheA1 transcripts during transmission [72]. Immunoblot and qRT-PCR analyses (Fig 9) suggest that the attenuated infectivity in mice upon depletion of CheA1 is likely due to dysregulated RpoS and RpoS-regulon expression, which are essential for B. burgdorferi survival, immune evasion, and persistence in mammalian hosts [42,43,45,69,78,85,86]. Protein turnover assays revealed that the stability of RpoS was significantly reduced in the absence of CheA1 (Fig 9D) and that this is likely due to dysregulated protease expression (Fig 10).

In other bacteria such as E. coli and Bacillus subtilis, in addition to the σ70 house keeping sigma factor, , multiple alternative sigma factors are employed to effectively regulate gene expression based on changes in the growth condition, such as stress response, starvation, and transition between motile and sessile stages [87,88]. The activity of each σ is controlled by the binding of anti-σ proteins such that the proper σ can be released for interacting with the RNA polymerase (RNAP) when required [8991]. The genome of B. burgdorferi encodes two alternative sigma factors, σ54 (RpoN) and σ38 (RpoS) [9]. How sigma factor partner switching occurs in B. burgdorferi remains unclear as no anti-sigma factors have yet been discovered. Preliminary RNA-seq analysis performed using in vitro cultured samples revealed significant up-regulation of many RpoS-repressed genes [76], such as bb_d18 and bb_a68 [76,92] in the cheA1mut mutant (S7 Fig). As shown in Fig 10A, the level of bb_d18 was significantly higher in cheA1mut as compared to WT when validated using qRT-PCR. Notably, the expression of bb_d18 remained elevated in the complemented strain cheA1com, which may explain why only a partial complementation was achieved in cheA1com (Table 1). The untimely expression of RpoS-repressed genes may have diluted the pool of RNAP-RpoD complex available for transcription of core genes, such as clpX and lon proteases, leading to global dysregulation of the CheA1 mutant transcriptome (S8 Fig). Additionally, BBD18 is known as a negative regulator of RpoS [9395] and upregulation of this gene can further repress the expression of RpoS-regulon in the mutant (Fig 9). In sum, the data strongly suggest that CheA1 affects the pathogenicity and transcriptome of B. burgdorferi by influencing the stability of RpoS potentially through ClpX and Lon proteases.

As discussed above, CheA1 may contribute to chemotaxis within the tick. It is also possible that it has a role to play in host dissemination. We did not detect any chemotaxis defect in the cheA1mut in vitro (Fig 4); however, the significance of CheA1 in chemotaxis may only be observable in vivo as lab cultivation rarely accurately represents any aspect of the natural infectious cycle of B. burgdorferi [96]. We had previously shown that CheW2 contributes to the full pathogenicity of B. burgdorferi as depletion of CheW2 resulted in attenuated dissemination and colonization of the spirochetes [18]. Given that CheW2 preferentially interacts with CheA1 [16], CheA1, CheW2, and CheY2 may constitute an active chemotaxis pathway during the vector transmission and host dissemination phases to guide the chemotactic responses and migration of B. burgdorferi. Notably, the phenotype of cheA1mut differs from what was observed in an infectious cheY2 and cheW2 mutant; Xu et al. showed that cheY2 mutant was cleared from the inoculation site after two weeks and failed to cause systemic infection [15]. In contrast, cheA1mut persisted at the initial injection site and was able to replicate as shown by our time course study (Fig 5). As for CheW2, even though deletion of CheW2 led to attenuated infectivity in immune-competent mice, its infectivity was not affected in SCID mice, which differs significantly from the in vivo phenotype of cheA1mut [18]. However, all three mutants exhibited wild-type chemotaxis behavior in vitro which raises the question as to whether CheA1-CheW2-CheY2 is required for chemotaxis under specific circumstances in vivo. cheA1 and cheY2 mutant survival in the tick vector was not significantly different from the wild type (Fig 8A), emphasizing that CheA1 and CheY2 [15] are not required for the persistence and survival of B. burgdorferi in the arthropod vector. How is cheA1mut able to persist and spread in the skin tissue but not to distal organs? Given that B. burgdorferi can spread through direct tissue spreading [81], it is possible that once cheA1mut establishes localized infection at the injection site, it is able to spread within the skin tissue to distal skin regions (e.g., ears) and avoid clearance by the host immune system via non-hematogenous spreading.

In order to colonize and survive in a mammalian host, B. burgdorferi has evolved to alter its surface antigen expression for distinct tissue tropism as well as in response to host immune reaction [45]. The disparity in host immune repones elicited by WT vs cheA1mut (Fig 11) is in congruence with the mouse infection results. The lack of key inflammatory mediators necessary for spirochete dissemination may dampen the progression of cheA1mut infection and dissemination outcome, leading to attenuated infectivity as observed (Tables 1, 2 and 3). The anti-inflammatory cytokine IL-10 plays a key role in the progression of B. burgdorferi infection [6062]. The failure of cheA1mut to elicit timely expression of IL-10 in mice may have contributed to its failure in dissemination and colonization in both immune-competent and–deficient mice (Tables 1 and 2). Despite this, the mutant was able to persist at the inoculation site (Fig 5) without the protection of an IL-10 cytokine effect which implies that additional bacterial and host factors may be required. Lack of an activated host immune reaction following cheA1mut infection (Fig 11B) could potentially allow the spirochetes to evade immune surveillance and persist at the inoculation site.

In summary, this study reveals that CheA1 is a non-canonical chemotaxis histidine kinase with a novel role in B. burgdorferi infectivity and control of virulence factor expression. As a histidine kinase, CheA1 does not directly influence gene expression but must act through a response regulatory protein in order to activate a given signaling pathway. Identifying the cognate response regulator for CheA1 is the next essential step in comprehending how CheA1 influences global transcriptomic changes via RpoS protein expression and stability as well as the greater roles that CheA1 plays in B. burgdorferi pathogenesis in vivo.

Material and methods

Ethics statement

All animal experimentation was conducted following the NIH guidelines for housing and care of laboratory animals and performed in accordance with the Virginia Commonwealth University institutional regulations after review and approval by the Institutional Animal Care and Use Committees [IACUC approval number: #AD10001779].

Bacterial strains and growth conditions

Infectious clone A3-68 and A3-68∆bbe02 (wild type), two derivative strains from the B. burgdorferi sensu stricto B31-A3, were used in this study [25]. They were a kind gift from P. Rosa (Rocky Mountain Laboratories, NIAID, NIH). Cells were grown in Barbour-Stoenner-Kelly (BSK-II) medium as previously described [97] in appropriate antibiotic(s) for selective pressure as needed: streptomycin (50 μg/ml), kanamycin (300 μg/ml), and/or gentamicin (40 μg/ml). To determine the expression of CheA1, 105 stationary phase wild type cells were inoculated into 10 ml of fresh BSK-II medium and then cultured at 23°C/pH 7.6 (unfed tick, UF) or 34°C/pH 7.6 (a routine laboratory cultural condition). To examine the expression of RpoS and RpoS-regulon, 105 stationary phase wild type, mutants, and complemented strains were inoculated into 10 ml of fresh BSK-II medium and then cultivated at 37°C/pH 6.8 with 5% CO2 to mimic the host condition. Cells were harvested for immunoblot analysis upon entry into stationary phase at ~108 cells/ml.

CheA1, CheA2, and TmCheA AlphaFold model generation and analysis

To build CheA1, CheA2, and TmCheA models, protein sequences for TmCheA (AAA96387.1), CheA1 (AAC66935.1), and CheA2 (AAC67024.1) were retrieved from the NCBI database and submitted for automated model-building using AlphaFold2 [20]. All parameters were kept at their default values for model building.

Construction of CheA1/pBBE22 for complementation of cheA1 mutant

pGA1kan, a previously constructed vector with an internal 393-bp deletion in cheA1 gene (nucleotides 1,016–1,409) [11], was used to inactivate the cheA1 in B31 A3-68 and A3-68∆bbe02 strains via allelic exchange mutagenesis. All the experiments performed in this report was done using mutants generated in the A3-68∆bbe02 background except for RpoS protein turnover analysis in which the cheA1 mutant was generated in the B31 A3-68 background because A3-68∆bbe02 carries a streptomycin resistant cassette. To complement the cheA1 mutant, CheA1/pBBE22 (Fig 2A), was constructed via PCR ligation using primer pair P1/P2 to amplify bb0566-cheA1 region and cloned into pBBE22G vector at BamHI restriction site [17,98,99]. All primers used in this study are listed in Table 5.

Construction of CheA1IFD for in-frame deletion of cheA1

To create an alternative cheA1 mutant, an in-frame deletion construct was made using PCR ligation method to in-frame replace the entire open reading frame (orf) of cheA1 with kanamycin (kan) cassette. The left and right arm of the deletion construct was PCR amplified using P27/P28 and P29/P30, respectively, and the kan cassette was amplified using P31/P32 (Table 5). The resulting three amplicons were then PCR ligated using P27/P30 and ligated to pJET1.2/blunt cloning vector (Thermo Scientific, Waltham, MA) yielding CheA1IFD (S4B Fig).

Localization of CheA1 and CheA2 in B. burgdorferi

For the localization of CheA1 and CheA2, cheA1-gfp/pBSV2G (S3A Fig) and cheA2-gfp/pBSV2G (S3B Fig) constructs were used to complement cheA1mut and a previously constructed cheA2 mutant strain [11], respectively. A control plasmid with gfp driven by the flgB promoter was included as a control for the localization study [100]. To determine the localization of CheA1 and CheA2, GFP construct expressing cells were inoculated into 10 ml of fresh BSK-II medium and then cultured at 23°C/pH 7.6 or 34°C/pH 7.6. Cells were observed every 48 hours for up to 2 weeks. Images were taken using a Zeiss Axiostar plus microscope and processed using Axiovision software (Zeiss, Germany) as previously described [100].

SDS-PAGE and immunoblots

10 to 20 μg of B. burgdorferi whole cell lysates were separated on 10 or 12% Stain-Free SDS-PAGE gel and transferred to PVDF membrane (Bio-Rad Laboratories, Hercules, CA). Stain-Free SDS-PAGE was first activated and imaged using ChemiDoc MP Imaging System (Bio-Rad Laboratories) prior to membrane transferring to visualize total proteins on the gel. The immunoblots were probed with antibodies against B. burgdorferi CheA1, CheA2 [11], P66 [101], BBK32 [102], RpoS, OspC, and DbpA [86]. DnaK was used as an internal control, as previously described [97]. Membranes were developed using horseradish peroxidase secondary antibody with an ECL luminol assay or with fluorescently labelled secondary antibodies. Signals were imaged using the ChemiDoc MP Imaging System and quantified with the Image Lab software (Bio-Rad Laboratories).

Bacterial motion tracking analysis, swimming plate, and capillary tube-based chemotaxis assays

The swimming velocity of B. burgdorferi cells was measured using a computer-based motion tracking system as previously described [31]. Swimming plate assays were conducted as previously documented [11,32,103]. The diameters of swimming rings were recorded in millimeters. Wild-type A3-68 Δbbe02 was used as a positive control. A non-motile strain, flgE mutant, [103] was used as a negative control to monitor the initial inoculum size. For the capillary tube assay, N-acetyl-D-glucosamine (NAG) was used as an attractant as previously reported [31]. The spirochete cells accumulated in the capillary tubes were enumerated using Petroff-Hausser counting chambers. The mean of three replicates was determined and the data are expressed as the mean relative increase over a buffer control containing no attractant for each group. A two- or more fold increase in the number of spirochetes in comparison to buffer control is considered as significant.

Quantitative reverse transcription PCR (qRT-PCR)

Ticks and mouse tissue samples for qRT-PCR analysis were prepared as described [17,97]. Briefly, total RNA from ticks or mouse tissues was isolated using TRIzol reagent (Invitrogen, Carlsbad, CA) and contaminating genomic DNA was removed using Turbo DNase (Ambion, Austin, TX). The DNase-treated RNAs were re-purified and converted to cDNA using SuperScript IV VILO Master Mix (Invitrogen) according to the manufacturer’s instructions. Quantitative PCR was then performed using Fast SYBR Green Master Mix (Applied Biosystems, Foster City, CA). The spirochete burdens within infected mice and ticks were expressed as the flaB transcript levels relative to the copy number of the mouse or tick β-actin transcript levels as described previously [17]. Primers used in qRT-PCR are summarized in Table 5. Data are expressed as the mean fold change of two replicates relative to WT ± SEM. The significance of the difference between different experimental groups was evaluated with ANOVA (P value < 0.05).

Protein turnover assay

To determine the stability of RpoS protein, protein turnover was performed as described previously [104] using wild-type A3-68 and a cheA1 mutant constructed in the same background. The mutant was confirmed to have the same antigen profile as cheA1mut [103,104]. Briefly 105 cells/ml of B. burgdorferi cells were inoculated into 50 ml of BSK-II medium, pH 6.8, and cultivated at 37°C to mid log phase (5x106 cells/ml) or stationary phase (108 cells/ml). 100 μg/ml of spectinomycin was added to the culture to arrest protein synthesis. 5 ml of cells were harvested at 0, 1, 2, 3, 4, 6, 8, and 24 hours post-protein arrest for SDS-PAGE followed by immunoblots against RpoS and DnaK.

Mouse infection studies

BALB/c or BALB/c SCID mice at 6–8 weeks of age (Jackson Laboratory, Bar Harbor, MN) were used in the needle infection study. The animal studies were carried out as previously described [17,97]. Briefly, mice were given a single subcutaneous injection of 105 spirochetes and sacrificed 3 weeks post-infection. For time point inoculation site study, mice were similarly infected with WT or cheA1mut and the initial inoculation sites were marked for harvesting at 96 hour, 14 day, 21 day, and 28 day post-infection. Skin specimens were subjected to qRT-PCR analysis as described above.

For tail-vein infection, 106 spirochetes re-suspended in 100 μl PBS was injected into the tail-vein of BALB/c mice. The non-motile non-infectious flaB mutant was included as a negative control [32]. Mice were sacrificed two weeks after infection. Tissues from the ear, skin, joint, and heart were harvested and placed into 1 ml BSK-II medium containing 5 μg/ml of rifampicin. The samples were incubated at 34°C for up to one month and microscopically monitored for the presence of spirochetes. Tissues were similarly collected for qRT-PCR analysis to assess for spirochete burdens. For bulk RNA-seq analysis, 107 spirochetes in 100 μl of BSK-II was injected into the dorsal side of BALB/c mice at three separate locations. For sham control, equal volume of BSK-II medium was injected instead. Mice were sacrificed at 72-hour post-injection. Skin tissues from the injection sites were harvested for RNA extraction as described above.

Serum bactericidal assays

Serum bactericidal assay was performed as previously described [43]. Normal human serum was purchased from BioIVT (Westbury, NY). (Briefly, wild-type A3-68∆bbe02, cheA1mut and cheA1com were grown to mid log phase at 34°C/pH 7.6. 106 cells in 80 μl of BSK-II medium were inoculated into a 96-well plate followed by the addition of 20 μl of normal human serum (NHS) or heat inactivated NHS (hiNHS) to get a final concentration of 20% serum in 100 μl. Cells were incubated at 34°C for 2 h. The percentage of viable cells after incubation were enumerated under dark field microscope using Petroff-Hausser counting chambers based on immobilization, cell lysis, and loss of cell membrane integrity. Experiments were repeated at least three times. Data is presented as percent viable cells relative to input cells.

Tick-mouse transmission study

The tick-mouse transmission study was carried out as previously described [17,41]. Briefly, naïve Ixodes scapularis nymph (Oklahoma State University, Stillwater, OK) were artificially infected by microinjection as previously described [39]. Two days after injection, infected nymphs were allowed to parasitize on naïve C3H mice (5 ticks/mouse and 3 mice for each B. burgdorferi strain) for 5 to 7 days and allowed to fall off. Engorged ticks were collected and subjected to qRT-PCR analysis to determine spirochete burdens. At day 14 after the tick feeding, mice were sacrificed and tissues from the skin, heart, joint, bladder and blood were harvested for quantification of spirochete burden via qRT-PCR analysis to determine transmission rate as described previously [41]. Spirochetal burden in fed ticks is presented as log copies of flaB / 103 tick β-actin copies while transmission data is shown as flaB copies per 106 mouse β-actin.

Tick acquisition studies

To examine the ability of cheA1mut to be acquired by feeding ticks, 1 x 104 of either WT or cheA1mut was injected subcutaneously into two naïve C3H mice each. Two weeks post-infection, skin biopsy was obtained from the site of injection to confirm present of live spirochetes via culturing. 2–3 naïve I. scapularis nymphs were placed onto the infected mice at the site of inoculation and allowed to feed to repletion. After 72 hours, fed nymphs were collected and tested individually for the presence of flaB via qPCR. Tick acquisition experiments were repeated twice and the mean of both data sets are presented as flaB copies per fed nymph.

Construction of an in-frame deletion mutant of clpX

An in-frame deletion construct of clpX was made using the same PCR ligation method as described for CheA1IFD. The orf of clpX was replaced with kan cassette. The left and right arm of the deletion construct was PCR amplified using P23/P24 and P25/P26, respectively and kan cassette was amplified using P31/P32 (Table 5). The resulting three amplicons were then PCR ligated using P39/P42 and ligated to pJET1.2/blunt cloning vector (Thermo Scientific) and used for transformation into wild-type A3-68 strain.

Mouse RNA-seq analysis

Total RNAs were extracted from mouse skin tissues using TRIzol reagent as described and treated with DNase to remove contaminating genomic DNA. RNA samples were sent to CD Genomics (CD Genomics, Shirley, NY) for RNA-seq analysis. Briefly, rRNA was depleted using Illumina Ribo-Zero Plus rRNA Depletion Kit. Library preparation was performed with NEBNext Ultra II RNA (non-directional) Kit. Paired-end reads with a length of ~150 nt were generated with NovaSeq 6000 platform. Quality control was performed with Fastp with low-quality reads filtered. Reads were aligned to the host reference genome (GRCm39). Quantification was then performed based on the alignment results, followed by differentially expressed gene analyses with DESeq2. Significant differentially expressed genes (DEG) between pairwise comparisons were identified with threshold |log2FC| > 0.5, and p < 0.01. All the significant DEGs were annotated with eggNOG, COG, Pfam, Swiss-Port, and KEGG databases. GO enrichment analysis and KEGG pathway analysis were performed with ClusterProfiler in R.

Statistical analysis

For the swimming plates, motion tracking, capillary tube assays, and the mouse and tick infection studies, the results are expressed as means ± standard error of the mean (SEM). The significance of the difference between different experimental groups was evaluated with an unpaired Student t test or ANOVA (P value < 0.05). For serum bactericidal assay, statistical significance was determined using multiple t test (P value < 0.05) followed by the Holm–Bonferroni method to correct for multiple comparisons. Data is depicted as the mean of all replicates ± SEM.

Supporting information

S1 Fig. Sequence and structural alignment of the P2 domains from BbCheA1 and BbCheA2.

(A) Sequence alignment of BbCheA2 P2α and P2β domains compared to TmCheA P2 domain. Conserved residues at the CheY:P2 domain binding site interface are indicated with black boxes. (B) Structural superimposition of BbCheA1 P2 (pink) and BbCheA2 P2α (green) domains with TmCheA P2:CheY complex (tan, PDB:1U0S) [5]. Multiple sequence alignment analysis (MSA) were generated using Clustal Omega [6] and figures were prepared in PyMol [7].

https://doi.org/10.1371/journal.ppat.1011752.s001

(TIFF)

S2 Fig. Multiple sequence alignment analysis (MSA) of Treponema spp. that contain a single CheA isoform.

(A) MSA of Treponema spp. CheA P2 domain sequences. BbCheA1 and BbCheA2 sequences are included for comparison. The location of BbCheA2 P2α is marked with a black box. (B) MSA of Treponema spp. CheA P3 domain sequences. BbCheA1 and BbCheA2 sequences are included for comparison. The location of BbCheA2 P3 domain is marked with a black box. Sequences collected using Annotree [1], MSA files generated using Clustal Omega [6].

https://doi.org/10.1371/journal.ppat.1011752.s002

(TIFF)

S3 Fig. Multiple sequence alignment analysis (MSA) of Borrelia spp. and Borreliella spp. CheA1 and CheA2 sequences.

(A) MSA of Borrelia (B_) and Borreliella (Ba_) spp. CheA1 and CheA2 P2 domain sequences. Location of BbCheA P2α marked with black box. (B) MSA of Borrelia (B_) and Borreliella (Ba_) spp. CheA1 and CheA2 P3 domain sequences. Location of BbCheA2 extended P3 domain sequence marking with black box. Sequences collected using Annotree [1], MSA files generated using Clustal Omega [6].

https://doi.org/10.1371/journal.ppat.1011752.s003

(TIFF)

S4 Fig. Detection of plasmid content in the cheA1com strain by PCR and construction of CheA1IFD for the in-frame deletion of cheA1 gene.

PCR was used to detect the plasmid profile of WT (A), cheA1mut (B) and cheA1com (C). The primers used were described previously [8]. (D) A diagram illustrating the construction of CheA1IFD. To construct an in frame deletion mutant, primer pair P27/P28 and P29/P30 were used to amplify the upstream and downstream flanking region of cheA1. Primer pair P31/P32 was used to amplify a promoterless kanamycin cassette (kan). cheA1 was in-frame replaced by kan via PCR fusion technique with primer pair P27/P30. The resulting PCR fusion amplicon was cloned into pJet1.2 vector forming CheA1IFD.

https://doi.org/10.1371/journal.ppat.1011752.s004

(TIFF)

S5 Fig. Construction of cheA1-gfp/pBSV2G and cheA2-gfp/pBSV2G and the expression of CheA GFP fusion proteins.

(A) Construction of cheA1-gfp/pBSV2G. The upstream flanking region (PcheW2, 417 bp) of cheW2 gene was PCR amplified and fused to cheA1 using the engineered restriction sites as shown. The fused PcheW2-cheA1 fragment was then fused to gfp gene containing a 5 x Gly linker at the indicated restriction sites and cloned into pBSV2G, a shuttle vector of B. burgdorferi [3], yielding cheA1-gfp/pBSV2G. (B) Construction of cheA2-gfp/pBSV6G. Similarly, the flgB promoter [4] was PCR amplified and fused to cheA2 gene followed by fusion to gfp with 5 x Gly linker at the indicated restriction sites prior to cloning into pBSV2G. The obtained construct was used to complement a previously constructed cheA2 mutant strain [9]. (C) Immunoblot analysis of GFP fusion proteins. Cell lysates from B. burgdorferi strains expressing GFP, CheA1-GFP, or CheA2-GFP cultured at UF conditions (23°C/pH 7.6) or routine laboratory cultural condition (34°C/pH 7.6) were analyzed on SDS-PAGE and probed with antibodies against GFP or DnaK (as loading control). No excessive degradation of GFP fusion proteins was observed under both culture conditions. (D) CheA2-GFP does not have a polar localization like CheA1-GFP. B. burgdorferi strain carrying CheA2-GFP construct was cultivated under UF tick condition. CheA2-GFP appeared diffused with no specific cellular localization observed.

https://doi.org/10.1371/journal.ppat.1011752.s005

(TIFF)

S6 Fig. CheA1-GFP localization is spatially- and temporally-regulated.

105 cells/ml of B. burgdorferi strains that express (A) GFP, (B) CheA1-GFP, or (C) CheA2-GFP were inoculated into 10 ml fresh BSK-II medium and cultivated at 34°C/pH 7.6. Images were taken every two days at ×200 magnification using a Zeiss Axiostar Plus microscope. Scale bars represent 10 μm.

https://doi.org/10.1371/journal.ppat.1011752.s006

(TIFF)

S7 Fig. Complementation of CheA1 restored RpoS protein stability in cheA1com.

The stability of RpoS protein was examined in stationary phase of B31 A3-68 cheA1com upon protein synthesis arrest with spectinomycin. Samples were harvested at the indicated time points and probed using antibody against RpoS and DnaK (as a loading control).

https://doi.org/10.1371/journal.ppat.1011752.s007

(TIFF)

S8 Fig. Volcano plot of differentially expressed genes (DEG) between WT and cheA1mut.

A total of 1391 DEG between WT and mutant were plotted. Three down-regulated protease genes (bb0613, bb0612, and bb0253) and six up-regulated RpoS-repressed genes (bb_a38, bb_a68, bb_d18, bb_i16, bb_i39, bb_j09) in cheA1mut were highlighted in the volcano plot.

https://doi.org/10.1371/journal.ppat.1011752.s008

(TIFF)

S1 Table. Oligonucleotide primers used in this study.

https://doi.org/10.1371/journal.ppat.1011752.s009

(PDF)

Acknowledgments

Special thank gives to Dr. Utpal Pal for his assistance with tick microinjection study. We thank Eamen Ho and Weigang Qiu of Hunter College of City University of New York for visualization of RNA-Seq data. We also thank Dr. Patricia Rosa for providing B. burgdorferi A3-68 and A3-68∆bbe02 strains, Dr. Jennifer Coburn for providing P66 antibody, Dr. Melissa Caimano for providing OspC and DbpA antibody, Dr. Janakiram Seshu for providing BBK32 antibody, and Dr. Zhiming Ouyang for providing the lon-1 mutant.

References

  1. 1. Wadhams GH, Armitage JP. Making sense of it all: bacterial chemotaxis. Nat Rev Mol Cell Biol. 2004;5(12):1024–37. pmid:15573139
  2. 2. Falke JJ, Bass RB, Butler SL, Chervitz SA, Danielson MA. The two-component signaling pathway of bacterial chemotaxis: a molecular view of signal transduction by receptors, kinases, and adaptation enzymes. Annu Rev Cell Dev Biol. 1997;13:457–512. pmid:9442881
  3. 3. Porter SL, Wadhams GH, Armitage JP. Signal processing in complex chemotaxis pathways. Nat Rev Microbiol. 2011;9(3):153–65. pmid:21283116
  4. 4. Sourjik V, Wingreen NS. Responding to chemical gradients: bacterial chemotaxis. Curr Opin Cell Biol. 2012;24(2):262–8. pmid:22169400
  5. 5. Kugeler KJ, Schwartz AM, Delorey MJ, Mead PS, Hinckley AF. Estimating the Frequency of Lyme Disease Diagnoses, United States, 2010–2018. Emerg Infect Dis. 2021;27(2):616–9. pmid:33496229
  6. 6. Murray TS, Shapiro ED. Lyme disease. Clin Lab Med. 2010;30(1):311–28. pmid:20513553
  7. 7. Stanek G, Wormser GP, Gray J, Strle F. Lyme borreliosis. Lancet. 2012;379(9814):461–73. pmid:21903253
  8. 8. Charon NW, Goldstein SF. Genetics of motility and chemotaxis of a fascinating group of bacteria: the spirochetes. Annu Rev Genet. 2002;36:47–73. pmid:12429686
  9. 9. Fraser CM, Casjens S, Huang WM, Sutton GG, Clayton R, Lathigra R, et al. Genomic sequence of a Lyme disease spirochaete, Borrelia burgdorferi. Nature. 1997;390(6660):580–6. pmid:9403685
  10. 10. Ge Y, Charon NW. Molecular characterization of a flagellar/chemotaxis operon in the spirochete Borrelia burgdorferi. FEMS Microbiol Lett. 1997;153(2):425–31. pmid:9271872
  11. 11. Li C, Bakker RG, Motaleb MA, Sartakova ML, Cabello FC, Charon NW. Asymmetrical flagellar rotation in Borrelia burgdorferi nonchemotactic mutants. Proc Natl Acad Sci U S A. 2002;99(9):6169–74. pmid:11983908
  12. 12. Motaleb MA, Miller MR, Li C, Bakker RG, Goldstein SF, Silversmith RE, et al. CheX is a phosphorylated CheY phosphatase essential for Borrelia burgdorferi chemotaxis. J Bacteriol. 2005;187(23):7963–9. pmid:16291669
  13. 13. Motaleb MA, Sultan SZ, Miller MR, Li C, Charon NW. CheY3 of Borrelia burgdorferi is the key response regulator essential for chemotaxis and forms a long-lived phosphorylated intermediate. J Bacteriol. 2011;193(13):3332–41. pmid:21531807
  14. 14. Zhang K, Liu J, Charon NW, Li C. Hypothetical Protein BB0569 Is Essential for Chemotaxis of the Lyme Disease Spirochete Borrelia burgdorferi. J Bacteriol. 2015;198(4):664–72. pmid:26644432
  15. 15. Xu H, Sultan S, Yerke A, Moon KH, Wooten RM, Motaleb MA. Borrelia burgdorferi CheY2 Is Dispensable for Chemotaxis or Motility but Crucial for the Infectious Life Cycle of the Spirochete. Infect Immun. 2017;85(1). pmid:27799336
  16. 16. Zhang K, Liu J, Tu Y, Xu H, Charon NW, Li C. Two CheW coupling proteins are essential in a chemosensory pathway of Borrelia burgdorferi. Mol Microbiol. 2012;85(4):782–94. pmid:22780444
  17. 17. Sze CW, Zhang K, Kariu T, Pal U, Li C. Borrelia burgdorferi needs chemotaxis to establish infection in mammals and to accomplish its enzootic cycle. Infect Immun. 2012;80(7):2485–92. pmid:22508862
  18. 18. Sze CW, Li C. Chemotaxis Coupling Protein CheW(2) Is Not Required for the Chemotaxis but Contributes to the Full Pathogenicity of Borreliella burgdorferi. Infect Immun. 2023;91(4):e0000823.
  19. 19. Jacobs MB, Norris SJ, Phillippi-Falkenstein KM, Philipp MT. Infectivity of the highly transformable BBE02- lp56- mutant of Borrelia burgdorferi, the Lyme disease spirochete, via ticks. Infect Immun. 2006;74(6):3678–81. pmid:16714602
  20. 20. Jumper J, Evans R, Pritzel A, Green T, Figurnov M, Ronneberger O, et al. Highly accurate protein structure prediction with AlphaFold. Nature. 2021;596(7873):583–9. pmid:34265844
  21. 21. Muok AR, Briegel A, Crane BR. Regulation of the chemotaxis histidine kinase CheA: A structural perspective. Biochim Biophys Acta Biomembr. 2020;1862(1):183030. pmid:31374212
  22. 22. Bilwes AM, Alex LA, Crane BR, Simon MI. Structure of CheA, a signal-transducing histidine kinase. Cell. 1999;96(1):131–41. pmid:9989504
  23. 23. Jahreis K, Morrison TB, Garzon A, Parkinson JS. Chemotactic signaling by an Escherichia coli CheA mutant that lacks the binding domain for phosphoacceptor partners. J Bacteriol. 2004;186(9):2664–72. pmid:15090507
  24. 24. Muok AR, Ortega DR, Kurniyati K, Yang W, Maschmann ZA, Sidi Mabrouk A, et al. Atypical chemoreceptor arrays accommodate high membrane curvature. Nat Commun. 2020;11(1):5763. pmid:33188180
  25. 25. Rego RO, Bestor A, Rosa PA. Defining the plasmid-borne restriction-modification systems of the Lyme disease spirochete Borrelia burgdorferi. J Bacteriol. 2011;193(5):1161–71. pmid:21193609
  26. 26. Casjens S, Palmer N, van Vugt R, Huang WM, Stevenson B, Rosa P, et al. A bacterial genome in flux: the twelve linear and nine circular extrachromosomal DNAs in an infectious isolate of the Lyme disease spirochete Borrelia burgdorferi. Mol Microbiol. 2000;35(3):490–516. pmid:10672174
  27. 27. Purser JE, Norris SJ. Correlation between plasmid content and infectivity in Borrelia burgdorferi. Proc Natl Acad Sci U S A. 2000;97(25):13865–70. pmid:11106398
  28. 28. Elias AF, Stewart PE, Grimm D, Caimano MJ, Eggers CH, Tilly K, et al. Clonal polymorphism of Borrelia burgdorferi strain B31 MI: implications for mutagenesis in an infectious strain background. Infect Immun. 2002;70(4):2139–50. pmid:11895980
  29. 29. Sze CW, Xu H, Motaleb MA, Wolgemuth CW, Liu J, Charon NW, et al. Dancing with the Star: Borrelia burgdorferi, a Solo Dancer with All the Right Moves. In: Radolf JD, Samuels DS, editors. Lyme Disease and Relapsing Fever Spirochetes: Genomics, Molecular Biology, Host Interactions and Disease Pathogenesis. U.K: Caister Academic Press; 2021. p. 221–50.
  30. 30. Carroll JA, Cordova RM, Garon CF. Identification of 11 pH-regulated genes in Borrelia burgdorferi localizing to linear plasmids. Infect Immun. 2000;68(12):6677–84. pmid:11083781
  31. 31. Bakker RG, Li C, Miller MR, Cunningham C, Charon NW. Identification of specific chemoattractants and genetic complementation of a Borrelia burgdorferi chemotaxis mutant: flow cytometry-based capillary tube chemotaxis assay. Appl Environ Microbiol. 2007;73(4):1180–8. pmid:17172459
  32. 32. Motaleb MA, Corum L, Bono JL, Elias AF, Rosa P, Samuels DS, et al. Borrelia burgdorferi periplasmic flagella have both skeletal and motility functions. Proc Natl Acad Sci U S A. 2000;97(20):10899–904. pmid:10995478
  33. 33. Kraiczy P, Skerka C, Kirschfink M, Zipfel PF, Brade V. Immune evasion of Borrelia burgdorferi: insufficient killing of the pathogens by complement and antibody. Int J Med Microbiol. 2002;291 Suppl 33:141–6. pmid:12141738
  34. 34. Kochi SK, Johnson RC. Role of immunoglobulin G in killing of Borrelia burgdorferi by the classical complement pathway. Infect Immun. 1988;56(2):314–21. pmid:3276623
  35. 35. Dulipati V, Meri S, Panelius J. Complement evasion strategies of Borrelia burgdorferi sensu lato. FEBS Lett. 2020;594(16):2645–56. pmid:32748966
  36. 36. Aslam B, Nisar MA, Khurshid M, Farooq Salamat MK. Immune escape strategies of Borrelia burgdorferi. Future Microbiol. 2017;12:1219–37. pmid:28972415
  37. 37. Skare JT, Garcia BL. Complement Evasion by Lyme Disease Spirochetes. Trends Microbiol. 2020;28(11):889–99. pmid:32482556
  38. 38. Anderson C, Brissette CA. The Brilliance of Borrelia: Mechanisms of Host Immune Evasion by Lyme Disease-Causing Spirochetes. Pathogens. 2021;10(3). pmid:33801255
  39. 39. Kariu T, Coleman AS, Anderson JF, Pal U. Methods for rapid transfer and localization of lyme disease pathogens within the tick gut. J Vis Exp. 2011(48). pmid:21372782
  40. 40. Koci J, Bernard Q, Yang X, Pal U. Borrelia burgdorferi surface protein Lmp1 facilitates pathogen dissemination through ticks as studied by an artificial membrane feeding system. Sci Rep. 2018;8(1):1910. pmid:29382879
  41. 41. Zhang X, Yang X, Kumar M, Pal U. BB0323 function is essential for Borrelia burgdorferi virulence and persistence through tick-rodent transmission cycle. J Infect Dis. 2009;200(8):1318–30. pmid:19754308
  42. 42. Caine JA, Lin YP, Kessler JR, Sato H, Leong JM, Coburn J. Borrelia burgdorferi outer surface protein C (OspC) binds complement component C4b and confers bloodstream survival. Cell Microbiol. 2017;19(12). pmid:28873507
  43. 43. Garcia BL, Zhi H, Wager B, Hook M, Skare JT. Borrelia burgdorferi BBK32 Inhibits the Classical Pathway by Blocking Activation of the C1 Complement Complex. PLoS Pathog. 2016;12(1):e1005404. pmid:26808924
  44. 44. Caine JA, Coburn J. Multifunctional and Redundant Roles of Borrelia burgdorferi Outer Surface Proteins in Tissue Adhesion, Colonization, and Complement Evasion. Front Immunol. 2016;7:442. pmid:27818662
  45. 45. Kenedy MR, Lenhart TR, Akins DR. The role of Borrelia burgdorferi outer surface proteins. FEMS Immunol Med Microbiol. 2012;66(1):1–19. pmid:22540535
  46. 46. Bunikis J, Noppa L, Bergstrom S. Molecular analysis of a 66-kDa protein associated with the outer membrane of Lyme disease Borrelia. FEMS Microbiol Lett. 1995;131(2):139–45. pmid:7557322
  47. 47. Sohaskey CD, Zuckert WR, Barbour AG. The extended promoters for two outer membrane lipoprotein genes of Borrelia spp. uniquely include a T-rich region. Mol Microbiol. 1999;33(1):41–51. pmid:10411722
  48. 48. Gur E, Biran D, Ron EZ. Regulated proteolysis in Gram-negative bacteria—how and when? Nat Rev Microbiol. 2011;9(12):839–48. pmid:22020261
  49. 49. Mahmoud SA, Chien P. Regulated Proteolysis in Bacteria. Annu Rev Biochem. 2018;87:677–96. pmid:29648875
  50. 50. Thompson C, Mason C, Parrilla S, Ouyang Z. The Lon-1 Protease Is Required by Borrelia burgdorferi To Infect the Mammalian Host. Infect Immun. 2020;88(6). pmid:32205400
  51. 51. Mason C, Thompson C, Ouyang Z. The Lon-2 protease of Borrelia burgdorferi is critical for infection in the mammalian host. Mol Microbiol. 2020;113(5):938–50. pmid:31955462
  52. 52. Coleman JL, Crowley JT, Toledo AM, Benach JL. The HtrA protease of Borrelia burgdorferi degrades outer membrane protein BmpD and chemotaxis phosphatase CheX. Mol Microbiol. 2013;88(3):619–33. pmid:23565798
  53. 53. Coleman JL, Toledo A, Benach JL. Borrelia burgdorferi HtrA: evidence for twofold proteolysis of outer membrane protein p66. Mol Microbiol. 2016;99(1):135–50.
  54. 54. Russell TM, Delorey MJ, Johnson BJ. Borrelia burgdorferi BbHtrA degrades host ECM proteins and stimulates release of inflammatory cytokines in vitro. Mol Microbiol. 2013;90(2):241–51.
  55. 55. Ye M, Sharma K, Thakur M, Smith AA, Buyuktanir O, Xiang X, et al. HtrA, a Temperature- and Stationary Phase-Activated Protease Involved in Maturation of a Key Microbial Virulence Determinant, Facilitates Borrelia burgdorferi Infection in Mammalian Hosts. Infect Immun. 2016;84(8):2372–81. pmid:27271745
  56. 56. Kumru OS, Bunikis I, Sorokina I, Bergstrom S, Zuckert WR. Specificity and role of the Borrelia burgdorferi CtpA protease in outer membrane protein processing. J Bacteriol. 2011;193(20):5759–65. pmid:21856844
  57. 57. Becker G, Klauck E, Hengge-Aronis R. Regulation of RpoS proteolysis in Escherichia coli: the response regulator RssB is a recognition factor that interacts with the turnover element in RpoS. Proc Natl Acad Sci U S A. 1999;96(11):6439–44. pmid:10339606
  58. 58. Schweder T, Lee KH, Lomovskaya O, Matin A. Regulation of Escherichia coli starvation sigma factor (sigma s) by ClpXP protease. J Bacteriol. 1996;178(2):470–6. pmid:8550468
  59. 59. Zhou Y, Gottesman S, Hoskins JR, Maurizi MR, Wickner S. The RssB response regulator directly targets sigma(S) for degradation by ClpXP. Genes Dev. 2001;15(5):627–37. pmid:11238382
  60. 60. Iliopoulou BP, Huber BT. Infectious arthritis and immune dysregulation: lessons from Lyme disease. Curr Opin Rheumatol. 2010;22(4):451–5. pmid:20375899
  61. 61. Chung Y, Zhang N, Wooten RM. Borrelia burgdorferi elicited-IL-10 suppresses the production of inflammatory mediators, phagocytosis, and expression of co-stimulatory receptors by murine macrophages and/or dendritic cells. PLoS One. 2013;8(12):e84980. pmid:24367705
  62. 62. Thompson D, Watt JA, Brissette CA. Host transcriptome response to Borrelia burgdorferi sensu lato. Ticks Tick Borne Dis. 2021;12(2):101638. pmid:33360384
  63. 63. Ramesh G, Santana-Gould L, Inglis FM, England JD, Philipp MT. The Lyme disease spirochete Borrelia burgdorferi induces inflammation and apoptosis in cells from dorsal root ganglia. J Neuroinflammation. 2013;10:88. pmid:23866773
  64. 64. Thompson D, Sorenson J, Greenmyer J, Brissette CA, Watt JA. The Lyme disease bacterium, Borrelia burgdorferi, stimulates an inflammatory response in human choroid plexus epithelial cells. PLoS One. 2020;15(7):e0234993. pmid:32645014
  65. 65. Murthy PK, Dennis VA, Lasater BL, Philipp MT. Interleukin-10 modulates proinflammatory cytokines in the human monocytic cell line THP-1 stimulated with Borrelia burgdorferi lipoproteins. Infect Immun. 2000;68(12):6663–9. pmid:11083779
  66. 66. Brown JP, Zachary JF, Teuscher C, Weis JJ, Wooten RM. Dual role of interleukin-10 in murine Lyme disease: regulation of arthritis severity and host defense. Infect Immun. 1999;67(10):5142–50. pmid:10496888
  67. 67. Sjowall J, Carlsson A, Vaarala O, Bergstrom S, Ernerudh J, Forsberg P, et al. Innate immune responses in Lyme borreliosis: enhanced tumour necrosis factor-alpha and interleukin-12 in asymptomatic individuals in response to live spirochetes. Clin Exp Immunol. 2005;141(1):89–98. pmid:15958074
  68. 68. Dennis VA, Jefferson A, Singh SR, Ganapamo F, Philipp MT. Interleukin-10 anti-inflammatory response to Borrelia burgdorferi, the agent of Lyme disease: a possible role for suppressors of cytokine signaling 1 and 3. Infect Immun. 2006;74(10):5780–9. pmid:16988256
  69. 69. Radolf JD, Caimano MJ, Stevenson B, Hu LT. Of ticks, mice and men: understanding the dual-host lifestyle of Lyme disease spirochaetes. Nat Rev Microbiol. 2012;10(2):87–99. pmid:22230951
  70. 70. Samuels DS. Gene regulation in Borrelia burgdorferi. Annu Rev Microbiol. 2011;65:479–99. pmid:21801026
  71. 71. Samuels DS, Lybecker MC, Yang XF, Ouyang Z, Bourret TJ, Boyle WK, et al. Gene Regulation and Transcriptomics. Curr Issues Mol Biol. 2021;42:223–66. pmid:33300497
  72. 72. Iyer R, Caimano MJ, Luthra A, Axline D Jr., Corona A, Iacobas DA, et al. Stage-specific global alterations in the transcriptomes of Lyme disease spirochetes during tick feeding and following mammalian host adaptation. Mol Microbiol. 2015;95(3):509–38. pmid:25425211
  73. 73. Novak EA, Sekar P, Xu H, Moon KH, Manne A, Wooten RM, et al. The Borrelia burgdorferi CheY3 response regulator is essential for chemotaxis and completion of its natural infection cycle. Cell Microbiol. 2016;18(12):1782–99. pmid:27206578
  74. 74. Sultan SZ, Sekar P, Zhao X, Manne A, Liu J, Wooten RM, et al. Motor rotation is essential for the formation of the periplasmic flagellar ribbon, cellular morphology, and Borrelia burgdorferi persistence within Ixodes scapularis tick and murine hosts. Infect Immun. 2015;83(5):1765–77. pmid:25690096
  75. 75. Charon NW, Cockburn A, Li C, Liu J, Miller KA, Miller MR, et al. The unique paradigm of spirochete motility and chemotaxis. Annu Rev Microbiol. 2012;66:349–70. pmid:22994496
  76. 76. Grassmann AA, Tokarz R, Golino C, McLain MA, Groshong AM, Radolf JD, et al. BosR and PlzA reciprocally regulate RpoS function to sustain Borrelia burgdorferi in ticks and mammals. J Clin Invest. 2023. pmid:36649080
  77. 77. Kurokawa C, Lynn GE, Pedra JHF, Pal U, Narasimhan S, Fikrig E. Interactions between Borrelia burgdorferi and ticks. Nat Rev Microbiol. 2020;18(10):587–600. pmid:32651470
  78. 78. Helble JD, McCarthy JE, Hu LT. Interactions between Borrelia burgdorferi and its hosts across the enzootic cycle. Parasite Immunol. 2021;43(5):e12816. pmid:33368329
  79. 79. Takacs CN, Wachter J, Xiang Y, Ren Z, Karaboja X, Scott M, et al. Polyploidy, regular patterning of genome copies, and unusual control of DNA partitioning in the Lyme disease spirochete. Nat Commun. 2022;13(1):7173. pmid:36450725
  80. 80. Dunham-Ems SM, Caimano MJ, Eggers CH, Radolf JD. Borrelia burgdorferi requires the alternative sigma factor RpoS for dissemination within the vector during tick-to-mammal transmission. PLoS Pathog. 2012;8(2):e1002532. pmid:22359504
  81. 81. Hyde JA. Borrelia burgdorferi Keeps Moving and Carries on: A Review of Borrelial Dissemination and Invasion. Front Immunol. 2017;8:114. pmid:28270812
  82. 82. Norman MU, Moriarty TJ, Dresser AR, Millen B, Kubes P, Chaconas G. Molecular mechanisms involved in vascular interactions of the Lyme disease pathogen in a living host. PLoS Pathog. 2008;4(10):e1000169.
  83. 83. Kumar D, Ristow LC, Shi M, Mukherjee P, Caine JA, Lee WY, et al. Intravital Imaging of Vascular Transmigration by the Lyme Spirochete: Requirement for the Integrin Binding Residues of the B. burgdorferi P66 Protein. PLoS Pathog. 2015;11(12):e1005333.
  84. 84. Moriarty TJ, Shi M, Lin YP, Ebady R, Zhou H, Odisho T, et al. Vascular binding of a pathogen under shear force through mechanistically distinct sequential interactions with host macromolecules. Mol Microbiol. 2012;86(5):1116–31. pmid:23095033
  85. 85. Brooks CS, Hefty PS, Jolliff SE, Akins DR. Global analysis of Borrelia burgdorferi genes regulated by mammalian host-specific signals. Infect Immun. 2003;71(6):3371–83. pmid:12761121
  86. 86. Caimano MJ, Groshong AM, Belperron A, Mao J, Hawley KL, Luthra A, et al. The RpoS Gatekeeper in Borrelia burgdorferi: An Invariant Regulatory Scheme That Promotes Spirochete Persistence in Reservoir Hosts and Niche Diversity. Front Microbiol. 2019;10:1923. pmid:31507550
  87. 87. Kazmierczak MJ, Wiedmann M, Boor KJ. Alternative sigma factors and their roles in bacterial virulence. Microbiol Mol Biol Rev. 2005;69(4):527–43. pmid:16339734
  88. 88. Rodriguez Ayala F, Bartolini M, Grau R. The Stress-Responsive Alternative Sigma Factor SigB of Bacillus subtilis and Its Relatives: An Old Friend With New Functions. Front Microbiol. 2020;11:1761. pmid:33042030
  89. 89. Brown KL, Hughes KT. The role of anti-sigma factors in gene regulation. Mol Microbiol. 1995;16(3):397–404. pmid:7565101
  90. 90. Hughes KT, Mathee K. The anti-sigma factors. Annu Rev Microbiol. 1998;52:231–86. pmid:9891799
  91. 91. Paget MS. Bacterial Sigma Factors and Anti-Sigma Factors: Structure, Function and Distribution. Biomolecules. 2015;5(3):1245–65. pmid:26131973
  92. 92. Iyer R, Schwartz I. Microarray-Based Comparative Genomic and Transcriptome Analysis of Borrelia burgdorferi. Microarrays (Basel). 2016;5(2). pmid:27600075
  93. 93. Dulebohn DP, Hayes BM, Rosa PA. Global repression of host-associated genes of the Lyme disease spirochete through post-transcriptional modulation of the alternative sigma factor RpoS. PLoS One. 2014;9(3):e93141. pmid:24671196
  94. 94. Sarkar A, Hayes BM, Dulebohn DP, Rosa PA. Regulation of the virulence determinant OspC by bbd18 on linear plasmid lp17 of Borrelia burgdorferi. J Bacteriol. 2011;193(19):5365–73. pmid:21784941
  95. 95. Wachter J, Cheff B, Hillman C, Carracoi V, Dorward DW, Martens C, et al. Coupled induction of prophage and virulence factors during tick transmission of the Lyme disease spirochete. Nat Commun. 2023;14(1):198. pmid:36639656
  96. 96. Stevenson B, von Lackum K, Riley SP, Cooley AE, Woodman ME, Bykowski T. Evolving models of Lyme disease spirochete gene regulation. Wien Klin Wochenschr. 2006;118(21–22):643–52. pmid:17160602
  97. 97. Sze CW, Li C. Inactivation of bb0184, which encodes carbon storage regulator A, represses the infectivity of Borrelia burgdorferi. Infect Immun. 2011;79(3):1270–9. pmid:21173314
  98. 98. Purser JE, Lawrenz MB, Caimano MJ, Howell JK, Radolf JD, Norris SJ. A plasmid-encoded nicotinamidase (PncA) is essential for infectivity of Borrelia burgdorferi in a mammalian host. Mol Microbiol. 2003;48(3):753–64. pmid:12694619
  99. 99. Xu Q, McShan K, Liang FT. Essential protective role attributed to the surface lipoproteins of Borrelia burgdorferi against innate defences. Mol Microbiol. 2008;69(1):15–29. pmid:18452586
  100. 100. Li C, Xu H, Zhang K, Liang FT. Inactivation of a putative flagellar motor switch protein FliG1 prevents Borrelia burgdorferi from swimming in highly viscous media and blocks its infectivity. Mol Microbiol. 2010;75(6):1563–76. pmid:20180908
  101. 101. Cugini C, Medrano M, Schwan TG, Coburn J. Regulation of expression of the Borrelia burgdorferi beta(3)-chain integrin ligand, P66, in ticks and in culture. Infect Immun. 2003;71(2):1001–7. pmid:12540584
  102. 102. Seshu J, Esteve-Gassent MD, Labandeira-Rey M, Kim JH, Trzeciakowski JP, Hook M, et al. Inactivation of the fibronectin-binding adhesin gene bbk32 significantly attenuates the infectivity potential of Borrelia burgdorferi. Mol Microbiol. 2006;59(5):1591–601. pmid:16468997
  103. 103. Sal MS, Li C, Motalab MA, Shibata S, Aizawa S, Charon NW. Borrelia burgdorferi uniquely regulates its motility genes and has an intricate flagellar hook-basal body structure. J Bacteriol. 2008;190(6):1912–21. pmid:18192386
  104. 104. Zhang K, Qin Z, Chang Y, Liu J, Malkowski MG, Shipa S, et al. Analysis of a flagellar filament cap mutant reveals that HtrA serine protease degrades unfolded flagellin protein in the periplasm of Borrelia burgdorferi. Mol Microbiol. 2019;111(6):1652–70. pmid:30883947
  105. 105. The PyMOL Molecular Graphics System. 2.0 ed: Schrödinger, LLC.
  106. 106. Jutras BL, Scott M, Parry B, Biboy J, Gray J, Vollmer W, et al. Lyme disease and relapsing fever Borrelia elongate through zones of peptidoglycan synthesis that mark division sites of daughter cells. Proc Natl Acad Sci U S A. 2016;113(33):9162–70. pmid:27506799
  107. 107. Pal U, Yang X, Chen M, Bockenstedt LK, Anderson JF, Flavell RA, et al. OspC facilitates Borrelia burgdorferi invasion of Ixodes scapularis salivary glands. J Clin Invest. 2004;113(2):220–30. pmid:14722614