By utilizing an experimental approach that monitors the immediate global response of Chlamydia trachomatis to iron starvation, clues to long-standing issues in Chlamydia biology are revealed, including how Chlamydia adapts to this stress. We determined that this pathogen initiates a transcriptional program that prioritizes replenishment of nutrient stores over replication, possibly in preparation for rapid growth once optimal iron levels are restored. Transcription of genes for biosynthesis of metabolic precursors was generally upregulated, while those involved in multiple steps of translation were downregulated. We also observed an increase in transcription of genes involved in DNA repair and neutralizing oxidative stress, indicating that Chlamydia employs an “all-or-nothing” strategy. Its small genome limits its ability to tailor a specific response to a particular stress. Therefore, the “all-or-nothing” strategy may be the most efficient way of surviving within the host, where the pathogen likely encounters multiple simultaneous immunological and nutritional insults.
KEYWORDS: Chlamydia, microbiology, global regulatory networks, intracellular bacteria, iron reduction, stress response, stringent response, systems, transcriptional regulation, translational control
ABSTRACT
Iron is essential for growth and development of Chlamydia. Its long-term starvation in cultured mammalian cells leads to production of aberrant noninfectious chlamydial forms, also known as persistence. Immediate transcriptional responses to iron limitation have not been characterized, leaving a knowledge gap of how Chlamydia regulates its response to changes in iron availability. We used the fast-chelating agent 2,2′-bipyridyl (BPDL) to homogeneously starve Chlamydia trachomatis serovar L2 of iron, starting at 6 or 12 h postinfection. Immediate transcriptional responses were monitored after only 3 or 6 h of BPDL treatment, well before formation of aberrant Chlamydia. The first genomewide transcriptional response of C. trachomatis to iron starvation was subsequently determined utilizing RNA sequencing. Only 7% and 8% of the genome were differentially expressed in response to iron starvation at the early and middle stages of development, respectively. Biological pathway analysis revealed an overarching theme. Synthesis of macromolecular precursors (deoxynucleotides, amino acids, charged tRNAs, and acetyl coenzyme A [acetyl-CoA]) was upregulated, while energy-expensive processes (ABC transport and translation) were downregulated. A large fraction of differentially downregulated genes are involved in translation, including those encoding ribosome assembly and initiation and termination factors, which could be analogous to the translation downregulation triggered by stress in other prokaryotes during stringent responses. Additionally, transcriptional upregulation of DNA repair, oxidative stress, and tryptophan salvage genes reveals a possible coordination of responses to multiple antimicrobial and immunological insults. These responses of replicative-phase Chlamydia to iron starvation indicate a prioritization of survival over replication, enabling the pathogen to “stock the pantry” with ingredients needed for rapid growth once optimal iron levels are restored.
IMPORTANCE By utilizing an experimental approach that monitors the immediate global response of Chlamydia trachomatis to iron starvation, clues to long-standing issues in Chlamydia biology are revealed, including how Chlamydia adapts to this stress. We determined that this pathogen initiates a transcriptional program that prioritizes replenishment of nutrient stores over replication, possibly in preparation for rapid growth once optimal iron levels are restored. Transcription of genes for biosynthesis of metabolic precursors was generally upregulated, while those involved in multiple steps of translation were downregulated. We also observed an increase in transcription of genes involved in DNA repair and neutralizing oxidative stress, indicating that Chlamydia employs an “all-or-nothing” strategy. Its small genome limits its ability to tailor a specific response to a particular stress. Therefore, the “all-or-nothing” strategy may be the most efficient way of surviving within the host, where the pathogen likely encounters multiple simultaneous immunological and nutritional insults.
INTRODUCTION
The sexually transmitted bacterium Chlamydia trachomatis infects the mucosal epithelium of the endocervix, urethra, and anogenital tract. These infections usually resolve spontaneously, and most are asymptomatic and thus underreported. Over 1.5 million cases of C. trachomatis genital infections were reported in the United States in 2015 alone (1). As many as 17% of females infected with C. trachomatis develop long-term infections in the genital tract, which can result in serious complications such as pelvic inflammatory disease (PID), fallopian-tube scarring, and ectopic pregnancy, all of which are major risk factors for tubal factor infertility (TFI) (2). Rectal infections with lymphogranuloma venereum (LGV) serovars of C. trachomatis can be invasive and, if untreated, can lead to complications such as proctocolitis, inguinal adenopathy, reactive arthropathy, and colorectal ulcers (3). In some patients, infection persists even after antibiotic treatment (4, 5). The ability of C. trachomatis to survive over the long term in some individuals despite host immunity and antibiotic treatment is not well understood and may be associated with Chlamydia’s ability to become persistent (6). While aberrant chlamydial forms have been identified in cervical samples, the clinical relevance of this phenomenon is not well understood (6, 7).
Chlamydiae are obligate intracellular Gram-negative bacteria that undergo a biphasic developmental cycle that includes both nonreplicative and replicative forms (8). Infection begins when the small, metabolically quiescent chlamydial elementary body (EB) binds to mucosal epithelial cells and translocates virulence factors that induce its endocytic uptake. Within 2 h of entry, the EB differentiates into its replicative form, the reticulate body (RB). Continued secretion of effectors leads to modification of the endocytic vesicle such that it avoids fusion with the lysosome and enables capture of nutrient-rich vesicles. This unique intracellular niche, called the inclusion, continues to expand as RBs replicate. In response to unknown signals present at around 24 h postinfection (p.i.). RBs then differentiate into infectious EBs, followed by EB release 36 to 72 h postinfection (8). Under conditions of exposure to certain forms of stress in cell culture (e.g., penicillin treatment, interferon gamma [IFN-γ] treatment, iron depletion, or tryptophan [Trp] depletion), RBs do not differentiate into EBs but instead enter into a state of persistence characterized by aberrant, enlarged morphology (9–13). Persistent Chlamydia bacteria are resistant to both antibiotics and host immunity mechanisms and can recover from this state upon removal of stress or addition of missing nutrients (14–17).
Chlamydiae have undergone reductive evolution as they have adapted to intracellular growth in mammalian cells, discarding metabolic genes responsible for synthesizing factors that could be acquired from the host (18). The core genome of C. trachomatis serovar L2 includes only 889 open reading frames, making Chlamydia dependent on its host for lipids, nucleotides, amino acids, and metal cofactors (18). Exposure of Chlamydia-infected cells to immune mediators, such as IFN-γ, reduces the availability of these factors and results in reduced RB division and differentiation (12, 14). IFN-γ induces intracellular depletion of tryptophan by increasing levels of indolamine 2,3-dioxygenase (IDO), which is responsible for catabolizing tryptophan into kynurenines, which cannot be utilized in tryptophan metabolism (19).
Induction of inflammatory cytokines such as IFN-γ and interleukin-6 (IL-6) in response to chlamydial infection likely causes sequestration of free iron by the activity of the mononuclear phagocytic system, which includes both cellular and systemic regulatory pathways (20–27). Readers are referred to two comprehensive reviews of the coordinated regulation of iron homeostasis by systemic and cellular mechanisms (26, 28). In the context of Chlamydia infection of the genital epithelium, iron availability in infected cells is likely limited by downregulation of transferrin receptor and upregulation of the iron-storage factor ferritin (29). Iron levels in the female genital tract can also fluctuate throughout the menstrual cycle, in part due to hormone-induced expression of lactoferrin (30, 31). Iron is essential for growth and development of Chlamydia, and its acquisition and accumulation must be carefully regulated. In mammals, readily usable ferrous iron (Fe2+) is tied up in molecular complexes, limiting their interaction with hydrogen peroxide to form damaging hydroxyl radicals through the Fenton reaction (32). Eukaryotic stores of ferric iron (Fe3+) are strictly regulated to restrict access by pathogenic bacteria (33). Extracellular bacteria such as Pseudomonas and Yersinia utilize multiple redundant iron-binding molecules called siderophores that compete with mammalian transferrin for ferrous iron (34, 35). Intracellular bacteria, such as Mycobacterium, Francisella, and Chlamydia, can obtain iron by subverting host vesicles that contain holo-transferrin bound to transferrin receptor (36–39). Using a combination of endocytic markers and chemical inhibitors, members of our laboratory discovered that Chlamydia specifically recruits transferrin-containing vesicles from the slow-recycling endocytic pathway (38). Once delivered into the inclusion, iron is likely imported into bacteria through an ABC transporter system, encoded by ytgABCD, which is the only known iron acquisition system in Chlamydia species (40–42). The C terminus of the YtgC permease, referred to as YtgR, is homologous to the Corynebacterium repressor DtxR and has been recently characterized as an iron-dependent repressor of the ytgABCD iron acquisition operon (42). A recent review highlights the differences between the iron acquisition strategies of Chlamydia and those of other intracellular bacteria (27).
Conversion to the aberrant phenotype in response to iron starvation reduces the infectious potential of Chlamydia, since only a portion of RBs recover from stress and complete development into infectious EBs once iron is added back into the media (43). Previous studies have characterized aberrant C. trachomatis after long-term treatment with the iron chelator deferoxamine and have detected increased expression of the iron-binding protein YtgA, indicating its role in iron uptake (41, 44, 45). However, the researchers who performed those studies added deferoxamine at the time of infection and did not monitor transcriptional or proteomic patterns until ≥24 h postinfection, making it difficult to determine whether the upregulation was part of the initial response to iron starvation. Immediate genomewide transcriptional responses to iron limitation have not yet been characterized in detail, leaving a gap in the current knowledge of how Chlamydia regulates its response to changes in iron availability. We utilized the chelator 2,2-bipyridyl (BPDL), which can quickly and efficiently chelate free iron from both bacterial and mammalian cells, to induce an immediate transcriptional response to iron starvation by C. trachomatis (43, 46–48).
This report provides the first global profile of the C. trachomatis transcriptional response to iron starvation. Our short-term, effective treatment regimen, in combination with deep RNA sequencing (RNA-seq), revealed the immediate response of Chlamydia to iron starvation in the logarithmic phase of growth when the bacteria are in the RB form. Here, we utilized this data set to map the specific biological pathways altered in response to iron starvation. Taken together, our results provide important clues to how Chlamydia survives iron limitation. Accumulation of metabolite precursors is prioritized over macromolecular biosynthesis. In addition, the transcriptional induction of genes involved in adaptation to other stress factors, e.g., oxidative stress and amino acid starvation, points to the inability of Chlamydia to tailor its transcriptional response to a specific stress. Lastly, the global transcriptomic profile of iron-starved Chlamydia provides valuable insights into how the biphasic developmental cycle might irreversibly switch to persistence.
RESULTS
Treatment optimization to detect the immediate chlamydial response to iron starvation.
The bivalent chelator 2,2-bipyridyl (BPDL) has been shown to deplete both ferrous iron and ferric iron from Chlamydia-infected cells during long-term treatment, and it induces the development of aberrant forms more consistently and homogenously than the previously used ferric iron chelator, deferoxamine (43). Here, we determined the optimal duration of BPDL treatment to induce iron-responsive transcription without inducing morphological abnormalities in Chlamydia. We chose to begin starvation during midcycle development (12 h p.i.) instead of at the beginning of infection for two reasons: (i) to test the response of actively replicating Chlamydia bacteria that are able to maximally respond to stress and (ii) to ensure that the treated and mock-treated Chlamydia bacteria remained in the same stage of development (RB). We monitored chlamydial morphology, growth, and transcriptional responses after 3, 6, and 12 h of BPDL treatment (Fig. 1A). Indirect immunofluorescent confocal microscopy revealed similar morphologies for the mock-treated and BPDL-treated forms for up to 12 h of BPDL treatment (Fig. 1B). Interestingly, observation of BPDL-treated cultures by light microscopy revealed an obvious decrease in Brownian movement within inclusions after 6 or more hours of treatment (data not shown). This observation is consistent with our findings showing that chlamydial growth is reduced compared to that seen with mock treatment after only 6 h of BPDL treatment, as determined by quantitative PCR (qPCR) analysis of chlamydial genomes (Fig. 1C).
We monitored the transcriptional response of the known iron-responsive genes ytgA and ahpC by reverse transcriptase quantitative PCR (RT-qPCR) to validate the iron starvation protocol (40, 41, 43, 49). Elevated (1.5-fold and 1.7-fold) transcription of both iron starvation markers was detected after only 6 h of BPDL treatment compared to the mock treatment results. Maximum differences in transcription of both markers (2.8-fold and 3.7-fold) were detected after 12 h of BPDL treatment (Fig. 1D). In the same experiment, we monitored the transcriptional profile of the early-stage gene euo, whose transcription decreases during late stages of normal development. Multiple persistence models have demonstrated dysregulated euo transcription, with high levels of euo mRNA detected late in development under persistence-inducing conditions (15, 43, 45). After 12 h of BPDL treatment, we observed that euo transcript levels remained elevated relative to the mock-treated control results, indicating dysregulated transcription or a possible delay in development (Fig. 1E, top). The idea that a delay in development occurs during longer BPDL treatment is supported by the lack of recoverable inclusion-forming units (IFUs) detected after 12 or 24 h of BPDL treatment compared to mock-treated controls (Fig. 1F), indicating a possible lack of RB-to-EB differentiation. Because 6 h of BPDL treatment is sufficient to induce iron-responsive transcription without inducing the patterns of morphology and transcription associated with persistence, we chose that as the optimal duration of iron starvation for our genomewide transcriptional studies. We also included 3 h of BPDL treatment to detect the earliest possible response to iron starvation prior to BPDL-induced changes in growth.
Global transcriptional response of C. trachomatis to iron starvation during midcycle development.
The primary global response of C. trachomatis to midcycle iron starvation was determined by RNA sequencing (RNA-seq). We utilized an Ion Proton chip for sequencing, which allowed easy and rapid scaling of time points based on observed yields of mapped reads. This approach is relevant to the study of Chlamydia transcription because chlamydial mRNA represents a small proportion of the total RNA at the time points analyzed, even after significant enrichment steps. For midcycle iron starvation studies, we aimed for greater than 10× coverage of 100% of the C. trachomatis genome, with a minimum of 3 biological replicates per sample. The sequencing reads were trimmed to exclude adaptor sequences and polyclonal reads, followed by exclusion of reads less than 30 nucleotides (nt) in length. The remaining reads were aligned to the C. trachomatis genome and plasmid, with 2% to 23% of trimmed reads mapping. Average read lengths ranged from 92 to 134 nucleotides, requiring an average of 108,837 mapped reads to reach our coverage goal. A summary of the read and mapping statistics for all of our samples can be found in Table S1 in the supplemental material. Alignments, comparisons, and normalization of aligned reads were done with CLC Genomics Workbench version 9.0 according to default settings. All midcycle conditions (12 h untreated, 12 + 3 h BPDL, 12 + 3 h mock, 12 h + 6 h BPDL, 12 + 6 h mock) were compared using the CLC Genomics experiment tool, normalized by quantile scaling, and analyzed for differential gene expression levels using EdgeR statistical analysis with false-discovery-rate (FDR) calculation. Because we included ribosomal rRNA, eukaryotic mRNA, and small (<100-nt) RNA depletion steps when preparing chlamydial mRNA for RNA-seq, we also excluded tRNAs, rRNAs, and genes with <10 mean reads in a sample group prior to normalization and analysis.
The genomewide profile of mock-treated and BPDL-treated gene expression during midcycle development (12 h to 18 h postinfection) is displayed as a heat map of normalized expression values (Fig. 2A). The raw and normalized data for these individual replicates can be found in Table S4. The mock treatment (left) and BPDL treatment (right) profiles were remarkably similar across all genes whose expression significantly changed during normal midcycle development of Chlamydia (based on comparisons of data from 12 h versus 15 h, 15 h versus 18 h, or 12 h versus 18 h; P value of ≤0.01). The annotated expression heat map and EdgeR comparisons for normal growth can be found in Fig. S1 and Table S2 in the supplemental material, respectively. The entire RNA-seq data set of normal development can be found in Table S3. The similarity between global expression profiles indicates that the normal development of Chlamydia is not dysregulated after only 3 h or 6 h of BPDL treatment. However, EdgeR analysis of BPDL-treated cultures compared to mock-treated samples (at equivalent time points postinfection) revealed that 8% (76/889) of the genome was differentially expressed after 3 h BPDL treatment and 1% (12/889) was differentially expressed after 6 h of BPDL treatment. Genes that were differentially expressed with a maximum P value of 0.01 are displayed in a heat map of fold change differences between BPDL-treated and mock-treated samples (Fig. 2B). Examples of decreased transcription after 3 h and 6 h of BPDL treatment include the ribosomal subunit genes rpsO and rpsT and the type III secretion genes copB and scc2, respectively. Transcription of the tryptophan salvage pathway operon trpBA and the ribonucleotide reductase operon nrdAB was significantly increased after both 3 and 6 h of treatment. Iron-responsive genes that were differentially expressed with a P value of <0.01 after 3 or 6 h of BPDL treatment are listed in Tables 1 and 2, respectively. The fully annotated heat map can be found in Fig. S2, and the full set of RNA-seq results for midcycle iron starvation can be found in Table S4.
TABLE 1 .
Feature ID |
Locus tag |
Fold change |
P value | Annotation | Functional category |
UniProtKB ID |
---|---|---|---|---|---|---|
CTL0013 | CTL0013 | 3.45 | 3.10E−4 | Hypothetical, YGGT family | Hypothetical | A0A0H3MB25 |
trpB | CTL0423 | 3.18 | 3.23E−7 | Tryptophan synthase subunit B | Amino acid biosynthesis | A0A0H3MD30 |
glgA | CTL0167 | 2.95 | 1.24E−4 | Glycogen synthase | Energy metabolism | B0B925 |
murB | CTL0203 | 2.52 | 6.41E−3 | UDP-N-acetylenolpyruvoylglucosamine reductase | Other | B0B960 |
CTL0525 | CTL0525 | 2.38 | 4.53E−5 | TPR-containing domain | Hypothetical | A0A0H3MKX6 |
recA | CTL0018 | 2.21 | 1.98E−3 | Recombinase A | DNA replication and repair | B0B8M5 |
CTL0339 | CTL0339 | 2.2 | 2.29E−3 | Phosphatidylcholine-hydrolyzing phospholipase D | Other | A0A0H3MCY1 |
aroL | CTL0621 | 2.13 | 6.56E−3 | Shikimate kinase 2 | Amino acid biosynthesis | A0A0H3MDG7 |
hemE | CTL0116 | 2.09 | 1.68E−3 | Uroporphyrinogen decarboxylase | Cofactor biosynthesis | B0B8X3 |
CTL0408 | CTL0408 | 2.08 | 1.36E−5 | MIR, MAC/perforin domain-containing protein | Other | A0A0H3MGT6 |
nrdA | CTL0199 | 2 | 1.48E−5 | Ribonucleoside-diphosphate reductase | DNA replication and repair | A0A0H3MCP2 |
mqnD | CTL0514 | 2 | 2.45E−3 | 1,4-Dihydroxy-6-naphtoate synthase | Cofactor biosynthesis | A0A0H3MC13 |
CTL0704 | CTL0704 | 1.99 | 0.01 | Hypothetical | Hypothetical | A0A0H3MCH5 |
trpA | CTL0424 | 1.94 | 4.15E−3 | Troptophan synthase subunit A | Amino acid biosynthesis | A0A0H3MKP4 |
CTL0255 | CTL0255 | 1.93 | 4.37E−3 | Hypothetical | Hypothetical | A0A0H3MGJ2 |
pepF | CTL0367 | 1.9 | 6.17E−4 | Endopeptidase F | Protein processing and folding | A0A0H3MKK2 |
rnc | CTL0549 | 1.89 | 0.01 | RNase III | Transcriptional regulation | B0B7L3 |
CTL0823 | CTL0823 | 1.88 | 7.21E−4 | Hypothetical | Hypothetical | A0A0H3MLF6 |
CTL0301 | CTL0301 | 1.81 | 1.32E−4 | Probable cytosol aminopeptidase PepA | Protein processing and folding | B0B9F3 |
CTL0884 | CTL0884 | 1.81 | 7.04E−3 | Hypothetical | Hypothetical | A0A0H3MCL0 |
CTL0885 | CTL0885 | 1.77 | 3.48E−3 | Hypothetical effector | Type III secretion | A0A0H3MHM9 |
lpdA | CTL0820 | 1.76 | 3.80E−3 | Dihydrolipoyl dehydrogenase | Energy metabolism | A0A0H3MHJ2 |
CTL0096 | CTL0096 | 1.72 | 1.98E−4 | Putative cation transporting ATPase | Nutrient transport | A0A0H3MCG9 |
gp6D | CTL0846 | 1.72 | 2.77E−3 | Virulence plasmid pGP6-D related protein | Hypothetical | A0A0H3MLH1 |
dnaQ | CTL0513 | 1.69 | 0.01 | DNA polymerase III subunit epsilon | DNA replication and repair | A0A0H3MKW6 |
CTL0512 | CTL0512 | 1.68 | 4.52E−3 | MCSC, secretion chaperone | Type III secretion | A0A0H3MDA7 |
CTL0847 | CTL0847 | 1.63 | 6.69E−3 | Hypothetical | Hypothetical | A0A0H3MCR9 |
oppA3 | CTL0450 | 1.61 | 9.20E−3 | Oligopeptide transporter | Nutrient transport | A0A0H3MKR3 |
CTL0102 | CTL0102 | 1.6 | 7.09E−3 | Putative exported protein | Hypothetical | A0A0H3MG91 |
lipA | CTL0821 | 1.59 | 8.90E−3 | Lipioic acid synthase | Other | B0B8D2 |
pheT | CTL0736 | 1.59 | 0.01 | Phenylalanine-tRNA ligase beta subunit | Translation | A0A0H3MHF2 |
cysS | CTL0151 | 1.56 | 2.47E−3 | Cysteinyl-tRNA synthase | Translation | A0A0H3MGC2 |
CTL0055 | CTL0055 | 1.51 | 7.04E−3 | Hypothetical | Hypothetical | A0A0H3MAN0 |
pal | CTL0863 | 1.51 | 7.29E−3 | Peptidogycan-associated lipoprotein | Other | A0A0H3MDU9 |
thrS | CTL0844 | 1.48 | 6.31E−3 | Threonine-tRNA ligase | Translation | B0B8F5 |
CTL0476 | CTL0476 | 1.48 | 0.01 | Candidate inclusion membrane protein | Hypothetical | A0A0H3MKT3 |
CTL0043 | CTL0043 | 0.01 | 0.01 | Type III secretion structural protein | Type III secretion | A0A0H3MG59 |
glmS | CTL0188 | 1.46 | 0.01 | Glutamine–fructose-6-phosphate aminotransferase | Energy metabolism | A0A0H3MCN0 |
CTL0684 | CTL0684 | 1.45 | 8.29E−3 | Hypothetical | Hypothetical | A0A0H3MCF9 |
nusA | CTL0352 | −1.48 | 4.90E−3 | Transcription termination factor | Transcriptional regulation | A0A0H3MGQ1 |
prtA | CTL0278 | −1.56 | 9.49E−3 | Peptide chain release factor RF1 | Translation | B0B9D0 |
rplW | CTL0788 | −1.58 | 5.43E−4 | Ribosomal subunit | Translation | B0B8A0 |
CTL0326 | CTL0326 | −1.62 | 6.61E−3 | YtgD, ABC transport protein, membrane permease | Nutrient transport | A0A0H3MGN6 |
trpF | CTL0581 | −1.67 | 2.26E−3 | N-(5′-Phosphoribosyl)anthranilate isomerase | Cofactor biosynthesis | B0B7P4 |
folX | CTL0878 | −1.68 | 7.23E−3 | Dihydroneopterin triphosphate 2′-epimerase | Cofactor biosynthesis | A0A0H3MCK6 |
pgsA_2 | CTL0757 | −1.71 | 1.74E−3 | CDP–diacylglycerol-glycerol-3-phosphate 3-phosphatidyl-transferase |
Other | A0A0H3MHG9 |
sodM | CTL0546 | −1.74 | 1.46E−3 | Superoxide dismutase | Redox homeostasis | A0A0H3MKY6 |
CTL0138 | CTL0138 | −1.74 | 4.48E−3 | Ribosomal silencing factor RafS | Translation | A0A0H3MAQ8 |
CTL0061 | CTL0061 | −1.75 | 2.26E−4 | Inorganic phosphate transporter | Nutrient transport | A0A0H3MG71 |
CTL0486 | CTL0486 | −1.77 | 3.48E−3 | Putative membrane transport protein | Nutrient transport | A0A0H3MBY7 |
gltT | CTL0658 | −1.85 | 3.07E−3 | Sodium:dicarboxylate symport protein | Nutrient transport | A0A0H3MCC1 |
smpB | CTL0332 | −1.88 | 0.01 | SsrA-binding protein | Translation | A0A0H3MBC6 |
aroA | CTL0620 | −1.91 | 1.85E−4 | 3-Phosphoshikimate 1-carboxyvinyltransferase | Energy metabolism | B0B7T5 |
CTL0132 | CTL0132 | −1.92 | 3.03E−3 | UPF0109-containing putative RNA-binding protein | Translation | A0A0H3MAQ7 |
CTL0548 | CTL0548 | −1.98 | 1.14E−4 | DcrA, putative nonheme Fe(II) 2-oxoglutarate | Hypothetical | A0A0H3MBY2 |
secG | CTL0606 | −2.01 | 0.01 | Protein export membrane protein SecG | Protein processing and folding | A0A0H3MH61 |
rpsK | CTL0770 | −2.07 | 1.59E−6 | Ribosomal subunit | Translation | B0B882 |
rplT | CTL0207 | −2.07 | 1.62E−6 | Ribosomal subunit | Translation | B0B964 |
rplN | CTL0780 | −2.08 | 3.98E−3 | Ribosomal subunit | Translation | B0B892 |
CTL0720 | CTL0720 | −2.12 | 5.09E−3 | SWIB domain-containing protein | Hypothetical | A0A0H3MC95 |
gcsH | CTL0534 | −2.14 | 9.08E−4 | Glycine cleavage system H protein | Amino acid biosynthesis | B0B7J8 |
CTL0680 | CTL0680 | −2.14 | 1.59E−3 | Putative rRNA processing peptide | Translation | A0A0H3MHB1 |
rpmL | CTL0206 | −2.14 | 1.75E−3 | Ribosomal subunit | Translation | A0A0H3MCP7 |
fer | CTL0315 | −2.26 | 6.48E−5 | Ferredoxin | Redox homeostasis | A0A0H3MCW5 |
CTL0552 | CTL0552 | −2.27 | 1.09E−3 | TPR-containing domain | Hypothetical | Pseudogene |
CTL0222 | CTL0222 | −2.27 | 6.58E−3 | Hypothetical | Hypothetical | A0A0H3MK96 |
infA2 | CTL0575 | −2.29 | 6.70E−6 | Translation initiation factor IF-1 | Translation | A0A0H3MDD9 |
rpsT | CTL0881 | −2.38 | 1.13E−4 | Ribosomal subunit | Translation | B0B8J2 |
CTL0335 | CTL0335 | −2.43 | 0.01 | Putative integral membrane protein | Hypothetical | A0A0H3MCX6 |
ltuA | CTL0631 | −2.68 | 3.38E−3 | Late transcription unit A protein | Hypothetical | A0A0H3MH76 |
rpsO | CTL0215 | −2.72 | 4.14E−4 | Ribosomal subunit | Translation | B0B972 |
ndk | CTL0762 | −2.76 | 7.97E−4 | Nucleoside diphosphate kinase | DNA replication and repair | B0B874 |
pGP8-D | pL2-02 | −2.99 | 0.01 | Virulence plasmid integrase pGP8-D | DNA replication and repair | B0BCM4 |
dut | CTL0544 | −3.02 | 6.06E−6 | Deoxyuridine 5′-triphosphate nucleotidohydrolase | DNA replication and repair | B0B7K8 |
rpmJ | CTL0154 | −3.03 | 3.18E−4 | Ribosomal subunit | Translation | B0B912 |
CTL0021 | CTL0021 | −3.16 | 0.01 | Hypothetical | Hypothetical | A0A0H3MG42 |
FDR-corrected P values can be found in Table S4. Data from upregulated and downregulated genes are shown in the top and bottom halves of the table, respectively. These data were exported from CLC Genomics Workbench 9.5.3. rRNAs, tRNAs, and features (genes) with fewer than 10 reads in all samples were eliminated from the data set prior to normalization and EDGE analysis, and the data include only those genes that were differentially expressed with a significance P value of ≤0.01. ID, identifier.
TABLE 2 .
Feature ID |
Locus tag |
P value | Fold change |
Annotation | Functional category |
UniProtKB ID |
---|---|---|---|---|---|---|
trpB | CTL0423 | 2.61E−7 | 3.5 | Tryptophan synthase subunit B | Amino acid biosynthesis | A0A0H3MD30 |
trpA | CTL0424 | 9.26E−6 | 3.21 | Tryptophan synthase subunit A | Amino acid biosynthesis | A0A0H3MKP4 |
nrdA | CTL0199 | 3.50E−6 | 2.39 | Ribonucleoside-diphosphate reductase | DNA replication and repair | A0A0H3MCP2 |
CTL0071 | CTL0071 | 6.73E−3 | 1.73 | Hypothetical | Hypothetical | A0A0H3MCG5 |
nrdB | CTL0200 | 9.34E−3 | 1.63 | Ribonucleoside-diphosphate reductase | DNA replication and repair | A0A0H3MK81 |
CTL0619 | CTL0619 | 1.68E−3 | −1.83 | Hypothetical integral membrane protein | Hypothetical | A0A0H3MH71 |
copD | CTL0842 | 3.17E−3 | −2.13 | Type III secretion system protein | Type III secretion | A0A0H3MHK7 |
scc2 | CTL0839 | 1.37E−3 | −2.18 | Type III secretion system chaperone | Type III secretion | A0A0H3MLG7 |
tsp | CTL0700 | 1.28E−3 | −2.34 | Tail-specific protease | Protein processing and folding | A0A0H3MDM0 |
CTL0185 | CTL0185 | 6.85E−3 | −2.76 | Hypothetical membrane protein | Hypothetical | A0A0H3MAV2 |
copB | CTL0841 | 6.79E−5 | −2.81 | Type III secretion system membrane protein | Type III secretion | A0A0H3MCF0 |
CTL0840 | CTL0840 | 2.15E−3 | −2.97 | Hypothetical | Hypothetical | A0A0H3MCQ4 |
FDR-corrected P values can be found in Table S4. These data were exported from CLC Genomics Workbench 9.5.3. rRNAs, tRNAs, and features (genes) with fewer than 10 reads in all samples were eliminated from the data set prior to normalization and EDGE analysis. These data include only genes that were differentially expressed with a significance P value of ≤0.01.
Functional categorization of genes differentially expressed during midcycle response to iron starvation.
Annotations and functional categories of genes differentially expressed during midcycle iron starvation were retrieved from UniProt and are listed in Tables 1 and 2. Genes differentially expressed, with a minimum P value of 0.01, after only 3 h of BPDL treatment are grouped by functional category of induced and reduced transcripts (Table 1; Fig. 3) (50). Of the 39 genes significantly induced after only 3 h of iron starvation, representing 4% of the genome, five categories were equally represented with 3 genes each: energy metabolism (glgA, lpdA, and glmS), amino acid biosynthesis (trpA, trpB, and aroL), DNA replication and repair (nrdA, recA, and dnaQ), type III secretion (mcsC, CTL0085, and CTL0043), and translation (pheT, cysS, and thrS). Of the 37 genes that were significantly reduced in response to 3 h of BPDL treatment, representing 4% of the genome, the majority (39%) are associated with translation (prfA, rplW, rsfS, smpB, CTL0132, rpsK, rplT, rplN, CTL0680, rpmI, rmpJ, infa2, rpsT, and rpsO), and 11% are associated with nutrient transport (CTL0061, CTL0485, gltT, and ytgD). Transcript levels of only 5 genes (trpB, trpA, nrdA, CTL0071, and nrdB) were significantly increased after 6 h of BPDL treatment, while transcript levels of 7 genes (CTL0185, CTL0619, tsp, and the entire scc2-CTL0840-cobB-copD operon) were decreased (Table 2).
To independently confirm the midcycle response detected by RNA sequencing, we utilized RT-qPCR. Increased transcription in response to iron starvation was confirmed for all of the transcripts tested by RT-qPCR, with the exception of recA (Fig. 4A). None of the tested downregulated genes were significantly reduced in expression compared to controls as determined by RT-qPCR, likely due to the fact that the genes tested were very low in abundance at the time points tested (Fig. 4B).
Functional categorization of genes differentially expressed during early-cycle response to iron starvation.
Chlamydia infections of the genital tract are asynchronous. Thus, Chlamydia could be exposed to host-induced stress at any point in the developmental cycle. For this reason, we extended our analysis of the immediate response to iron starvation to an earlier point in the developmental cycle. Chlamydia-infected cells were treated with BPDL starting at 6 h postinfection, which is a time point after the initial EB-to-RB differentiation and at the beginning of the logarithmic-growth phase. RNA and genomic DNA (gDNA) were collected at 9 h postinfection for both treated and mock-treated samples. RNA-seq analyses and alignments were performed as described above. A summary of mapped reads and coverage can be found in Table S1.
Genes differentially expressed during the early-cycle response (6 + 3 h BPDL treatment versus 6 + 3 h mock treatment), with a maximum P value of 0.01, are grouped by functional categories of induced and reduced transcripts (Fig. 5). Data corresponding to the full set of differentially expressed genes and their annotations can be found in Table 3. Similarly to the results of analysis of the midcycle response, transcription of 4% of the genome, including genes involved in DNA replication and repair (nrdA, nrdB, mutS, dnaQ, and recA), amino acid biosynthesis (trpB, trpA, aspC_1, and glyA), and translation (CTL0111, trpS, thrS, and aspS), was induced during the early-cycle response to iron starvation. Uniquely, genes involved in redox homeostasis (pdi, ahpC, and sodM) were also upregulated in response to iron starvation starting at 6 h postinfection but not during the midcycle response. Of the 23 genes with reduced transcription during the early-cycle response to iron starvation (3% of the genome), 17% are associated with translation (rplW, prfA, rplC, and ftsY) and 13% with DNA replication and repair (pGP8D, amn, and dnaX_1).
TABLE 3 .
Feature ID |
Locus tag |
Fold change |
P value | Annotation | Functional category |
UniProtKB ID |
---|---|---|---|---|---|---|
CTL0149 | CTL0149 | 6.71 | 3.71E−3 | Protein disulfide isomerase | Redox homeostasis | A0A0H3MAT1 |
CTL0184 | CTL0184 | 3.4 | 4.44E−5 | Hypothetical inclusion membrane protein | Hypothetical | A0A0H3MBD4 |
trpA | CTL0424 | 3.12 | 3.27E−4 | Tryptophan synthase subunit A | Amino acid biosynthesis | A0A0H3MKP4 |
CTL0388 | CTL0388 | 2.93 | 3.02E−3 | Hypothetical methyltransferase | Hypothetical | A0A0H3MKL6 |
CTL0111 | CTL0111 | 2.89 | 0.000877 | rRNA methyltransferase TrmA | Translation | A0A0H3MAQ4 |
hemN_1 | CTL0115 | 2.47 | 1.00E−2 | Coproporphyrinogen-III oxidase | Cofactor biosynthesis | A0A0H3MJY5 |
mip | CTL0803 | 2.37 | 3.97E0-04 | Peptidyl-prolyl cis-trans-isomerase | Protein processing and folding | A0A0H3MDR1 |
trpB | CTL0423 | 2.28 | 2.55E−3 | Tryptophan synthase subunit B | Amino acid biosynthesis | A0A0H3MD30 |
lpxB | CTL0668 | 2.28 | 0.00923 | Lipid-A-disaccharide synthase | Other | A0A0H3MDJ6 |
nrdB | CTL0200 | 2.12 | 1.57E−6 | Ribonucleoside-diphosphate reductase subunit B | DNA replication and repair | A0A0H3MK81 |
nrdA | CTL0199 | 2.09 | 4.10E−12 | Ribonucleoside-diphosphate reductase subunit A | DNA replication and repair | A0A0H3MCP2 |
CTL0874 | CTL0874 | 2.04 | 2.07E−6 | CADD, PABA synthase | Cofactor biosynthesis | A0A0H3MHM3 |
CTL0360 | CTL0360 | 2.04 | 8.59E−3 | Hypothetical | Hypothetical | A0A0H3MKJ7 |
mutS | CTL0160 | 1.95 | 6.41E−6 | DNA mismatch repair protein | DNA replication and repair | B0B918 |
dnaQ | CTL0513 | 1.88 | 2.83E−3 | DNA polymerase III subunit epsilon | DNA replication and repair | A0A0H3MKW6 |
CTL0164 | CTL0164 | 1.86 | 1.20E−3 | Hypothetical exported protein | Hypothetical | A0A0H3MBC3 |
CTL0791 | CTL0791 | 1.82 | 1.78E−7 | Hypothetical membrane protein | Hypothetical | A0A0H3MCL8 |
trpS | CTL0848 | 1.81 | 1.46E−3 | Tryptophan-tRNA ligase | Translation | A0A0H3MCF4 |
aspC_1 | CTL0005 | 1.77 | 1.92E−5 | Aminotransferase | Amino acid biosynthesis | A0A0H3MG09 |
eno | CTL0850 | 1.75 | 2.51E−3 | Enolase | Energy metabolism | B0B8G1 |
CTL0408 | CTL0408 | 1.73 | 3.50E−4 | MIR, MAC/perforin domain-containing protein | Other | A0A0H3MGT6 |
recA | CTL0018 | 1.72 | 1.49E−3 | Recombinase A | DNA replication and repair | B0B8M5 |
sodM | CTL0546 | 1.65 | 7.17E−3 | Superoxide dismutase | Redox homeostasis | A0A0H3MKY6 |
brnQ | CTL0817 | 1.62 | 3.60E−4 | Branched-chain amino acid transporter | Nutrient transport | A0A0H3MLF2 |
greA | CTL0004 | 1.59 | 2.63E−4 | Transcription elongation factor | Transcriptional regulation | A0A0H3MAD9 |
CTL0102 | CTL0102 | 1.58 | 4.63E−3 | Hypothetical exported protein | Hypothetical | A0A0H3MG91 |
ahpC | CTL0866 | 1.52 | 2.54E−3 | Thio-specific antioxidant peroxidase | Redox homeostasis | A0A0H3MCJ5 |
thrS | CTL0844 | 1.52 | 0.00501 | Threonine-tRNA ligase | Translation | B0B8F5 |
aspS | CTL0804 | 1.52 | 6.89E−3 | Aspartate-tRNA ligase | Translation | B0B8B6 |
rpoD | CTL0879 | 1.51 | 1.00E−2 | RNA polymerase sigma factor RpoD | Transcriptional regulation | A0A0H3MHM6 |
glyA | CTL0691 | 1.47 | 0.00242 | Serine hydroxymethyltransferase | Amino acid biosynthesis | B0B804 |
sctJ | CTL0.822 | 1.43 | 7.93E−3 | Type III secretion protein | Type III secretion | A0A0H3MDS0 |
rpoC | CTL0566 | −1.28 | 7.87E−3 | DNA-directed RNA polymerase subunit beta′ | Transcriptional regulation | B0B7N0 |
pGP8-D | L2b_RS04755 | −1.36 | 4.91E−3 | Virulence plasmid integrase pGP8-D | DNA replication and repair | B0BCM4 |
rplW | CTL0788 | −1.41 | 0.0035 | Ribosomal subunit | Translation | B0B8A0 |
prfA | CTL0278 | −1.44 | 5.80E−3 | Peptide chain release factor RF1 | Translation | B0B9D0 |
rplC | CTL0790 | −1.52 | 4.30E−4 | Ribosomal subunit | Translation | B0B8A2 |
CTL0061 | CTL0061 | −1.52 | 2.95E−3 | Inorganic phosphate transporter PHO4 | Nutrient transport | A0A0H3MG71 |
CTL0659 | CTL0659 | −1.57 | 9.75E−4 | Tetraacyldisaccharide 4′-kinase LpxK | Other | A0A0H3MC42 |
CTL0473 | CTL0473 | −1.57 | 1.22E−3 | Hypothetical exported protein | Hypothetical | A0A0H3MBQ9 |
incD | CTL0370 | −1.59 | 3.33E−3 | Inclusion membrane protein D | Other | B0B9M3 |
plsX | CTL0182 | −1.59 | 1.00E−2 | Phosphate acyltransferase | Other | B0B939 |
CTL0613 | CTL0613 | −1.6 | 1.34E−3 | Hypothetical inner membrane protein | Hypothetical | A0A0H3MC14 |
pmpA | CTL0669 | −1.63 | 0.00534 | Probable outer membrane protein PmpA | Other | A0A0H3ML49 |
CTL0548 | CTL0548 | −1.66 | 1.01E−3 | Hypothetical nonheme Fe(II) 2-oxoglutarate | Hypothetical | A0A0H3MBY2 |
CTL0541 | CTL0541 | −1.67 | 1.00E−2 | Hypothetical membrane protein | Hypothetical | A0A0H3MC35 |
sucB_2 | CTL0311 | −1.7 | 7.85E−3 | Dihydrolipoyllysine-residue succinyltransferase | Energy metabolism | A0A0H3ML42 |
amn | CTL0120 | −1.71 | 4.71E−3 | AMP nucleosidase | DNA replication and repair | A0A0H3MGA3 |
mrsA | CTL0547 | −1.77 | 1.00E−2 | Phosphoglucomutase | Other | A0A0H3MC40 |
ftsY | CTL0192 | −1.82 | 4.26E−4 | Signal recognition particle receptor | Translation | A0A0H3MGE5 |
CTL0609 | CTL0609 | −1.85 | 7.29E−6 | Hypothetical exported protein | Hypothetical | A0A0H3MDF7 |
dnaX_1 | CTL0439 | −1.92 | 5.15E−3 | DNA polymerase III subunit gamma/tau | DNA replication and repair | A0A0H3MBM8 |
CTL0314 | CTL0314 | −2.04 | 2.67E−4 | Hypothetical membrane protein | Hypothetical | A0A0H3MGM7 |
CTL0430 | CTL0430 | −3.85 | 5.46E−5 | Hypothetical integral membrane protein | Hypothetical | A0A0H3MBV0 |
CTL0063 | CTL0063 | −3.89 | 2.61E−3 | Hypothetical | Hypothetical | A0A0H3MCG4 |
FDR-corrected P values can be found in Table S5. These data were exported from CLC Genomics Workbench 9.5.3. rRNAs, tRNA, and features (genes) with fewer than 10 reads in all samples were eliminated from the data set prior to normalization and EDGE analysis. These data include only genes that were differentially expressed with a significance P value of ≤0.01.
Upregulation of trpA transcription during early-cycle iron starvation was confirmed by RT-qPCR, while only modest increases were observed for the other upregulated genes tested (Fig. 6A). Downregulation of CTL0430, CTL0063, and incD during iron starvation could not be confirmed by RT-qPCR (Fig. 6B). We reasoned that early-cycle responses were not detected by RT-qPCR for most of our tested genes due to the limit of detection of the technique. The raw values detected for most of our early-cycle transcripts fell at or below the lowest concentrations in our standard curves. Between 6 and 9 h postinfection, chlamydial mRNA represents a very small proportion of the total RNA. This limitation was overcome for RNA-seq experiments by depleting rRNAs and eukaryotic RNA prior to synthesizing cDNA. However, cDNA used in RT-qPCR was prepared from total RNA because mRNA enrichment would have made it impossible for us to normalize our RT-qPCR data to chlamydial genomes. The overwhelming proportion of eukaryotic RNA present in the undiluted cDNA used as the template may have impeded accurate detection of transcripts.
Network and biological pathway analysis.
To further analyze the relevance of these gene expression changes to chlamydial survival, we utilized the bioinformatics tool STRING-db v.10.5 to generate networks of functionally associated genes (51). The representation of differentially expressed gene sets (P value, ≤0.05) corresponding to short-term iron starvation (3 h) reveals gene networks with intersecting pathway clusters (see the manually added gray circles). We chose to use the less stringent P value to allow entire pathways to emerge (the pathways would not have been quite as apparent with a more stringent cutoff value). Consistent with our predicted functional categories, network analysis of both early (Fig. 7A) and midcycle (Fig. 7B) responses to iron starvation revealed clusters that include amino acid biosynthesis, DNA replication and repair, and translation. Functional clustering of the midcycle response also revealed the entire cluster of genes necessary to convert pyruvate to acetyl coenzyme A (acetyl-CoA), as well as gene clusters involved in tRNA modification and charging.
The locus identifiers of genes in each identified cluster were submitted to KEGGMapper v2.8 to determine possible roles in specific biological pathways (52). For example, genes from the early-cycle DNA replication and repair cluster (Fig. 7A) were mapped to multiple pathways, including the purine metabolism (5), pyrimidine metabolism (5), mismatch repair (5), replication (4), homologous recombination (3), double-stranded break repair (2), and base excision repair (1) pathways.
Nucleotide metabolism.
We modified the KEGGMapper output for pyrimidine metabolism to indicate the direction of change in midcycle gene expression during iron starvation (Fig. 7C). Under all iron starvation conditions, ribonucleotide reductase gene nrdA was upregulated. Ribonucleotide diphosphates (NDPs) bound to NrdA are converted by NrdB to deoxynucleotide diphosphates (dNDPs). These dNDPs are not likely further converted to deoxynucleotide triphosphates (dNDPs), as indicated by the downregulation of the ndk nucleotide diphosphate kinase gene. Available dUMP would likely be derived from the UDP pool, instead of from the dUTP pool, since transcription of the dUTP pyrophosphatase gene, dut, is downregulated during iron starvation. Taking the data together, these transcriptional changes would result in a net increase in levels of dNDPs, enabling rapid DNA replication when iron levels and ndk expression return to normal (Fig. 7C).
Amino acid biosynthesis.
Functional clustering also indicates that Chlamydia prioritizes maintenance of amino acid pools during iron starvation. Multiple amino acid synthesis, interconversion, and uptake mechanisms were upregulated in response to short-term iron starvation. Transcriptional upregulation of the branched-chain amino acid transporter gene brnQ, the aspartate aminotransferase gene aspC, and the serine hydroxymethyltransferase gene glyA may increase the diversity of the amino acid pool such that Chlamydia can quickly adapt to fluctuations in amino acids. Surprisingly, the tryptophan salvage pathway genes, trpB and trpA, were consistently upregulated during short-term iron starvation. Tryptophan synthase subunit TrpB catalyzes the beta-replacement of indole with serine to form tryptophan (Trp), while TrpA facilitates the interaction of TrpB with indole (53). Their role in recovery from IFN-γ and Trp starvation stresses is well documented, but differential regulation in response to iron starvation is novel (54, 55). While the biological relevance of trpBA induction during iron starvation is unclear, we reason that Chlamydia could in fact prepare for further immune insult (e.g., IFN-γ induction of indoleamine 2,3-dioxygenase expression) by increasing intracellular Trp levels. Taking the data together, iron starvation may increase levels of serine, aspartate, glutamate, branched-chain amino acids, and tryptophan, many of which are essential for normal development (56–60). Amino acid biosynthetic genes were significantly overrepresented (4.38-fold; P value = 0.0464) in the set of differentially expressed midcycle genes as determined by the PANTHER overexpression test (61).
Translation.
The largest cluster generated from STRING-db included translation factors of the midcycle response (Fig. 7B). Based on protein annotations in Uniprot and Biocyc databases, it is evident that C. trachomatis responds to iron starvation by shutting down factors involved in every step of translation: ribosome assembly, initiation, elongation, termination, ribosome recycling, and peptide targeting (50, 90) (Table 4). While preventing the assembly and function of translational machinery, Chlamydia also responds to iron starvation by increasing the levels of factors important for synthesis and modification of tRNAs, in addition to increasing transcription of rnC, the product of which cleaves rRNA transcripts into ribosomal subunit precursors (Table 4). Translation genes were significantly overrepresented (3.26-fold, P value = 0.0243) in the set of midcycle differentially expressed genes as determined by the PANTHER overexpression test (61).
TABLE 4 .
Genea | Expression change |
Locus tag |
Protein ID |
Annotation | Interacting component |
Interaction function | Expression change (6 + 3 h BPDL) |
---|---|---|---|---|---|---|---|
tRNA processing |
|||||||
rnpA_1 | Decrease | CTL0153 | RnP1 | RNase P protein component |
tRNA | Cleaves 5′ end of pre-tRNA | |
rnpA_2 | Decrease | CTL0153A | RnP2 | RNase P protein component |
tRNA | Cleaves 5′ end of pre-tRNA | |
tRNA biogenesis |
|||||||
cysS | Increase | CTL0151 | CysS | Cysteine-tRNA ligase | tRNA (Cys) | Charges tRNA with cysteine | |
pheT | Increase | CTL0736 | PheT | Phenylalanine-tRNA ligase beta |
tRNA (Phe) | Charges tRNA with phenylalanine | |
glyQ | Increase | CTL0165 | GlyQ | Glycine-tRNA ligase alpha subunit |
tRNA (Gly) | Charges tRNA with glycine | |
aspS | Increase | CTL0804 | AspS | Aspartate-tRNA ligase | tRNA (Asp) | Charges tRNA with aspartate | Increase |
thrS | Increase | CTL0844 | ThrS | Threonine-tRNA ligase | tRNA (Thr) | Charges tRNA with threonine | Increase |
truA | Increase | CTL0723 | TruA | tRNA pseudouridine synthase A |
tRNA-anticodon loop |
Converts uridines at 38, 39, and 40 to pseudouridine |
|
miaA | Increase | CTL0135 | MiaA | tRNA dimethylallyltransferase |
tRNA-anticodon loop |
Converts adenine (37) to N6- (dimethylallyl)adenosine |
|
rRNA processing |
|||||||
rnc | Increase | CTL0549 | Rnc | RNase III | 30S transcript | Cleaves 30S precursor transcript to 16S and 23S |
Increase |
Subunit assembly |
|||||||
rpsO | Decrease | CTL0215 | S15 | Ribosomal protein | 16S rRNA | Assembly of 30S subunit | |
rpsT | Decrease | CTL0881 | S20 | Ribosomal protein | 16S rRNA | Assembly of 30S subunit | |
rpsK | Decrease | CTL0770 | S11 | Ribosomal protein | 30S subunit | Forms Shine-Dalgarno cleft | |
rplT | Decrease | CTL0207 | L20 | Ribosomal protein | 23S rRNA | Assembly of 50S subunit | |
rplN | Decrease | CTL0780 | L14 | Ribosomal protein | 23S rRNA | Forms bridge between 30S and 50S | |
rplW | Decrease | CTL0788 | L23 | Ribosomal protein | 23S rRNA | Forms docking site for trigger factor | Decrease |
Initiation | |||||||
infA2 | Decrease | CTL0575 | IF-1 | Initiation factor | 30S-RpsA | Recruited to 30S by RpsA | |
Decrease | IF-1 | Initiation factor | IF-3 | IF-1 and IF-3 recruit IF-2 to 30S, IF-2 recruits mRNA and tRNA |
|||
rplN | Decrease | CTL0780 | L14 | Ribosomal protein | 23S rRNA | Forms bridge between 30S and 50S | Decrease |
rsfS | Decrease | CTL0138 | RsfS | Ribosomal silencing factor | RplN | Inhibits 70S assembly | |
rplL | Decrease | CTL0568 | L7/L12 | Ribosomal protein | GTPases | Binds GTPases required for IF-3 recruitment |
|
Elongation | |||||||
rplL | Decrease | CTL0568 | L7/L12 | Ribosomal protein | GTPases | Binds GTPases required for EF-Tu and EF-G recruitment |
|
Termination | |||||||
prfA | Decrease | CTL0278 | RF-1 | Ribosome release factor | Increases termination at UAA and UAG stop codons |
Decrease | |
rplL | Decrease | CTL0568 | L7/L12 | Ribosomal protein | GTPases | GTPase activity required for RF-3 recruitment |
|
Recycling | |||||||
rrf | Decrease | CTL0046 | RrF | Ribosome recycling factor | Causes disassembly of stalled ribosomes |
||
CTL0791 | Increase | CTL0634 | HflX | GTPase HflX | 50S subunit | Binds to E-site of 70S, disassembles ribosome |
|
Nascent peptide folding/ targeting |
|||||||
ftsY | Decrease | CTL0192 | FtsY | Signal recognition particle receptor |
SRP-RNC | Targets nascent membrane proteins to Sec translocase |
|
secG | Decrease | CTL0606 | SecG | Protein export membrane protein |
SecY | Forms SecYEG translocation channel | |
smpB | Decrease | CTL0332 | SmpB | SsrA-binding protein | tmRNA | Guides tmRNA into tRNA A site, rescuing stalled ribosomes and tagging nascent peptides for degradation |
|
Unknown function |
|||||||
rpmE | Decrease | CTL0277 | L31 | Ribosomal protein | 23S rRNA | Unknown | |
rplQ | Decrease | CTL0768 | L15 | Ribosomal protein | 23S rRNA | Unknown | |
rplY | Decrease | CTL0168 | L25 | Ribosomal protein | 5S rRNA | Binds to 5S in central protuberance |
These genes were shown to be differentially regulated by EdgeR analysis in CLC Genomics Workbench with a P value of ≤0.05 during midcycle iron starvation (12 + 3 BPDL versus 15 h). Annotations and functions were retrieved from the UniProt and BioCyc databases.
Acetyl-CoA synthesis.
Transcription of the entire set of genes necessary for conversion of pyruvate to acetyl-CoA was induced during the midcycle response to BPDL treatment (Fig. 7D). This set includes the lipoylation enzyme genes lipA and lpdA and the genes corresponding to the entire pyruvate dehydrogenase complex, pdhABC. In addition, transcription of the tricarboxylic acid (TCA) cycle gene mdhC and the glycolysis gene eno was induced, likely driving formation of pyruvate from different carbon sources. Acetyl-CoA can be converted to malonyl-CoA for fatty acid biosynthesis or utilized in the formation of N-acetylglucosamine-1-phosphate for peptidoglycan synthesis, both of which are required for rapid growth of Chlamydia (62, 63). Since the levels of transcription of the peptidoglycan-modifying enzymes encoded by glmS and murB were also increased during iron starvation, acetyl-CoA is likely used to form new peptidoglycan. Expression of fatty acid synthesis genes was unchanged during iron starvation.
Pathway analysis of both early and midcycle responses to iron starvation (Fig. 7A and B) revealed that downregulation of translation and upregulation of amino acid synthesis and nucleotide synthesis may be important for surviving this stress. Similarly, a core set of 13 genes (trpB, trpA, nrdA, recA, dnaQ, CTL0704, CTL0102, thrS, prfA, rplW, CTL0061, CTL0548, and pGP8-D) showed differential expression after 3 h of BPDL treatment, during both the early and midcycle responses, while trpB, trpA, and nrdA were upregulated in all BPDL treatments (Fig. S3). This overlap in differential gene expression data is displayed as a Venn diagram in Fig. S3.
DISCUSSION
We monitored the immediate global transcriptional response of Chlamydia trachomatis serovar L2 to short-term iron starvation during early and midcycle (RB-phase) development. In contrast to previous studies of iron starvation in Chlamydia, our short-term treatment performed with BPDL did not cause the hallmark changes in morphology and euo transcription associated with persistence. This approach enabled us to detect a response specific to iron starvation as Chlamydia tries to adapt to stress, rather than detect the transcriptome of the aberrant bacterium. By deep RNA sequencing, we were able to identify novel primary transcriptional responses, representing 7% to 8% of the genome, after only 3 h of iron starvation with BPDL. It is possible that a portion of the detected BPDL-responsive regulon was actually due to chelation of metals other than iron. Cu2+ is chelated at affinities similar to those seen with Fe2+ and Fe3+, while Zn2+ is chelated at a level of affinity 2 to 3 logs lower than that seen with iron ions (64). We suspect that Zn2+ was not efficiently depleted during the short-term BPDL treatments used in this study but cannot exclude the possibility that we had detected transcriptional changes that represent responses to altered availability of other metals. It is also possible that a more immediate response could be detected with even shorter-term BPDL treatments, though we expect a longer duration is required to chelate both free iron and iron bound to protein complexes in intracellular Chlamydia. Since only 12 genes were differentially expressed after 6 h of BPDL treatment, a longer duration of treatment may be necessary to detect the full secondary response, which may not be obvious until the effects of the primary transcriptional response are realized at the protein level. This conjecture is supported by the fact that 6 h of BPDL treatment maintains induction of the primary response operons, trpBA and nrdAB, while reducing or delaying expression of some late-cycle genes (scc2 CTL0840 copB copD, tsp). Decreased or delayed late gene expression has also been observed during long-term iron starvation (43, 49, 65, 66).
In agreement with proteomic observations of deferoxamine-treated C. trachomatis after 24 h and C. pneumoniae after 48 h postinfection, we observed upregulation of CTL0874 (CADD gene), ahpC, eno, and htrA during short-term BPDL treatment (45, 67). In contrast to previous iron starvation studies, we did not detect a significant increase in ytgA expression in our RNA-seq analyses. We expected the ytgABCD iron acquisition operon to be induced immediately in response to iron starvation, since its repression by YtgR is dependent on the presence of available iron (42). Expression of the ytgABCD operon peaks during midcycle development, indicating that the iron-dependent repressor YtgR may be inactive or present at low levels during the early cycle and midcycle (15, 42). It is possible that we did not observe significant differences in the expression of the operon during iron starvation because it was already maximally expressed in the mock-treated controls (see Table S3 in the supplemental material). Global detection of YtgR repression by chromatin immunoprecipitation (ChIP) sequencing or targeted analysis of specific promoters will be necessary to delineate the contribution of YtgR activity to that of the detected iron-responsive regulon. Recent work to define targets of known transcription factors in Waddlia chondophila discovered binding sites of YtgC by ChIP sequencing (68). Interestingly, the level of the most frequent target, hrcA, was also increased during iron starvation in our study, indicating that it may also be a target of YtgC in C. trachomatis.
Transcriptional responses to iron starvation in most bacteria typically include upregulation of iron acquisition systems and virulence factors (69–71). While expression of the ytgABCD operon was not upregulated during short-term iron starvation, other unidentified iron uptake and iron-dependent repression mechanisms may exist and thus could be represented in our set of iron starvation-induced genes. The virulence factor CADD (CTL0874) gene and the MACPF (CTL0408) gene were induced in both the early and midcycle responses to iron starvation. CADD overexpression induces apoptosis under conditions of expression in cultured human epithelial cells but has also been demonstrated to play a role in folate biosynthesis (72, 73). MACPF contains a domain that may enable perforin activity but so far has only been shown to undergo cleavage upon infection and become inserted into bacterial membranes (74). Several type III secretion structural components (mcsC, sctJ, sctR, fliF, cdsN [CTL0043], and cdsD [CTL0033]) and effectors (CTL0884, CTL0476, CTL0184, and CTL0081) were also transcriptionally upregulated during iron starvation, which could potentially alter interactions between the host and chlamydial inclusion.
Similarly to the upregulation of the ribonucleotide reductase operon nrdHIEF seen during iron starvation in Escherichia coli and Yersinia pestis, the ribonucleotide nrdA and nrdB reductase genes are consistently upregulated during under short-term iron starvation (70, 75). This upregulation indicates that deoxynucleotides may be important for Chlamydia to survive this stress. However, since NrdB requires iron for its function, deoxynucleotide levels may not increase until iron becomes available. Instead, high levels of inactive NrdA-B complexes may actually impede replication and development by inducing stalling at replication forks, providing a possible explanation for the decreased replication observed during iron starvation (76, 77).
The immediate transcriptional response of Chlamydia to iron starvation is remarkably similar to the stringent responses seen in other bacteria, which enable rapid adaptation to various stresses by diverting resources from macromolecular biosynthesis, e.g., translation, and from growth to immediate survival, often resulting in a quiescent state (78, 79). This rapid transcriptional response is achieved through synthesis of the chemical alarmone (p)ppGpp, which interacts with RNA polymerase and DksA to globally modify transcriptional activity (80, 81). During amino acid starvation in bacteria, uncharged tRNAs in the A-site of ribosomes are sensed by RelA, which responds by synthesizing (p)ppGpp from ATP and GDP or GTP (82, 83). (p)ppGpp can also be synthesized and hydrolyzed by SpoT under other stress conditions. However, since Chlamydia lacks the RelA and SpoT homologues necessary for (p)ppGpp synthesis, it likely evolved alternative mechanisms to reduce growth and increase survival responses during stress (17, 84, 85). Iron starvation has been shown to induce a stringent response in Bacillus subtilis that upregulates transcription of amino acid biosynthesis genes (86).
Multiple amino acid synthesis, interconversion, and uptake mechanisms were upregulated in response to short-term iron starvation. Surprisingly, the primary response included an increase in expression of transcripts involved in tryptophan salvage, trpB and trpA, but not in expression of the tryptophan-dependent repressor trpR gene. TrpR-dependent regulation of the polycistronic transcript trpRBA has been extensively studied during tryptophan starvation and IFN-γ treatment but rarely, if ever, in the context of iron starvation (54, 87, 88). Notably, trpB levels, but not trpR levels, were also increased under conditions of estradiol-induced persistence, suggesting that a trpR-independent mechanism for inducing tryptophan salvage transcription may exist (89).
Pathway analysis clearly indicates that transcripts involved in all steps of translation from initiation to ribosome recycling are downregulated during iron starvation. This reduction in translation factors might lead to an eventual shutdown or modification of translation activity that could increase survival during stress. By shutting down energy-expensive protein synthesis, ATP and GTP pools can be rerouted to immediate survival responses (tRNA charging, transcription). Similarly, iron starvation reduces the transcription of several ABC transporter genes which require ATP for their function. Uncoupled RNA and protein levels in Chlamydia have also been observed during IFN-γ stress (17). The apparent decrease in translation during IFN-γ exposure could be exacerbated by decreases in the levels of components of the translation machinery in response to simultaneous iron starvation. However, decreased expression of translation factors during the primary response to iron starvation may not be apparent until preexisting ribosome-protein complexes are degraded or destabilized. This may explain why ≥24 h of iron starvation is required to induce the development of aberrant RBs (43). Downregulation of translation factors during iron starvation will have to be examined at the protein level to determine its contribution to adaptation to iron starvation and development of persistence.
In contrast to downregulation of translation, iron starvation increases transcription of amino-acyl synthesis genes (cysS, pheT, glyQ, aspS, thrS), which are responsible for charging tRNAs with amino acids. The apparent disconnect between increased levels of aminoacyl-tRNA pools and decreased translation indicates possible survival mechanisms. Charged tRNAs might be utilized in an immediate survival response to iron starvation, prior to the turnover of ribosomal subunits. Alternatively, Chlamydia might accumulate charged tRNAs for recovery and resumption of development when normal levels of iron and translation factors are restored.
A major theme that emerged from our gene expression analysis is that Chlamydia likely perceives iron starvation as a signal to prepare for further nutrient deprivation and immune insult. Transcriptional upregulations of tryptophan salvage pathway (trpB, trpA), oxidative stress (ahpC, pdi, and sodM), and DNA repair (mutS, mutL, ssb, ung, recA) genes indicate a protective response to antimicrobial insults of the inflammatory immune response (e.g., IDO activation, reactive oxygen species). As an obligate intracellular pathogen, Chlamydia has undergone reductive evolution with constant selective pressure from the host immune system and its multiple antichlamydial effectors. Due to its small (~1-Mbp) genome, Chlamydia may not have the capability to induce a specific transcriptional response to each particular stressor, and the simultaneous deployment of stress responses may have been the most parsimonious route of adaptation to immune insult. In this case, we would expect that iron-starved Chlamydia would be better protected from damage by antimicrobial insults than mock-treated Chlamydia. Immediate transcriptional responses to other stress conditions will need to be monitored to determine if this coordination of antimicrobial responses is unique to iron starvation.
This report provides the first evidence of a global iron-dependent regulon for C. trachomatis. By using a system approach to delineate Chlamydia’s transcriptional response to iron starvation, we have been able to detect biological pathways and place them in the context of chlamydial development. These findings are novel and add to previous studies of iron-dependent transcriptional and proteomic profiling in aberrant RBs, revealing transcriptional adaptive strategies prior to the development of a persistent state. Additionally, our results include a high-resolution profile of midcycle development of C. trachomatis serovar L2, including relevant time points for monitoring shifts in gene expression in the early, middle, and late cycles. We expect that this data set will prove useful for future studies that seek to determine the immediate transcriptional response of Chlamydia to other chemical and/or nutrient stresses. Our findings include previously unrecognized shifts in energy utilization and downregulation of translation that resemble a stringency-like survival response. Chlamydia may utilize a two-stage approach of increasing transcription of survival genes in the short term to delay development and survive during iron starvation, followed by an eventual shutdown of translation at later times of sustained stress. The latter might account for the observed irreversibility of the persistent state during long-term starvation for iron or tryptophan.
MATERIALS AND METHODS
Cell culture and infection.
HeLa monolayers were infected with C. trachomatis strain L2 434/Bu in 6-well plates at a multiplicity of infection (MOI) of 2 for RNA and genomic DNA (gDNA) collection experiments and on coverslips in 24-well plates for morphology studies. Cells were grown in Dulbecco’s modified Eagle’s medium (DMEM) supplemented with 10% fetal bovine serum (FBS), 2 mM glutamine, and 10 µg/ml gentamycin in 5% CO2 at 37°C. HeLa cells used in this study were started from P1 stocks from ATCC and were regularly checked for contamination by DAPI (4′,6-diamidino-2-phenylindole) staining and the use of a Universal Mycoplasma Detection kit (ATCC).
RNA sequencing.
RNA was collected and pooled from 2 or 4 T75 flasks of C. trachomatis-infected HeLa monolayers that had been treated with 100 µM 2,2-bipyridyl (BPDL) starting at 6 or 12 h postinfection (6 h + 3 h BPDL, 12 h + 3 h BPDL, 12 h + 6 h BPDL) and from mock-treated samples at equivalent time points postinfection (9 h, 12 h, 15 h, 18 h). RNA was purified using a RiboPure Bacteria (Ambion) kit per the instructions of the manufacturer. Total RNA was further enriched for transcripts over 100 nucleotides in length by the use of a MegaClear kit (Ambion). Mammalian transcripts and rRNAs were removed using a MicrobEnrich kit (Ambion), and bacterial rRNAs were removed using a MicrobExpress kit (Ambion), repeating 2 to 3 times. The integrity and quantity of total and depleted RNA were monitored with an AATI fragment analyzer. cDNA libraries were prepared with Ion Total RNA-seq kit V2, sequencing beads were prepared using an Ion Chef system, and sequencing was performed on an Ion Proton chip with HiQ chemistry. Primary sequence analysis and trimming and binning of reads were performed using Torrent Suite Software version 5.0.5. Remaining reads were mapped to the combined core genome of C. trachomatis strain L2 434/Bu (GenBank accession no. AM884176) and the plasmid of C. trachomatis L2b CS784/08 (NZ_CP009926) using CLC Genomics Workbench 9, requiring reads be at least 30 nucleotides in length, with default alignment parameters.
The EdgeR algorithm in CLC Genomics was used to determine differential gene expression levels during development and iron starvation, assuming a false-discovery rate of 10% and P values of ≤0.05. tRNAs and ribosomal RNAs were filtered from the reads to account for differences in depletion efficiency, and only genes with at least 5 mapped reads were included in the analysis. Differentially expressed genes were confirmed for selected transcripts by RT-qPCR.
qPCR and RT-qPCR.
C. trachomatis-infected HeLa monolayers were treated with 100 µM BPDL starting at 6 or 12 h hours postinfection (6 h + 3 h BPDL, 12 h + 3 h BPDL, 12 h + 6 h BPDL) and mock-treated samples at equivalent time points postinfection (6 h, 9 h, 12 h, 15 h, 18 h). RNA and gDNA were collected with RiboPure Bacteria and the DNeasy Blood and Tissue (Qiagen) kits, respectively. cDNA was generated with Superscript IV reverse transcriptase (Life Technologies, Inc.) using 200 to 500 ng RNA per the instructions of the manufacturer, except with the use of random nonamers instead of hexamers. Transcripts were amplified with a PowerUp SYBR green system from undiluted cDNA for early-cycle samples (6 to 9 h) or diluted 1:10 in 10 mM Tris for midcycle samples (12 to 18 h) and detected with an Applied Biosystems 7300 RT-qPCR system.
Chlamydial morphology.
Chlamydiae were monitored for 3, 6, or 12 h for changes in morphology in response to mock treatment or treatment with 100 µM BPDL starting at 12 h postinfection. Infected cultures were fixed on coverslips and stained with pooled human serum (Sigma; H4522) at 1:750 followed by goat anti-human antibody conjugated to Alexa Fluor 488 (Thermo Fisher) at 1:1,000. DNA was stained with DAPI at 5 µg/ml. Images were taken on a Leica SP8 confocal microscope with a 63× oil objective and 4× zoom.
IFU assay.
Chlamydiae were monitored starting at 12 h postinfection for 12 or 24 h for changes in infectivity in response to mock treatment or treatment with 100 µM BPDL at an MOI of 1. Infected cultures were scraped into 300 μl SPG (succinic acid, sodium dihydrogen phosphate, glycine) and stored at −80 C for later testing. Thawed lysates were serially diluted into complete DMEM, centrifuged onto HeLa monolayers in 24-well plates, washed with Hanks balanced salt solution (HBSS), and allowed to infect for 24 h. Infected cultures were fixed and stained with pooled human serum at 1:750 followed by goat anti-human antibody conjugated to Alexa Fluor 488 (Thermo Fisher) at 1:1,000. Inclusions were counted by fluorescence microscopy, and levels of inclusion-forming units (IFU) were calculated as previously described.
Visual analysis of differentially expressed genes.
Functional categories were assigned for all genes differentially regulated with a P value of ≤0.01 by referring to the GO terms listed on UniProt. Pie charts were generated using the “pie” function in Rstudio. Heat maps were generated in Rstudio using the package “pheatmaps,” with parameters set to average clustering and Euclidean distance. The PANTHER overexpression test was done in PANTHER v12.0 on differentially regulated gene sets (total) with P values of ≤0.01, using the default parameters and Bonferroni correction. Pathway analysis was performed on differentially regulated genesets with P values of ≤0.05 and a minimum of 10 mapped reads, with STRING-db v.10.5 set to a confidence value ≥0.7. StringDB maps were slightly modified to make space to increase font size, to indicate the direction of change by color coding, and to add pathway labels without altering network relationships. Clustered genes detected with StringDB were further analyzed using KeggMapper v.2.8, and pathway maps were generated based on KeggMapper output using Affinity Designer v1.4.1.
Data availability.
Raw and processed sequencing files were submitted to the NCBI Gene Expression Omnibus (GEO) as a Superseries, and the midcycle and early-cycle projects can be found using accession number GSE106763.
ACKNOWLEDGMENTS
We acknowledge Scot Ouellette and Nicholas Pokorzynski for critical reading of the manuscript.
This work was funded by startup funds for R.A.C. from the School of Molecular Biosciences, College of Veterinary Medicine, Washington State University; by NIH grant 2R01AI065545-06A1 awarded to R.A.C.; and by NIH training grant for A.J.B.
REFERENCES
- 1.CDC 2016. Sexually transmitted disease surveillance 2015. Department of Health and Human Services, Atlanta, GA. [Google Scholar]
- 2.Price MJ, Ades AE, Welton NJ, Simms I, Macleod J, Horner PJ. 2016. Proportion of pelvic inflammatory disease cases caused by Chlamydia trachomatis: consistent picture from different methods. J Infect Dis 214:617–624. doi: 10.1093/infdis/jiw178. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Stoner BP, Cohen SE. 2015. Lymphogranuloma venereum 2015: clinical presentation, diagnosis, and treatment, Table 1. Clin Infect Dis 61:S865–SS873. doi: 10.1093/cid/civ756. [DOI] [PubMed] [Google Scholar]
- 4.Kjaer HO, Dimcevski G, Hoff G, Olesen F, Ostergaard L. 2000. Recurrence of urogenital Chlamydia trachomatis infection evaluated by mailed samples obtained at home: 24 weeks’ prospective follow up study. Sex Transm Infect 76:169–172. doi: 10.1136/sti.76.3.169. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Batteiger BE, Tu W, Ofner S, Van Der Pol B, Stothard DR, Orr DP, Katz BP, Fortenberry JD. 2010. Repeated Chlamydia trachomatis genital infections in adolescent women. J Infect Dis 201:42–51. doi: 10.1086/648734. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Wyrick PB. 2010. Chlamydia trachomatis persistence in vitro: an overview. J Infect Dis 201:S88–S95. doi: 10.1086/652394. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Lewis ME, Belland RJ, AbdelRahman YM, Beatty WL, Aiyar AA, Zea AH, Greene SJ, Marrero L, Buckner LR, Tate DJ, McGowin CL, Kozlowski PA, O’Brien M, Lillis RA, Martin DH, Quayle AJ. 2014. Morphologic and molecular evaluation of Chlamydia trachomatis growth in human endocervix reveals distinct growth patterns. Front Cell Infect Microbiol 4:71. doi: 10.3389/fcimb.2014.00071. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.AbdelRahman YM, Belland RJ. 2005. The chlamydial developmental cycle. FEMS Microbiol Rev 29:949–959. doi: 10.1016/j.femsre.2005.03.002. [DOI] [PubMed] [Google Scholar]
- 9.Matsumoto A, Manire GP. 1970. Electron microscopic observations on the effects of penicillin on the morphology of Chlamydia psittaci. J Bacteriol 101:278–285. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Kazar J, Gillmore JD, Gordon FB. 1971. Effect of interferon and interferon inducers on infections with a nonviral intracellular microorganism, Chlamydia trachomatis. Infect Immun 3:825–832. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Raulston JE. 1997. Response of Chlamydia trachomatis serovar E to iron restriction in vitro and evidence for iron-regulated chlamydial proteins. Infect Immun 65:4539–4547. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Byrne GI, Lehmann LK, Landry GJ. 1986. Induction of tryptophan catabolism is the mechanism for gamma-interferon-mediated inhibition of intracellular Chlamydia psittaci replication in T24 cells. Infect Immun 53:347–351. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Beatty WL, Byrne GI, Morrison RP. 1993. Morphologic and antigenic characterization of interferon gamma-mediated persistent Chlamydia trachomatis infection in vitro. Proc Natl Acad Sci U S A 90:3998–4002. doi: 10.1073/pnas.90.9.3998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Beatty WL, Morrison RP, Byrne GI. 1995. Reactivation of persistent Chlamydia trachomatis infection in cell culture. Infect Immun 63:199–205. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Belland RJ, Nelson DE, Virok D, Crane DD, Hogan D, Sturdevant D, Beatty WL, Caldwell HD. 2003. Transcriptome analysis of chlamydial growth during IFN-gamma-mediated persistence and reactivation. Proc Natl Acad Sci U S A 100:15971–15976. doi: 10.1073/pnas.2535394100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Reveneau N, Crane DD, Fischer E, Caldwell HD. 2005. Bactericidal activity of first-choice antibiotics against gamma interferon-induced persistent infection of human epithelial cells by Chlamydia trachomatis. Antimicrob Agents Chemother 49:1787–1793. doi: 10.1128/AAC.49.5.1787-1793.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Ouellette SP, Hatch TP, AbdelRahman YM, Rose LA, Belland RJ, Byrne GI. 2006. Global transcriptional upregulation in the absence of increased translation in Chlamydia during IFNgamma-mediated host cell tryptophan starvation. Mol Microbiol 62:1387–1401. doi: 10.1111/j.1365-2958.2006.05465.x. [DOI] [PubMed] [Google Scholar]
- 18.Stephens RS, Kalman S, Lammel C, Fan J, Marathe R, Aravind L, Mitchell W, Olinger L, Tatusov RL, Zhao Q, Koonin EV, Davis RW. 1998. Genome sequence of an obligate intracellular pathogen of humans: Chlamydia trachomatis. Science 282:754–759. doi: 10.1126/science.282.5389.754. [DOI] [PubMed] [Google Scholar]
- 19.Chen W. 2011. IDO: more than an enzyme. Nat Immunol 12:809–811. doi: 10.1038/ni.2088. [DOI] [PubMed] [Google Scholar]
- 20.Byrd TF, Horwitz MA. 1989. Interferon gamma-activated human monocytes downregulate transferrin receptors and inhibit the intracellular multiplication of Legionella pneumophila by limiting the availability of iron. J Clin Invest 83:1457–1465. doi: 10.1172/JCI114038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Nemeth E, Valore EV, Territo M, Schiller G, Lichtenstein A, Ganz T. 2003. Hepcidin, a putative mediator of anemia of inflammation, is a type II acute-phase protein. Blood 101:2461–2463. doi: 10.1182/blood-2002-10-3235. [DOI] [PubMed] [Google Scholar]
- 22.Ludwiczek S, Aigner E, Theurl I, Weiss G. 2003. Cytokine-mediated regulation of iron transport in human monocytic cells. Blood 101:4148–4154. doi: 10.1182/blood-2002-08-2459. [DOI] [PubMed] [Google Scholar]
- 23.Mpiga P, Ravaoarinoro M. 2006. Chlamydia trachomatis persistence: an update. Microbiol Res 161:9–19. doi: 10.1016/j.micres.2005.04.004. [DOI] [PubMed] [Google Scholar]
- 24.Theurl I, Theurl M, Seifert M, Mair S, Nairz M, Rumpold H, Zoller H, Bellmann-Weiler R, Niederegger H, Talasz H, Weiss G. 2008. Autocrine formation of hepcidin induces iron retention in human monocytes. Blood 111:2392–2399. doi: 10.1182/blood-2007-05-090019. [DOI] [PubMed] [Google Scholar]
- 25.Paradkar PN, De Domenico I, Durchfort N, Zohn I, Kaplan J, Ward DM. 2008. Iron depletion limits intracellular bacterial growth in macrophages. Blood 112:866–874. doi: 10.1182/blood-2007-12-126854. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Nairz M, Haschka D, Demetz E, Weiss G. 2014. Iron at the interface of immunity and infection. Front Pharmacol 5:152. doi: 10.3389/fphar.2014.00152. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Pokorzynski ND, Thompson CC, Carabeo RA. 2017. Ironing out the unconventional mechanisms of iron acquisition and gene regulation in chlamydia. Front Cell Infect Microbiol 7:394. doi: 10.3389/fcimb.2017.00394. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Hentze MW, Muckenthaler MU, Galy B, Camaschella C. 2010. Two to tango: regulation of mammalian iron metabolism. Cell 142:24–38. doi: 10.1016/j.cell.2010.06.028. [DOI] [PubMed] [Google Scholar]
- 29.Vardhan H, Bhengraj AR, Jha R, Singh Mittal A. 2009. Chlamydia trachomatis alters iron-regulatory protein-1 binding capacity and modulates cellular iron homeostasis in HeLa-229 cells. J Biomed Biotechnol 2009:342032. doi: 10.1155/2009/342032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Kim I, Yetley EA, Calvo MS. 1993. Variations in iron-status measures during the menstrual cycle. Am J Clin Nutr 58:705–709. doi: 10.1093/ajcn/58.5.705. [DOI] [PubMed] [Google Scholar]
- 31.Kelver ME, Kaul A, Nowicki B, Findley WE, Hutchens TW, Nagamani M. 1996. Estrogen regulation of lactoferrin expression in human endometrium. Am J Reprod Immunol 36:243–247. doi: 10.1111/j.1600-0897.1996.tb00171.x. [DOI] [PubMed] [Google Scholar]
- 32.Aisen P, Enns C, Wessling-Resnick M. 2001. Chemistry and biology of eukaryotic iron metabolism. Int J Biochem Cell Biol 33:940–959. doi: 10.1016/S1357-2725(01)00063-2. [DOI] [PubMed] [Google Scholar]
- 33.Skaar EP. 2010. The battle for iron between bacterial pathogens and their vertebrate hosts. PLoS Pathog 6:e1000949. doi: 10.1371/journal.ppat.1000949. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Caza M, Kronstad JW. 2013. Shared and distinct mechanisms of iron acquisition by bacterial and fungal pathogens of humans. Front Cell Infect Microbiol 3:80. doi: 10.3389/fcimb.2013.00080. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Hood MI, Skaar EP. 2012. Nutritional immunity: transition metals at the pathogen-host interface. Nat Rev Microbiol 10:525–537. doi: 10.1038/nrmicro2836. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Pan X, Tamilselvam B, Hansen EJ, Daefler S. 2010. Modulation of iron homeostasis in macrophages by bacterial intracellular pathogens. BMC Microbiol 10:64. doi: 10.1186/1471-2180-10-64. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Scidmore MA, Fischer ER, Hackstadt T. 2003. Restricted fusion of Chlamydia trachomatis vesicles with endocytic compartments during the initial stages of infection. Infect Immun 71:973–984. doi: 10.1128/IAI.71.2.973-984.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Ouellette SP, Carabeo RA. 2010. A functional slow recycling pathway of transferrin is required for growth of Chlamydia. Front Microbiol 1:112. doi: 10.3389/fmicb.2010.00112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Boradia VM, Malhotra H, Thakkar JS, Tillu VA, Vuppala B, Patil P, Sheokand N, Sharma P, Chauhan AS, Raje M, Raje CI. 2014. Mycobacterium tuberculosis acquires iron by cell-surface sequestration and internalization of human holo-transferrin. Nat Commun 5:4730. doi: 10.1038/ncomms5730. [DOI] [PubMed] [Google Scholar]
- 40.Raulston JE, Miller JD, Davis CH, Schell M, Baldwin A, Ferguson K, Lane H. 2007. Identification of an iron-responsive protein that is antigenic in patients with Chlamydia trachomatis genital infections. FEMS Immunol Med Microbiol 51:569–576. doi: 10.1111/j.1574-695X.2007.00336.x. [DOI] [PubMed] [Google Scholar]
- 41.Miller JD, Sal MS, Schell M, Whittimore JD, Raulston JE. 2009. Chlamydia trachomatis YtgA is an iron-binding periplasmic protein induced by iron restriction. Microbiology 155:2884–2894. doi: 10.1099/mic.0.030247-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Thompson CC, Nicod SS, Malcolm DS, Grieshaber SS, Carabeo RA. 2012. Cleavage of a putative metal permease in Chlamydia trachomatis yields an iron-dependent transcriptional repressor. Proc Natl Acad Sci U S A 109:10546–10551. doi: 10.1073/pnas.1201398109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Thompson CC, Carabeo RA. 2011. An optimal method of iron starvation of the obligate intracellular pathogen, Chlamydia trachomatis. Front Microbiol 2:20. doi: 10.3389/fmicb.2011.00020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.LaRue RW, Dill BD, Giles DK, Whittimore JD, Raulston JE. 2007. Chlamydial Hsp60-2 is iron responsive in Chlamydia trachomatis serovar E-infected human endometrial epithelial cells in vitro. Infect Immun 75:2374–2380. doi: 10.1128/IAI.01465-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Dill BD, Dessus-Babus S, Raulston JE. 2009. Identification of iron-responsive proteins expressed by Chlamydia trachomatis reticulate bodies during intracellular growth. Microbiology 155:210–219. doi: 10.1099/mic.0.022731-0. [DOI] [PubMed] [Google Scholar]
- 46.Breuer W, Epsztejn S, Cabantchik ZI. 1995. Iron acquired from transferrin by K562 cells is delivered into a cytoplasmic pool of chelatable iron(II). J Biol Chem 270:24209–24215. doi: 10.1074/jbc.270.41.24209. [DOI] [PubMed] [Google Scholar]
- 47.Wilderman PJ, Sowa NA, FitzGerald DJ, FitzGerald PC, Gottesman S, Ochsner UA, Vasil ML. 2004. Identification of tandem duplicate regulatory small RNAs in Pseudomonas aeruginosa involved in iron homeostasis. Proc Natl Acad Sci U S A 101:9792–9797. doi: 10.1073/pnas.0403423101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Klitgaard K, Friis C, Angen O, Boye M. 2010. Comparative profiling of the transcriptional response to iron restriction in six serotypes of Actinobacillus pleuropneumoniae with different virulence potential. BMC Genomics 11:698. doi: 10.1186/1471-2164-11-698. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Timms P, Good D, Wan C, Theodoropoulos C, Mukhopadhyay S, Summersgill J, Mathews S. 2009. Differential transcriptional responses between the interferon-gamma-induction and iron-limitation models of persistence for Chlamydia pneumoniae. J Microbiol Immunol Infect 42:27–37. [PubMed] [Google Scholar]
- 50.UniProt Consortium 2017. UniProt: the universal protein KnowledgeBase. Nucleic Acids Res 45:D158–D169. doi: 10.1093/nar/gkw1099. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Szklarczyk D, Franceschini A, Wyder S, Forslund K, Heller D, Huerta-Cepas J, Simonovic M, Roth A, Santos A, Tsafou KP, Kuhn M, Bork P, Jensen LJ, von Mering C. 2015. STRING v10: protein-protein interaction networks, integrated over the tree of life. Nucleic Acids Res 43:D447–D452. doi: 10.1093/nar/gku1003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Kanehisa M, Goto S, Sato Y, Furumichi M, Tanabe M. 2012. KEGG for integration and interpretation of large-scale molecular data sets. Nucleic Acids Res 40:D109–D114. doi: 10.1093/nar/gkr988. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Fehlner-Gardiner C, Roshick C, Carlson JH, Hughes S, Belland RJ, Caldwell HD, McClarty G. 2002. Molecular basis defining human Chlamydia trachomatis tissue tropism. A possible role for tryptophan synthase. J Biol Chem 277:26893–26903. doi: 10.1074/jbc.M203937200. [DOI] [PubMed] [Google Scholar]
- 54.Wood H, Fehlner-Gardner C, Berry J, Fischer E, Graham B, Hackstadt T, Roshick C, McClarty G. 2003. Regulation of tryptophan synthase gene expression in Chlamydia trachomatis. Mol Microbiol 49:1347–1359. doi: 10.1046/j.1365-2958.2003.03638.x. [DOI] [PubMed] [Google Scholar]
- 55.Caldwell HD, Wood H, Crane D, Bailey R, Jones RB, Mabey D, Maclean I, Mohammed Z, Peeling R, Roshick C, Schachter J, Solomon AW, Stamm WE, Suchland RJ, Taylor L, West SK, Quinn TC, Belland RJ, McClarty G. 2003. Polymorphisms in Chlamydia trachomatis tryptophan synthase genes differentiate between genital and ocular isolates. J Clin Invest 111:1757–1769. doi: 10.1172/JCI17993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Karayiannis P, Hobson D. 1981. Amino acid requirements of a Chlamydia trachomatis genital strain in McCoy cell cultures. J Clin Microbiol 13:427–432. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Al-Younes HM, Gussmann J, Braun PR, Brinkmann V, Meyer TF. 2006. Naturally occurring amino acids differentially influence the development of Chlamydia trachomatis and Chlamydia (Chlamydophila) pneumoniae. J Med Microbiol 55:879–886. doi: 10.1099/jmm.0.46445-0. [DOI] [PubMed] [Google Scholar]
- 58.Braun PR, Al-Younes H, Gussmann J, Klein J, Schneider E, Meyer TF. 2008. Competitive inhibition of amino acid uptake suppresses chlamydial growth: involvement of the chlamydial amino acid transporter BrnQ. J Bacteriol 190:1822–1830. doi: 10.1128/JB.01240-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Kane CD, Vena RM, Ouellette SP, Byrne GI. 1999. Intracellular tryptophan pool sizes may account for differences in gamma interferon-mediated inhibition and persistence of chlamydial growth in polarized and nonpolarized cells. Infect Immun 67:1666–1671. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Leonhardt RM, Lee SJ, Kavathas PB, Cresswell P. 2007. Severe tryptophan starvation blocks onset of conventional persistence and reduces reactivation of Chlamydia trachomatis. Infect Immun 75:5105–5117. doi: 10.1128/IAI.00668-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Mi H, Huang X, Muruganujan A, Tang H, Mills C, Kang D, Thomas PD. 2017. PANTHER version 11: expanded annotation data from Gene Ontology and Reactome pathways, and data analysis tool enhancements. Nucleic Acids Res 45:D183–D189. doi: 10.1093/nar/gkw1138. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Liechti GW, Kuru E, Hall E, Kalinda A, Brun YV, VanNieuwenhze M, Maurelli AT. 2014. A new metabolic cell-wall labelling method reveals peptidoglycan in Chlamydia trachomatis. Nature 506:507–510. doi: 10.1038/nature12892. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Soupene E, Wang D, Kuypers FA. 2015. Remodeling of host phosphatidylcholine by Chlamydia acyltransferase is regulated by acyl-CoA binding protein ACBD6 associated with lipid droplets. Microbiologyopen 4:235. doi: 10.1002/mbo3.234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Smith R, Martell AE. 2002. NIST critically selected stability constants of metal complexes database. NIST Standard Reference Database, 46, version 8.0 https://www.nist.gov/sites/default/files/documents/srd/46_8.pdf.
- 65.Al-Younes HM, Rudel T, Brinkmann V, Szczepek AJ, Meyer TF. 2001. Low iron availability modulates the course of Chlamydia pneumoniae infection. Cell Microbiol 3:427–437. doi: 10.1046/j.1462-5822.2001.00125.x. [DOI] [PubMed] [Google Scholar]
- 66.Mäurer AP, Mehlitz A, Mollenkopf HJ, Meyer TF. 2007. Gene expression profiles of Chlamydophila pneumoniae during the developmental cycle and iron depletion-mediated persistence. PLoS Pathog 3:e83. doi: 10.1371/journal.ppat.0030083. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Mukhopadhyay S, Miller RD, Sullivan ED, Theodoropoulos C, Mathews SA, Timms P, Summersgill JT. 2006. Protein expression profiles of Chlamydia pneumoniae in models of persistence versus those of heat shock stress response. Infect Immun 74:3853–3863. doi: 10.1128/IAI.02104-05. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.de Barsy M, Frandi A, Panis G, Théraulaz L, Pillonel T, Greub G, Viollier PH. 2016. Regulatory (pan-)genome of an obligate intracellular pathogen in the PVC superphylum. ISME J 10:2129–2144. doi: 10.1038/ismej.2016.23. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Brickman TJ, Cummings CA, Liew SY, Relman DA, Armstrong SK. 2011. Transcriptional profiling of the iron starvation response in Bordetella pertussis provides new insights into siderophore utilization and virulence gene expression. J Bacteriol 193:4798–4812. doi: 10.1128/JB.05136-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Zhou D, Qin L, Han Y, Qiu J, Chen Z, Li B, Song Y, Wang J, Guo Z, Zhai J, Du Z, Wang X, Yang R. 2006. Global analysis of iron assimilation and fur regulation in Yersinia pestis. FEMS Microbiol Lett 258:9–17. doi: 10.1111/j.1574-6968.2006.00208.x. [DOI] [PubMed] [Google Scholar]
- 71.Kunkle CA, Schmitt MP. 2003. Analysis of the Corynebacterium diphtheriae DtxR regulon: identification of a putative siderophore synthesis and transport system that is similar to the Yersinia high-pathogenicity island-encoded yersiniabactin synthesis and uptake system. J Bacteriol 185:6826–6840. doi: 10.1128/JB.185.23.6826-6840.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Stenner-Liewen F, Liewen H, Zapata JM, Pawlowski K, Godzik A, Reed JC. 2002. CADD, a Chlamydia protein that interacts with death receptors. J Biol Chem 277:9633–9636. doi: 10.1074/jbc.C100693200. [DOI] [PubMed] [Google Scholar]
- 73.Adams NE, Thiaville JJ, Proestos J, Juárez-Vázquez AL, McCoy AJ, Barona-Gómez F, Iwata-Reuyl D, de Crécy-Lagard V, Maurelli AT. 2014. Promiscuous and adaptable enzymes fill “holes” in the tetrahydrofolate pathway in Chlamydia species. mBio 5:e01378-14. doi: 10.1128/mBio.01378-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Taylor LD, Nelson DE. 2014. Chlamydial MACPF protein CT153. Subcell Biochem 80:255–269. doi: 10.1007/978-94-017-8881-6_13. [DOI] [PubMed] [Google Scholar]
- 75.McHugh JP, Rodríguez-Quinoñes F, Abdul-Tehrani H, Svistunenko DA, Poole RK, Cooper CE, Andrews SC. 2003. Global iron-dependent gene regulation in Escherichia coli. A new mechanism for iron homeostasis. J Biol Chem 278:29478–29486. doi: 10.1074/jbc.M303381200. [DOI] [PubMed] [Google Scholar]
- 76.Rosenkranz HS, Gutter B, Becker Y. 1973. Studies on the developmental cycle of Chlamydia trachomatis: selective inhibition by hydroxyurea. J Bacteriol 115:682–690. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Odsbu I, Morigen S, Skarstad K. 2009. A reduction in ribonucleotide reductase activity slows down the chromosome replication fork but does not change its localization. PLoS One 4:e7617. doi: 10.1371/journal.pone.0007617. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Jain V, Kumar M, Chatterji D. 2006. ppGpp: stringent response and survival. J Microbiol 44:1–10. [PubMed] [Google Scholar]
- 79.Potrykus K, Murphy H, Philippe N, Cashel M. 2011. ppGpp is the major source of growth rate control in E. coli. Environ Microbiol 13:563–575. doi: 10.1111/j.1462-2920.2010.02357.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Perederina A, Svetlov V, Vassylyeva MN, Tahirov TH, Yokoyama S, Artsimovitch I, Vassylyev DG. 2004. Regulation through the secondary channel—structural framework for ppGpp-DksA synergism during transcription. Cell 118:297–309. doi: 10.1016/j.cell.2004.06.030. [DOI] [PubMed] [Google Scholar]
- 81.Cashel M. 1969. The control of ribonucleic acid synthesis in Escherichia coli. IV. Relevance of unusual phosphorylated compounds from amino acid-starved stringent strains. J Biol Chem 244:3133–3141. [PubMed] [Google Scholar]
- 82.Haseltine WA, Block R. 1973. Synthesis of guanosine tetra- and pentaphosphate requires the presence of a codon-specific, uncharged transfer ribonucleic acid in the acceptor site of ribosomes. Proc Natl Acad Sci U S A 70:1564–1568. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Traxler MF, Summers SM, Nguyen HT, Zacharia VM, Hightower GA, Smith JT, Conway T. 2008. The global, ppGpp-mediated stringent response to amino acid starvation in Escherichia coli. Mol Microbiol 68:1128–1148. doi: 10.1111/j.1365-2958.2008.06229.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Mittenhuber G. 2001. Comparative genomics and evolution of genes encoding bacterial (p)ppGpp synthetases/hydrolases (the Rel, RelA and SpoT proteins). J Mol Microbiol Biotechnol 3:585–600. [PubMed] [Google Scholar]
- 85.Ouellette SP, Rueden KJ, Rucks EA. 2016. Tryptophan codon-dependent transcription in Chlamydia pneumoniae during gamma interferon-mediated tryptophan limitation. Infect Immun 84:2703–2713. doi: 10.1128/IAI.00377-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Miethke M, Westers H, Blom EJ, Kuipers OP, Marahiel MA. 2006. Iron starvation triggers the stringent response and induces amino acid biosynthesis for bacillibactin production in Bacillus subtilis. J Bacteriol 188:8655–8657. doi: 10.1128/JB.01049-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Akers JC, Tan M. 2006. Molecular mechanism of tryptophan-dependent transcriptional regulation in Chlamydia trachomatis. J Bacteriol 188:4236–4243. doi: 10.1128/JB.01660-05. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Carlson JH, Wood H, Roshick C, Caldwell HD, McClarty G. 2006. In vivo and in vitro studies of Chlamydia trachomatis TrpR:DNA interactions. Mol Microbiol 59:1678–1691. doi: 10.1111/j.1365-2958.2006.05045.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Amirshahi A, Wan C, Beagley K, Latter J, Symonds I, Timms P. 2011. Modulation of the Chlamydia trachomatis in vitro transcriptome response by the sex hormones estradiol and progesterone. BMC Microbiol 11:150. doi: 10.1186/1471-2180-11-150. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Caspi R, Billington R, Ferrer L, Foerster H, Fulcher CA, Keseler IM, Kothari A, Krummenacker M, Latendresse M, Mueller LA, Ong Q, Paley S, Subhraveti P, Weaver DS, Karp PD. 2016. The MetaCyc database of metabolic pathways and enzymes and the BioCyc collection of pathway/genome databases. Nucleic Acids Res 44:D471–D480. doi: 10.1093/nar/gkv1164. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
Raw and processed sequencing files were submitted to the NCBI Gene Expression Omnibus (GEO) as a Superseries, and the midcycle and early-cycle projects can be found using accession number GSE106763.