Skip to main content
Advertisement
  • Loading metrics

Bringing genetics to heretofore intractable obligate intracellular bacterial pathogens: Chlamydia and beyond

  • Magnus Ölander,

    Affiliations Department of Molecular Biology, Umeå University, Umeå, Sweden, The Laboratory for Molecular Infection Medicine Sweden (MIMS), Umeå University, Umeå, Sweden, Umeå Centre for Microbial Research (UCMR), Umeå University, Umeå, Sweden

  • Barbara S. Sixt

    barbara.sixt@umu.se

    Affiliations Department of Molecular Biology, Umeå University, Umeå, Sweden, The Laboratory for Molecular Infection Medicine Sweden (MIMS), Umeå University, Umeå, Sweden, Umeå Centre for Microbial Research (UCMR), Umeå University, Umeå, Sweden

What are obligate intracellular bacteria and why should we care?

As their designation implies, obligate intracellular bacteria are microbes that have developed lifestyles so closely entwined with the cells of the hosts they infect that they can reproduce only within the confines of these cells. Among this group of bacteria are several pathogenic species with a significant impact on human health. For instance, Chlamydia spp. are responsible for millions of cases of urogenital, ocular, and respiratory infections every year [1]. Moreover, Coxiella burnetii is the agent of Q fever [2], whereas members of the order Rickettsiales (specifically the genera Rickettsia, Ehrlichia, and Anaplasma) cause life-threatening vector-borne diseases, such as spotted fever and typhus [3].

While these pathogens are all restricted to life in an intracellular niche, they differ in the ways they reproduce and engage with their host cells. Chlamydia spp., Ehrlichia spp., Anaplasma spp., and C. burnetii have biphasic life cycles, in which the bacteria alternate between infectious (environmentally stable) and replicative (metabolically highly active) forms inside specialized membrane-enclosed vacuoles [13]. In contrast, Rickettsia spp. escape their vacuoles after uptake and replicate in the host cytosol without undergoing developmental transitions [3]. The various obligate intracellular lifestyles are enabled by distinct sets of virulence factors, which, for instance, engage in the subversion of host defenses and the hijacking of host resources and machineries.

Studying the molecular basis of these intimate relationships between host and bacteria can uncover unknown aspects of host–pathogen interactions and new targets for pharmacological intervention to help relieve the significant disease burden of intracellular infections. However, historically, the study of obligate intracellular bacteria was neglected due to the impeding lack of tools enabling their molecular genetic manipulation. While whole-genome sequencing provided the opportunity to predict their virulence traits, in the absence of genetic tools, direct links to specific bacterial genes could not be formally demonstrated.

Fortunately, recent years have seen a significant expansion of our genetic toolbox for these important pathogens. Here, we summarize these developments, with a focus on Chlamydia spp., along with a brief overview for other obligate intracellular bacteria. Furthermore, we provide examples of insights into chlamydial biology enabled by the expanded toolbox, and we conclude with a discussion on future perspectives for the molecular genetic manipulation of this group of bacteria.

Why is the genetic manipulation of obligate intracellular bacteria so challenging?

A major factor restricting the genetic tractability of obligate intracellular bacteria is their main defining characteristic, i.e., their intracellular lifestyle. Delivery of DNA into bacteria residing within host cells is difficult. Therefore, most approaches that depend on DNA delivery require the bacteria first to be isolated before they can be modified extracellularly and then reintroduced into suitable host cells [47]. However, the distinct developmental forms of some species can be differentially receptive for DNA uptake, with the infective forms being generally considered less amenable to modification due to their rigid cell walls, condensed nucleoids, and/or reduced metabolic activities [8,9].

As a consequence of the resulting low transformation efficiencies, selection and recovery of modified bacteria typically requires the expansion of a very low number of transformed bacteria through passage in cell culture, a process that can take days to weeks. Notably, antibiotic resistance genes are commonly used as selection markers, but the list of antibiotics that can be used is limited by regulatory prohibitions. Moreover, antibiotics that cannot access the intracellular site of bacterial replication in an active form, or induce spontaneous resistance at a high frequency, cannot be used. On an encouraging note, several nonantibiotic selection markers, for instance, based on herbicide resistance or complementation of amino acid auxotrophies, have already been introduced in some obligate intracellular species [1012].

The strict intracellular lifestyle also complicates the recovery of clonal strains of modified bacteria because in the absence of a possibility for host cell-free cultivation, clones cannot be obtained simply by picking colonies from agar plates. Plaque purification is applicable to certain obligate intracellular bacteria, such as some Chlamydia and Rickettsia spp. [13,14]. However, when plaque purification is not possible, clonal strains instead need to be recovered by the slow process of limiting dilution [15] or by technically more demanding techniques, such as micromanipulation [16], laser microdissection [17], or cell sorting [18].

Overall, while becoming technically increasingly feasible, the genetic manipulation of obligate intracellular bacteria remains laborious and time-consuming. For example, clones of transformed Escherichia coli may be obtained overnight, while the generation of plaque-purified transformed Chlamydia spp. takes at least a month (Fig 1). Moreover, the genetic manipulation of obligate intracellular bacteria is not only complicated by technical difficulties, but also by the fact that many genes with roles in host–pathogen interactions are essential for the ability of these pathogens to replicate and can therefore not be disrupted easily.

thumbnail
Fig 1. A timeline for the generation of transformed bacteria in C. trachomatis and E. coli.

https://doi.org/10.1371/journal.ppat.1010669.g001

Clearly, these hurdles could be overcome if host cell-free cultivation was possible. Indeed, this is well illustrated by the case of C. burnetii, whose genetic toolbox has been greatly expanded following the development of an axenic medium [19]. Unfortunately, efforts to repeat this feat for other obligate intracellular bacteria have so far only resulted in axenic media that enable survival with detectable metabolic activity [2022], while media that allow these bacteria to replicate under host cell-free conditions still await development.

What tools for genetic manipulation of obligate intracellular bacteria have been developed so far?

Transformation protocols

Genetic manipulation of bacteria often starts with a transformation step in which the bacteria are stimulated to take up and incorporate exogenous DNA. A CaCl2-based chemical transformation method has been widely adopted for transformation of host cell-free (purified) Chlamydia spp. [4,2325], while electroporation is the method of choice for other obligate intracellular bacteria [57]. In attempts to bypass the need for purified bacteria, polyamidoamine dendrimers were successfully tested in Chlamydia spp. [26,27] and Anaplasma phagocytophilum [28] to introduce plasmids into bacteria residing in host cells. Yet, dendrimer-based protocols have not yet been further developed and thus remain at a proof-of-principle stage. Of note, in Chlamydia trachomatis, the fact that DNA, including plasmids, can be naturally transferred between coinfecting strains through lateral gene transfer has also been exploited to introduce DNA [29].

Shuttle vectors for gene expression

Shuttle vectors are plasmids that can be engineered in a convenient system such as E. coli and then transferred through transformation to the host of interest, where the plasmids are stably maintained to enable expression of genes. Shuttle vectors have been developed for several species of obligate intracellular bacteria with C. trachomatis currently having one of the most versatile arsenals of vectors. These vectors are commonly based on the endogenous plasmid of C. trachomatis serovar L2 and contain various selection markers and promoter systems, including such that enable inducible gene expression [4,30,31]. More recently, shuttle vectors for other Chlamydia spp. have been developed as well [2325], and attempts are underway to develop a broad-spectrum vector system [32]. For Rickettsia spp., vectors were constructed based on endogenous plasmids found in the spotted fever group species Rickettsia amblyommatis and were shown to be stably maintained in members from both the spotted fever and typhus groups of the genus [33,34]. Finally, vectors have also been created for C. burnetii based on plasmid backbones previously used in Legionella pneumophila [35,36] and, more recently, based on an endogenous plasmid from C. burnetii itself [37].

Targeted modifications in bacterial chromosomes

Site-specific mutations in genomes of obligate intracellular bacteria have principally been generated by using site-specific transposons, the proprietary TargeTron system, or allelic exchange approaches. A transposon system enabling single copy integration of genes at a specific chromosomal site has for instance been described for C. burnetii [38], while similar approaches have so far not been established in Chlamydia spp. In contrast, the TargeTron system, which takes advantage of the ability of mobile group II introns to insert themselves into the bacterial genome and which in contrast to the transposon system can be engineered to target a specific site of choice, has been widely applied in C. trachomatis, in particular to disrupt genes coding for secreted effector proteins [3941]. Beyond Chlamydia spp., however, the use of TargeTron in obligate intracellular bacteria has so far been limited to Ehrlichia chaffeensis [7] and Rickettsia rickettsii [42]. Targeted modification by allelic exchange, a process that replaces a selected stretch of DNA via homologous recombination and enables the most versatile modifications, was enabled in C. trachomatis by the development of fluorescence-reported allelic exchange mutagenesis (FRAEM) [43]. Its latest implementation, exploiting the Cre-lox system for site-specific recombination, even allows for the markerless deletion of operon-localized genes avoiding polar effects [29]. Allelic exchange has also been successfully used to disrupt genes in C. burnetii [44], E. chaffeensis [7], and Rickettsia prowazekii [45]. Finally, worth mentioning is that the feasibility of generating conditional knockdowns in C. trachomatis using CRISPR interference has recently been demonstrated [46,47], which now facilitates the study of essential genes that cannot be disrupted by the approaches mentioned above [48,49].

Random mutagenesis methods

Bacterial genomes can be randomly mutagenized by various means, but the main methods that have been used for obligate intracellular bacteria so far are transposon mutagenesis and exposure to chemical mutagens. Chemical methods have so far been the mainstay of random mutagenesis in Chlamydia spp. [5053], for which successful Himar1-based transposon mutagenesis was only recently reported [54,55]. The latter has, on the other hand, been more commonly used in other obligate intracellular bacteria, including C. burnetii [56], E. chaffeensis [7], A. phagocytophilum [6], and R. prowazekii [57]. Both mutagenesis methods allow the generation of mutant libraries, which can be used in forward and reverse genetic approaches, i.e., for the identification of mutants that have desired traits or defects in specific genes. An advantage of the chemical mutagenesis approach is that it can also generate hypomorphic alleles, which can facilitate the study of essential genes. In addition, it often introduces multiple mutations per strain; therefore, fewer mutants may need to be screened. However, associating phenotypes with specific chemically induced point mutations is more tedious compared to the identification of single transposon integration sites. In Chlamydia, this typically involves whole-genome sequencing in combination with genetic mapping approaches based on the generation of recombinant strains [51,58]. Moreover, strains with point mutations in specific genes can be found either through whole-genome sequencing of entire collections of mutated strains or through a targeted search approach named TILLING (targeting induced local lesions in genomes) [50,52].

What has the application of genetic tools taught us so far about Chlamydia spp.?

The ability to disrupt genes and to then complement the observed phenotypic defects by restoring their expression, allows us to unequivocally assign functions to specific genes. Moreover, the expression of tagged proteins facilitates studies of their localization and interactors, while the expression of modified protein variants can help identifying functional domains. Hence, a versatile genetic toolbox is a powerful asset for in-depth gene characterizations (Fig 2).

thumbnail
Fig 2. The benefits of a versatile genetic toolbox in the exploration of gene functions.

https://doi.org/10.1371/journal.ppat.1010669.g002

In C. trachomatis, the application of these novel approaches has in particular advanced our knowledge of its repertoire of secreted virulence factors and thus our understanding of key aspects of Chlamydia’s interaction with host cells. To give examples, two C. trachomatis effectors, i.e., TarP and TmeA, were shown to promote host cell invasion via remodeling of the actin cytoskeleton at the invasion site [29,5961]. Moreover, functions could be assigned to a significant number of the so-called inclusion membrane proteins (Incs), a class of secreted effectors that decorate the membrane of the Chlamydia vacuole (called inclusion). For instance, IncA was shown to mediate homotypic inclusion fusion [39], IncD and IncV to establish contact sites between the inclusion and the endoplasmic reticulum [62,63], and IncE and CpoS to modulate host membrane trafficking [41,6466]. CpoS deficiency had a particularly detrimental effect on bacterial replication, as it resulted in a premature death of the infected cell, presumably as a result of premature inclusion lysis [40,41]. Interestingly, inclusion instability was also noted in cells infected with various other Inc mutants [40], including such deficient for InaC, an Inc that promotes microtubule stabilization and actin cage formation at the periphery of the inclusion [52,67,68]. Finally, concerning Chlamydia exit from host cells, we learned that the Incs MrcA and CT228 regulate myosin II motor protein activity to promote or suppress exit by extrusion [69,70], while the secreted protease CPAF and the plasmid-encoded factor PGP4 contribute to bacterial egress by host cell lysis [71].

The possibility for genetic manipulation has also empowered the use of reporter genes. For instance, the expression of fluorescent proteins in Chlamydia spp., now routinely applied, can ease the analysis of infection-related processes, such as adhesion, entry, and bacterial growth [72]. Moreover, their use for monitoring promoter activities and developmental transitions now also allows us to decode the mysteries of Chlamydia’s developmental biology, which, for example, recently revealed that the transition of the replicative form of C. trachomatis into the infectious form is regulated by bacteria-intrinsic not external cues [73]. In addition, it was shown that the use of luciferase-expressing Chlamydia muridarum strains can enable in vivo imaging of Chlamydia infections in mice [74]. The application of this novel tool uncovered an unanticipated spread of the bacteria from the primary infection site, the genital tract, to other organs, followed by a long-lasting colonization of the mouse intestine [75]. Of note, the plasmid-encoded effector PGP3 and several chromosomal-encoded bacterial genes were later identified as virulence factors necessary for C. muridarum to survive the acidic conditions in the stomach and to colonize the colon [7678]. Because these virulence factors were also required for hydrosalpinx induction, intestinal colonization was proposed to be a key event in promoting upper genital tract pathology in infected mice [76,78].

What further developments can we expect in the future for the genetic manipulation of Chlamydia spp.?

In spite of all the progress made in the genetic manipulation of obligate intracellular bacteria, major technical hurdles remain that significantly restrict the biological insights we can gain. For instance, considering the case of C. trachomatis, it should now principally be feasible to integrate or replace genes or regulatory elements at any desired site in the chlamydial chromosome or to markerlessly remove genes or larger genomic regions. However, in practice, this remains technically difficult, as it requires recovery of rare double cross-over allelic exchange events. This could be significantly eased if counterselectable markers, already described for C. burnetii [44], would be available for Chlamydia spp. as well. The use of temperature-sensitive alleles has recently been proposed as one possible solution [58].

Another expected major milestone in Chlamydia will be the development of saturation mutagenesis enabling transposon-insertion sequencing. It appears that many secreted Chlamydia effectors are nonessential in cell culture, suggesting either an inbuilt redundancy of effector functions or more likely context-specific roles. For instance, certain functions might be relevant only in specific host species, tissues, or cell types, or only in the context of an in vivo infection, for instance, to promote immune evasion or dissemination. Transposon-insertion sequencing holds great promise in identifying essential genes and genes providing such context-specific fitness benefits. While the low transformation efficiencies in Chlamydia have been considered a major obstacle for establishing such approach, the recent development of inducible transposon mutagenesis in C. trachomatis suggests that it can be bypassed [79].

Finally, a better understanding of Chlamydia’s unique developmental biology will also require a further refinement of our capabilities to study the function of essential genes, such as by developing more tightly controlled inducible expression systems to improve conditional gene expression and disruption approaches. Again, most recent developments in the field, such as the description of riboswitches for translational control of gene expression in Chlamydia [79,80], provide a highly encouraging perspective.

Taken together, while there are still many obstacles in our way, we can be confident that the establishment and refinement of novel genetic tools for obligate intracellular bacteria, such as Chlamydia spp. and beyond, will continue at fast pace. Clearly, these advances will further revolutionize both the ease and depth with which we can decipher the secret ways by which these pathogens modulate host cell biology and cause diseases.

Acknowledgments

We would like to thank the research community for their heroic efforts in advancing the field and like to apologize to all those whose contributions could not be highlighted here due to space constraints.

References

  1. 1. Elwell C, Mirrashidi K, Engel J. Chlamydia cell biology and pathogenesis. Nat Rev Microbiol. 2016;14(6):385–400. pmid:27108705
  2. 2. van Schaik EJ, Chen C, Mertens K, Weber MM, Samuel JE. Molecular pathogenesis of the obligate intracellular bacterium Coxiella burnetii. Nat Rev Microbiol. 2013;11(8):561–573. pmid:23797173
  3. 3. Salje J. Cells within cells: Rickettsiales and the obligate intracellular bacterial lifestyle. Nat Rev Microbiol. 2021;19(6):375–390. pmid:33564174
  4. 4. Wang Y, Kahane S, Cutcliffe LT, Skilton RJ, Lambden PR, Clarke IN. Development of a transformation system for Chlamydia trachomatis: restoration of glycogen biosynthesis by acquisition of a plasmid shuttle vector. PLoS Pathog. 2011;7(9):e1002258. pmid:21966270
  5. 5. Rachek LI, Tucker AM, Winkler HH, Wood DO. Transformation of Rickettsia prowazekii to rifampin resistance. J Bacteriol. 1998;180(8):2118–2124. pmid:9555894
  6. 6. Felsheim RF, Herron MJ, Nelson CM, Burkhardt NY, Barbet AF, Kurtti TJ, et al. Transformation of Anaplasma phagocytophilum. BMC Biotechnol. 2006;6:42. pmid:17076894
  7. 7. Cheng CM, Nair ADS, Indukuri VV, Gong SZ, Felsheim RF, Jaworski D, et al. Targeted and random mutagenesis of Ehrlichia chaffeensis for the identification of genes required for in vivo infection. PLoS Pathog. 2013;9(2):e1003171. pmid:23459099
  8. 8. Beare PA, Sandoz KM, Omsland A, Rockey DD, Heinzen RA. Advances in genetic manipulation of obligate intracellular bacterial pathogens. Front Microbiol. 2011;2:97. pmid:21833334
  9. 9. Sixt BS, Valdivia RH. Molecular genetic analysis of Chlamydia species. Annu Rev Microbiol. 2016;70:179–198. pmid:27607551
  10. 10. Chavez ASO, Herron MJ, Nelson CM, Felsheim RF, Oliver JD, Burkhardt NY, et al. Mutational analysis of gene function in the Anaplasmataceae: challenges and perspectives. Ticks Tick Borne Dis. 2019;10(2):482–494. pmid:30466964
  11. 11. Sandoz KM, Beare PA, Cockrell DC, Heinzen RA. Complementation of arginine auxotrophy for genetic transformation of Coxiella burnetii by use of a defined axenic medium. Appl Environ Microbiol. 2016;82(10):3042–3051. pmid:26969695
  12. 12. Beare PA, Jeffrey BM, Long CM, Martens CM, Heinzen RA. Genetic mechanisms of Coxiella burnetii lipopolysaccharide phase variation. PLoS Pathog. 2018;14(3):e1006922. pmid:29481553
  13. 13. Wike DA, Tallent G, Peacock MG, Ormsbee RA. Studies of rickettsial plaque assay technique. Infect Immun. 1972;5(5):715–722. pmid:4629250
  14. 14. Matsumoto A, Izutsu H, Miyashita N, Ohuchi M. Plaque formation by and plaque cloning of Chlamydia trachomatis biovar trachoma. J Clin Microbiol. 1998;36(10):3013–3019. pmid:9738059
  15. 15. Suhan ML, Chen SY, Thompson HA. Transformation of Coxiella burnetti to ampicillin resistance. J Bacteriol. 1996;178(9):2701–2708. pmid:8626341
  16. 16. Beare PA, Howe D, Cockrell DC, Heinzen RA. Efficient method of cloning the obligate intracellular bacterium Coxiella burnetii. Appl Environ Microbiol. 2007;73(12):4048–4054. pmid:17468273
  17. 17. Podgorny OV, Polina NF, Babenko VV, Karpova IY, Kostryukova ES, Govorun VM, et al. Isolation of single Chlamydia-infected cells using laser microdissection. J Microbiol Methods. 2015;109:123–128. pmid:25546842
  18. 18. Alzhanov DT, Suchland RJ, Bakke AC, Stamm WE, Rockey DD. Clonal isolation of Chlamydia-infected cells using flow cytometry. J Microbiol Methods. 2007;68(1):201–208. pmid:16997404
  19. 19. Omsland A, Cockrell DC, Howe D, Fischer ER, Virtaneva K, Sturdevant DE, et al. Host cell-free growth of the Q fever bacterium Coxiella burnetii. Proc Natl Acad Sci U S A. 2009;106(11):4430–4434. pmid:19246385
  20. 20. Omsland A, Sager J, Nair V, Sturdevant DE, Hackstadt T. Developmental stage-specific metabolic and transcriptional activity of Chlamydia trachomatis in an axenic medium. Proc Natl Acad Sci U S A. 2012;109(48):19781–19785. pmid:23129646
  21. 21. Eedunuri VK, Zhang YT, Cheng CM, Chen L, Liu HT, Omsland A, et al. Protein and DNA synthesis demonstrated in cell-free Ehrlichia chaffeensis organisms in axenic medium. Sci Rep. 2018;8(1):9293. pmid:29915240
  22. 22. Zhang YT, Chen L, Kondethimmanahalli C, Liu HT, Ganta RR. Protein and DNA biosynthesis demonstrated in host cell-free phagosomes containing Anaplasma phagocytophilum or Ehrlichia chaffeensis in axenic media. Infect Immun. 2021;89(4):e00638–e00620. pmid:33431703
  23. 23. Song L, Carlson JH, Zhou B, Virtaneva K, Whitmire WM, Sturdevant GL, et al. Plasmid-mediated transformation tropism of chlamydial biovars. Pathog Dis. 2014;70(2):189–193. pmid:24214488
  24. 24. Shima K, Wanker M, Skilton RJ, Cutcliffe LT, Schnee C, Kohl TA, et al. The genetic transformation of Chlamydia pneumoniae. mSphere. 2018;3(5):e00412–e00418. pmid:30305318
  25. 25. Shima K, Weber MM, Schnee C, Sachse K, Kading N, Klinger M, et al. Development of a plasmid shuttle vector system for genetic manipulation of Chlamydia psittaci. mSphere. 2020;5(4):e00787–e00720. pmid:32848009
  26. 26. Gérard HC, Mishra MK, Mao G, Wang S, Hali M, Whittum-Hudson JA, et al. Dendrimer-enabled DNA delivery and transformation of Chlamydia pneumoniae. Nanomedicine. 2013;9(7):996–1008. pmid:23639679
  27. 27. Kannan RM, Gerard HC, Mishra MK, Mao G, Wang S, Hali M, et al. Dendrimer-enabled transformation of Chlamydia trachomatis. Microb Pathog. 2013;65:29–35. pmid:24075820
  28. 28. Oki AT, Seidman D, Lancina MG 3rd, Mishra MK, Kannan RM, Yang H, et al. Dendrimer-enabled transformation of Anaplasma phagocytophilum. Microbes Infect. 2015;17(11–12):817–822. pmid:26369714
  29. 29. Keb G, Hayman R, Fields KA. Floxed-cassette allelic exchange mutagenesis enables markerless gene deletion in Chlamydia trachomatis and can reverse cassette-induced polar effects. J Bacteriol. 2018;200(24):e00479–e00418. pmid:30224436
  30. 30. Wickstrum J, Sammons LR, Restivo KN, Hefty PS. Conditional gene expression in Chlamydia trachomatis using the tet system. PLoS ONE. 2013;8(10):e76743. pmid:24116144
  31. 31. Bastidas RJ, Valdivia RH. Emancipating Chlamydia: Advances in genetic engineering of a recalcitrant intracellular pathogen. Microbiol Mol Biol Rev. 2016;80(2):411–427. pmid:27030552
  32. 32. Garvin L, Vande Voorde R, Dickinson M, Carrell S, Hybiske K, Rockey D. A broad-spectrum cloning vector that exists as both an integrated element and a free plasmid in Chlamydia trachomatis. PLoS ONE. 2021;16(12):e0261088. pmid:34914750
  33. 33. Burkhardt NY, Baldridge GD, Williamson PC, Billingsley PM, Heu CC, Felsheim RF, et al. Development of shuttle vectors for transformation of diverse Rickettsia species. PLoS ONE. 2011;6(12):e29511. pmid:22216299
  34. 34. Wood DO, Hines A, Tucker AM, Woodard A, Driskell LO, Burkhardt NY, et al. Establishment of a replicating plasmid in Rickettsia prowazekii. PLoS ONE. 2012;7(4):e34715. pmid:22529927
  35. 35. Chen C, Banga S, Mertens K, Weber MM, Gorbaslieva I, Tan Y, et al. Large-scale identification and translocation of type IV secretion substrates by Coxiella burnetii. Proc Natl Acad Sci U S A. 2010;107(50):21755–21760. pmid:21098666
  36. 36. Voth DE, Beare PA, Howe D, Sharma UM, Samoilis G, Cockrell DC, et al. The Coxiella burnetii cryptic plasmid is enriched in genes encoding type IV secretion system substrates. J Bacteriol. 2011;193(7):1493–1503. pmid:21216993
  37. 37. Luo S, Lu S, Fan H, Sun Z, Hu Y, Li R, et al. The Coxiella burnetii QpH1 plasmid is a virulence factor for colonizing bone marrow-derived murine macrophages. J Bacteriol. 2021;203(9):e00588–e00520. pmid:33558394
  38. 38. Beare PA, Gilk SD, Larson CL, Hill J, Stead CM, Omsland A, et al. Dot/Icm type IVB secretion system requirements for Coxiella burnetii growth in human macrophages. MBio. 2011;2(4):e00175–e00111. pmid:21862628
  39. 39. Johnson CM, Fisher DJ. Site-specific, insertional inactivation of incA in Chlamydia trachomatis using a group II intron. PLoS ONE. 2013;8(12):e83989. pmid:24391860
  40. 40. Weber MM, Lam JL, Dooley CA, Noriea NF, Hansen BT, Hoyt FH, et al. Absence of specific Chlamydia trachomatis inclusion membrane proteins triggers premature inclusion membrane lysis and host cell death. Cell Rep. 2017;19(7):1406–1417. pmid:28514660
  41. 41. Sixt BS, Bastidas RJ, Finethy R, Baxter RM, Carpenter VK, Kroemer G, et al. The Chlamydia trachomatis inclusion membrane protein CpoS counteracts STING-mediated cellular surveillance and suicide programs. Cell Host Microbe. 2017;21(1):113–121. pmid:28041929
  42. 42. Noriea NF, Clark TR, Hackstadt T. Targeted knockout of the Rickettsia rickettsii OmpA surface antigen does not diminish virulence in a mammalian model system. MBio. 2015;6(2):e00323–e00315. pmid:25827414
  43. 43. Mueller KE, Wolf K, Fields KA. Gene deletion by fluorescence-reported allelic exchange mutagenesis in Chlamydia trachomatis. MBio. 2016;7(1):e01817–e01815. pmid:26787828
  44. 44. Beare PA, Larson CL, Gilk SD, Heinzen RA. Two systems for targeted gene deletion in Coxiella burnetii. Appl Environ Microbiol. 2012;78(13):4580–4589. pmid:22522687
  45. 45. Driskell LO, Yu XJ, Zhang L, Liu Y, Popov VL, Walker DH, et al. Directed mutagenesis of the Rickettsia prowazekii pld gene encoding phospholipase D. Infect Immun. 2009;77(8):3244–3248. pmid:19506016
  46. 46. Ouellette SP. Feasibility of a conditional knockout system for Chlamydia based on CRISPR interference. Front Cell Infect Microbiol. 2018;8:59. pmid:29535977
  47. 47. Ouellette SP, Blay EA, Hatch ND, Fisher-Marvin LA. CRISPR interference to inducibly repress gene expression in Chlamydia trachomatis. Infect Immun. 2021;89(7):e0010821. pmid:33875479
  48. 48. Brockett MR, Lee J, Cox JV, Liechti GW, Ouellette SP. A dynamic, ring-forming bactofilin critical for maintaining cell size in the obligate intracellular bacterium Chlamydia trachomatis. Infect Immun. 2021;89(8):e0020321. pmid:33941579
  49. 49. Wood NA, Blocker AM, Seleem MA, Conda-Sheridan M, Fisher DJ, Ouellette SP. The ClpX and ClpP2 orthologs of Chlamydia trachomatis perform discrete and essential functions in organism growth and development. MBio. 2020;11(5):e02016–e02020. pmid:32873765
  50. 50. Kari L, Goheen MM, Randall LB, Taylor LD, Carlson JH, Whitmire WM, et al. Generation of targeted Chlamydia trachomatis null mutants. Proc Natl Acad Sci U S A. 2011;108(17):7189–7193. pmid:21482792
  51. 51. Nguyen BD, Valdivia RH. Virulence determinants in the obligate intracellular pathogen Chlamydia trachomatis revealed by forward genetic approaches. Proc Natl Acad Sci U S A. 2012;109(4):1263–1268. pmid:22232666
  52. 52. Kokes M, Dunn JD, Granek JA, Nguyen BD, Barker JR, Valdivia RH, et al. Integrating chemical mutagenesis and whole-genome sequencing as a platform for forward and reverse genetic analysis of Chlamydia. Cell Host Microbe. 2015;17(5):716–725. pmid:25920978
  53. 53. Rajaram K, Giebel AM, Toh E, Hu S, Newman JH, Morrison SG, et al. Mutational analysis of the Chlamydia muridarum plasticity zone. Infect Immun. 2015;83(7):2870–2881. pmid:25939505
  54. 54. LaBrie SD, Dimond ZE, Harrison KS, Baid S, Wickstrum J, Suchland RJ, et al. Transposon mutagenesis in Chlamydia trachomatis identifies CT339 as a ComEC homolog important for DNA uptake and lateral gene transfer. MBio. 2019;10(4):e01343–e01319. pmid:31387908
  55. 55. Wang Y, LaBrie SD, Carrell SJ, Suchland RJ, Dimond ZE, Kwong F, et al. Development of transposon mutagenesis for Chlamydia muridarum. J Bacteriol. 2019;201(23):e00366–e00319. pmid:31501283
  56. 56. Beare PA, Howe D, Cockrell DC, Omsland A, Hansen B, Heinzen RA. Characterization of a Coxiella burnetii ftsZ mutant generated by Himar1 transposon mutagenesis. J Bacteriol. 2009;191(5):1369–1381. pmid:19114492
  57. 57. Liu ZM, Tucker AM, Driskell LO, Wood DO. Mariner-based transposon mutagenesis of Rickettsia prowazekii. Appl Environ Microbiol. 2007;73(20):6644–6649. pmid:17720821
  58. 58. Brothwell JA, Muramatsu MK, Toh E, Rockey DD, Putman TE, Barta ML, et al. Interrogating genes that mediate Chlamydia trachomatis survival in cell culture using conditional mutants and recombination. J Bacteriol. 2016;198(15):2131–2139. pmid:27246568
  59. 59. Ghosh S, Ruelke EA, Ferrell JC, Bodero MD, Fields KA, Jewett TJ. Fluorescence-reported allelic exchange mutagenesis-mediated gene deletion indicates a requirement for Chlamydia trachomatis Tarp during in vivo infectivity and reveals a specific role for the C terminus during cellular invasion. Infect Immun. 2020;88(5):e00841–e00819. pmid:32152196
  60. 60. McKuen MJ, Mueller KE, Bae YS, Fields KA. Fluorescence-reported allelic exchange mutagenesis reveals a role for Chlamydia trachomatis TmeA in invasion that is independent of host AHNAK. Infect Immun. 2017;85(12):e00640–e00617. pmid:28970272
  61. 61. Faris R, McCullough A, Andersen SE, Moninger TO, Weber MM. The Chlamydia trachomatis secreted effector TmeA hijacks the N-WASP-ARP2/3 actin remodeling axis to facilitate cellular invasion. PLoS Pathog. 2020;16(9):e1008878. pmid:32946535
  62. 62. Stanhope R, Flora E, Bayne C, Derre I. IncV, a FFAT motif-containing Chlamydia protein, tethers the endoplasmic reticulum to the pathogen-containing vacuole. Proc Natl Acad Sci U S A. 2017;114(45):12039–12044. pmid:29078338
  63. 63. Agaisse H, Derre I. Expression of the effector protein IncD in Chlamydia trachomatis mediates recruitment of the lipid transfer protein CERT and the endoplasmic reticulum-resident protein VAPB to the inclusion membrane. Infect Immun. 2014;82(5):2037–2047. pmid:24595143
  64. 64. Mirrashidi KM, Elwell CA, Verschueren E, Johnson JR, Frando A, Von Dollen J, et al. Global mapping of the Inc-human interactome reveals that retromer restricts Chlamydia infection. Cell Host Microbe. 2015;18(1):109–121. pmid:26118995
  65. 65. Meier K, Jachmann LH, Pérez L, Kepp O, Valdivia RH, Kroemer G, et al. The Chlamydia protein CpoS modulates the inclusion microenvironment and restricts the interferon response by acting on Rab35. bioRxiv. 2022.
  66. 66. Faris R, Merling M, Andersen SE, Dooley CA, Hackstadt T, Weber MM. Chlamydia trachomatis CT229 subverts Rab GTPase-dependent CCV trafficking pathways to promote chlamydial infection. Cell Rep. 2019;26(12):3380–3390. pmid:30893609
  67. 67. Wesolowski J, Weber MM, Nawrotek A, Dooley CA, Calderon M, St Croix CM, et al. Chlamydia hijacks ARF GTPases to coordinate microtubule posttranslational modifications and Golgi complex positioning. MBio. 2017;8(3):e02280–e02216. pmid:28465429
  68. 68. Haines A, Wesolowski J, Ryan NM, Monteiro-Brás T, Paumet FA-O. Cross talk between ARF1 and RhoA coordinates the formation of cytoskeletal scaffolds during Chlamydia infection. MBio. 2021;12(6):e0239721. pmid:34903051
  69. 69. Shaw JH, Key CE, Snider TA, Sah P, Shaw EI, Fisher DJ, et al. Genetic inactivation of Chlamydia trachomatis inclusion membrane protein CT228 alters MYPT1 recruitment, extrusion production, and longevity of infection. Front Cell Infect Microbiol. 2018;8:415. pmid:30555802
  70. 70. Nguyen PH, Lutter EI, Hackstadt T. Chlamydia trachomatis inclusion membrane protein MrcA interacts with the inositol 1,4,5-trisphosphate receptor type 3 (ITPR3) to regulate extrusion formation. PLoS Pathog. 2018;14(3):e1006911. pmid:29543918
  71. 71. Yang C, Starr T, Song L, Carlson JH, Sturdevant GL, Beare PA, et al. Chlamydial lytic exit from host cells is plasmid regulated. MBio. 2015;6(6):e01648–e01615. pmid:26556273
  72. 72. Vromman F, Laverriere M, Perrinet S, Dufour A, Subtil A. Quantitative monitoring of the Chlamydia trachomatis developmental cycle using GFP-expressing bacteria, microscopy and flow cytometry. PLoS ONE. 2014;9(6):e99197. pmid:24911516
  73. 73. Chiarelli TJ, Grieshaber NA, Omsland A, Remien CH, Grieshaber SS. Single-inclusion kinetics of Chlamydia trachomatis development. mSystems. 2020;5(5):e00689–e00620. pmid:33051378
  74. 74. Campbell J, Huang Y, Liu Y, Schenken R, Arulanandam B, Zhong G. Bioluminescence imaging of Chlamydia muridarum ascending infection in mice. PLoS ONE. 2014;9(7):e101634. pmid:24983626
  75. 75. Zhang Q, Huang Y, Gong S, Yang Z, Sun X, Schenken R, et al. In vivo and ex vivo imaging reveals a long-lasting chlamydial infection in the mouse gastrointestinal tract following genital tract inoculation. Infect Immun. 2015;83(9):3568–3577. pmid:26099591
  76. 76. Shao L, Zhang T, Melero J, Huang Y, Liu Y, Liu Q, et al. The genital tract virulence factor pGP3 is essential for Chlamydia muridarum colonization in the gastrointestinal tract. Infect Immun. 2018;86(1):e00429–e00417. pmid:29038127
  77. 77. Zhang T, Huo Z, Ma J, He C, Zhong G. The plasmid-encoded pGP3 promotes Chlamydia evasion of acidic barriers in both stomach and vagina. Infect Immun. 2019;87(5):e00844–e00818. pmid:30858342
  78. 78. Koprivsek JJ, Zhang T, Tian Q, He Y, Xu H, Xu Z, et al. Distinct roles of chromosome- versus plasmid-encoded genital tract virulence factors in promoting Chlamydia muridarum colonization in the gastrointestinal tract. Infect Immun. 2019;87(8):e00265–e00219. pmid:31160366
  79. 79. O’Neill CE, Skilton RJ, Forster J, Cleary DW, Pearson SA, Lampe DJ, et al. An inducible transposon mutagenesis approach for the intracellular human pathogen Chlamydia trachomatis. Wellcome Open Res. 2021;6:312. pmid:35087955
  80. 80. Grieshaber NA, Chiarelli TJ, Appa CR, Neiswanger G, Peretti K, Grieshaber SS. Translational gene expression control in Chlamydia trachomatis. PLoS ONE. 2022;17(1):e0257259. pmid:35085261