Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Oct 27.
Published in final edited form as: Curr Opin Microbiol. 2009 Dec 16;13(1):4–10. doi: 10.1016/j.mib.2009.11.002

Acquisition of nutrients by Chlamydiae: unique challenges of living in an intracellular compartment

Hector Alex Saka 1, Raphael H Valdivia 1
PMCID: PMC3202608  NIHMSID: NIHMS164550  PMID: 20006538

Summary

The Chlamydiae are obligate intracellular pathogen that replicate within a membrane-bound vacuole, termed the “inclusion”. From this compartment, bacteria acquire essential nutrients by selectively redirecting transport vesicles and hijacking intracellular organelles. Re-routing is achieved by several mechanisms including proteolysis-mediated fragmentation of the Golgi apparatus, recruitment of Rab GTPases and SNAREs, and translocation of cytoplasmic organelles into the inclusion lumen. Given Chlamydiae’s extended co-evolution with eukaryotic cells, it is likely that co-option of multiple cellular pathways is a strategy to provide redundancy in the acquisition of essential nutrients from the host and has contributed to the success of these highly adapted pathogens.

Introduction

The Chlamydiaceae comprise a distinct family of closely related, obligate intracellular bacteria that diverged early in evolution and parasitize a wide range of hosts [1]. The human pathogens are widely distributed and include Chlamydia trachomatis (leading cause of genital infections and infectious blindness), C. pneumoniae (common cause of respiratory tract infections and community-acquired pneumonia), and C. psittaci (a zoonotic pathogen that can cause severe pneumonia in humans) [26].

All Chlamydiaceae undergo a biphasic developmental cycle involving the infectious and environmentally resistant form or elementary body (EB), and the replicative, non-infectious form or reticulate body (RB). After entering their target eukaryotic cells, EBs differentiate into RBs, and replicate within a membrane-bound parasitophorous vacuole, termed the “inclusion”. Finally, RBs differentiate back into EBs, which are released to the extracellular medium to infect neighboring cells [7].

Chlamydia species have undergone massive genome condensation and lack in several biosynthetic pathways [812], suggesting that they have acquired compensatory mechanisms to allow the import and incorporation of nucleotides, amino acids, lipids and other nutrients from the host cell [1318]. Though it is a crucial step in their pathogenesis, the molecular mechanisms used by Chlamydia species to acquire nutrients are poorly characterized largely due to the lack of a system for mutational analysis and cell-free cultivation methods.

In this review, we discuss our understanding of Chlamydia nutrient acquisition pathways, based on information from genome sequencing and new cell biological observations detailing the extensive interactions between the inclusion and host organelles.

Nutrient uptake systems

The current estimation of Chlamydia’s metabolic capacity is founded on hints from genomic sequencing. As for any living organism, Chlamydia needs pyrimidine and purine nucleotides for energy transduction and nucleic acid biosynthesis [19,20], but is unable to synthesize them de novo. Chlamydia encodes enzymes that can generate ATP via substrate level phosphorylation, and CTP from UTP through a CTP synthetase; however, these organisms still require host cell–derived ATP, GTP and UTP [10,12,14,15,21,22]. The bacteria import these nucleotides by an unusual transport system that is found only in a small number of obligate intracellular bacteria and plant plastids [2325]. C. trachomatis has at least two nucleotide transport proteins (Npt): Npt1 and Npt2 [26]. Npt1 mediates the import of ATP from the host cell into the bacteria coupled with the export of ADP. Npt2 catalyzes the uptake of GTP, UTP, CTP and ATP, in a proton-dependent manner [26]. These transport systems are found in several Chlamydia species, Chlamydia-related amoeba symbionts, and Rickettsia [8,12,2629]. Chlamydia species differ in their ability to metabolize nucleotides. For instance, C. muridarum harbors guaAB-add and upp genes whose predicted products enable ATP to GTP conversion and the uracil phosphoribosyl tranferase-mediated biosynthesis of UTP from uracil, respectively. C. pneumoniae encodes udk, which may mediate uridine kinase-dependent synthesis of UTP from uracil, a UMP synthetase (PyrE), and a non-functional guaAB-add operon [11,12]. C. caviae encodes an intact guaAB-add operon and pyrE [11]. Chlamydia also lacks the components for NAD+ synthesis, indicating that this essential molecule must be scavenged from the host as has been demonstrated in the amoeba symbiont Parachlamydia UWE25 [23].

Genomic comparisons reveal that Chlamydia has several incomplete amino acid biosynthesis pathways [8,1012,30]. Not surprisingly, Chlamydia also contains several amino acid transporters including aat (neutral amino acid transporter), xasA (amino acid antiporter), brnQ-like (branched amino acid transporter), and a substantial number of ABC transporters (at least 13 in C. trachomatis) that are likely associated with amino acid and oligopeptide transport [8,1012,31]. The regulation of tryptophan biosynthesis is of particular interest for Chlamydia infections. Interferon-gamma (IFN-γ) secreted by immune cells activates indoleamine 2,3-dioxygenase, which inhibits chlamydial replication by depleting intracellular pools of tryptophan [32]. The ability of different Chlamydia species to synthesize or to acquire tryptophan precursors correlates with their susceptibility to IFN-γ-mediated killing and is linked to tissue tropism [3335]. IFN-γ inhibits the replication of C. pneumoniae, C. muridarum, and ocular strains of C. trachomatis, but has limited effect on genital strains that contain genes for converting scavenged indole tryptophan [33]. Indeed, the anti-chlamydial activity of IFN-γ in human cells can be abrogated in vitro by addition of tryptophan to the culture media [35]. Although Chlamydia has several efficient and specialized uptake systems, these systems are confined to bacterial membranes. How nutrients cross from the host cytoplasm through the inclusion membrane, which is not permissive to the diffusion of molecules > 520 Da [36], is largely unknown.

Co-option of host organelles and trafficking pathways promote delivery of nutrients to the inclusion

One potential mechanism for nutrient acquisition may involve the interception of vesicular transport intermediates. Even though the chlamydial inclusion is predominantly segregated from classical endo/lysosomal transport pathways, it can exploit membrane trafficking events [3744] and host lipases [45] to acquire lipids. The molecular basis for these events is unclear, but a subset of Rab GTPases and SNARE proteins, which regulate membrane trafficking, are recruited to the inclusion membrane [4650]. Indeed, recent findings indicate that interaction with multivesicular bodies, the Golgi-apparatus, mitochondria, and lipid droplets, are required for optimal chlamydial replication [5157] and suggests that membrane fusion events between host-derived vesicles and the inclusion may deliver nutrients to Chlamydia.

Interception of exocytic vesicles

Chlamydia acquires eukaryote-specific lipids [58] from the host cell. Sphingomyelin, glucosylceramide and cholesterol are delivered to the inclusion, and subsequently incorporated into chlamydial membranes [4345,59,60]. Cholesterol and sphingolipid transport to C. trachomatis are brefeldin A-sensitive and microtubule-dependent. Since brefeldin A inhibits anterograde vesicular traffic from the Golgi, Golgi-derived exocytic vesicles are likely involved in the delivery of these lipids to the bacteria. Furthermore, chlamydial protein synthesis is required for sphingomyelin uptake, suggesting that co-option of these vesicles is a bacteria-driven process [59]. The utilization of exocytic pathways is more pronounced when analyzing Chlamydia infection of polarized epithelial cells [40]. In these cells sphingomyelin, but not glucosylceramide, is retained by the Chlamydia inclusion and EBs due to preferential interception of basolaterally trafficked exocytic vesicles. However, the bulk protein cargo normally present within Golgi-derived secretory vesicles does not appear to accumulate in the inclusion, suggesting that a subset of vesicles may be preferentially targeted. The requirement of Golgi-derived vesicles for chlamydial replication is controversial because brefeldin A does not inhibit chlamydial replication [43]. However, RNAi screens identified COPI coat proteins as important for chlamydial growth [57,61]. The exact role of COPI in in bacterial survival, remains to be determined. It is possible that a subpopulation of brefeldin A-insensitive COPI vesicles deliver nutrients to the inclusion, or that other COPI-dependent membrane trafficking events (endosomal sorting, lipid droplet biogenesis) are required by Chlamydia. Recently, the GPI-anchored plasma membrane protein CD59 was found to traffic to the inclusion in a brefeldin A-insensitive manner, which suggests that Chlamydia also exploits Golgi-independent pathways to acquire nutrients [62].

Chlamydia-mediated reprogramming of the Golgi apparatus

The Golgi-apparatus processes and sorts newly synthesized proteins and lipids to their subcellular destinations [63]. Recent findings by Heuer and coworkers detail the importance of Golgi architecture in C. trachomatis replication [53]. At early stages of infection, Golgi ribbon-like structures of normal morphology are found closely associated with the chlamydial inclusion. These structures progressively fragment into ministacks during the course of infection and fragmentation correlates with the cleavage of the Golgi protein golgin-84. The processing of golgin-84 is sensitive to caspase and calpain inhibitors, but it is unclear if secreted bacterial proteases also contribute to this process. Nonetheless, disruption of golgin-84 processing with the caspase inhibitor,Z-WEHD-FMK impaired both bacterial replication and sphingolipid transport into the inclusion. Moreover, when Golgi-fragmentation was induced by knockdown of different Golgi matrix proteins, bacterial replication was increased. These results demonstrate that Golgi-fragmentation enhances chlamydial replication and is required for the efficient transport of sphingolipids into the inclusion.

Recruitment of Rab GTPases and SNARE proteins: subversion of vesicular trafficking

Rab GTPases are pivotal regulators of membrane trafficking processes and organelle identity [64] and SNARE proteins are key components of the intracellular membrane fusion machinery [65]. Scidmore and colleagues were the first to show that a subset Rab GTPases are recruited to the inclusion [46]. Rab1, 4 and 11 are recruited to inclusions of all chlamydial species tested, while recruitment of Rab6 is species-specific for C. trachomatis and Rab10 associates with both C. pneumoniae and C. muridarum. In addiiton, in RNAi screens several Rabs have been identified as important factors required for Chlamydia replication and development [57,61]. Recent observation indicate that this is partially due to the role of Rab GTPases in remodeling the Golgi-apparatus [54]. Depletion of Rab6 or Rab11 by RNAi inhibited Chlamydia-induced Golgi-fragmentation despite processing of golgin-84, suggesting that these Rabs act downstream of golgin-84 cleavage. Interestingly, the impairment of chlamydial replication in Rab6 or Rab11 silenced cells is reversed when Golgi-fragmentation is induced through disruption of the Golgi-tethering protein p115. These findings provide additional evidence for a functional link between Golgi-fragmentation and chlamydial replication.

How Chlamydia mediates recruitment of Rab proteins is not fully understood. A family of integral inclusion membrane proteins (Incs) are attractive candidates for bacterial factors that may facilitate interaction between Rabs and the inclusion membrane [66]. The C. trachomatis Inc CT229 interacts with Rab4 [47], and the C. pneumoniae inclusion membrane protein Cpn0585 interacts with Rabs 1, 10, and 11 [48]. These interactions are important for bacterial development as ectopic overexpression of Cpn0585 in C. pneumoniae-infected cells [48] or the microinjection of anti-CT229 in C. trachomatis-infected cells [47] negatively impact bacterial replication.

Like Rab GTPases, a group of SNARE proteins is targeted to the inclusion. Subtil and coworkers determined that the R-SNAREs Vamp3, Vamp7, and Vamp8 preferentially localize around the inclusion [49]. Through bioinformatics and structural modeling, SNARE-like motifs were identified in at least three C. trachomatis Incs (IncA, CT813 and CT223) suggesting the potential for interaction with host SNAREs. IncA interacts with Vamp3, Vamp7 and Vamp8, and SNARE recruitment to the inclusion is reduced in C. trachomatis isolates lacking IncA. Furthermore, ectopically expressed CT813 interacts with Vamp-7 and Vamp-8 in co-immunoprecipitation assays. Importantly, the SNARE-like motif of these proteins was essential for these interactions. More recently, recombinant IncA was shown to block SNARE-mediated membrane fusion in an in vitro liposome fusion assay [50]. These findings lead to the speculation that Chlamydia subverts host SNARE-mediated fusion events by expressing inhibitory SNARE-like proteins [50].

Modulation of signaling pathways

Chlamydial membranes contain lipid species that are normally associated with eukaryotic membranes [67]. Using CHO cell lines deficient in various phospholipid components, the lipid composition of chlamydial membranes was found to closely mimic that of the host [67]. The bacteria modify host-derived glycerophospholipids by replacing the straight chain fatty acid at the sn2 position with a bacteria-derived branched chain fatty acid [17]. The sn2 deacylation of phospholipids requires phospholipase A2 (PLA2) activity. Because chlamydial genomes do not encode any obvious PLA2 homologues, it is speculated that Chlamydia relies on a host PLA2 [45]. In support of this, pharmacological inhibition of the Ca2+-dependent cytosolic PLA2 (cPLA2) prevents Chlamydia uptake of host-derived phospholipids and severely limits replication [45]. Furthermore, cPLA2 is activated by the extracellular signal-regulated map kinase (ERK1/2) which is fully activated during chlamydial infections [45]. These results suggest that Chlamydia manipulates ERK/cPLA2 signaling pathways to facilitate the acquisition of glycerophospholipids. However, this story is likely more complex than originally proposed, as components of these signaling pathways have not been recognized as necessary for infection by RNAi-based screens. Indeed, we have found through genetic ablation experiments that cPLA2 has an innate immune signaling function in Chlamydia-infected cells, and that these processes vary between host cell species (unpublished observation).

Targeting of endosomal compartments

Multivesicular bodies (MVB) are a specialized subset of late endosomes involved in the sorting of cargo proteins and lipids to lysosomes for degradation [68,69]. By immunofluorescence and immunoelectron microscopy, two different MVB markers, CD63 and MLN64, and the lipid LBPA (lysobisphosphatidic acid, highly enriched in intra-luminal vesicles of MVBs) were found in the inclusion lumen. Inhibition of MVB biogenesis with pharmacological inhibitors or by exogenous addition of anti-CD63 antibodies impairs chlamydial replication and disrupts the traffic of sphingomyelin and cholesterol into the inclusion [51,52]. However, RNAi-mediated silencing of CD63 synthesis does not prevent MVB interactions with the inclusion [52], and MVB components have not been identified in the host genome RNAi screens [57,61]. It is likely that Golgi and MVBs partially overlap in their ability to deliver essential nutrients to replicating Chlamydia. This would explain why disruption of individual components of either pathway is not sufficient to inhibit chlamydial replication.

Co-option of lipid droplets

Lipid droplets (LD) are endoplasmic reticulum-derived organelles composed of a neutral lipid core surrounded by a phospholipid monolayer [70]. These lipid storage compartments are being increasingly recognized as dynamic organelles involved in a multiple biological processes [71]. These organelles proliferate and are recruited to the periphery of the inclusion during Chlamydia infection [55,56]. Three chlamydial proteins (Lda1 to Lda3) were identified as potentially involved in targeting LDs based on their ability to localize to LDs when expressed exogenously in mammalian cells [55]. Strikingly, LDs enter into the inclusion lumen [56]. These observations are significant because they demonstrate that intact organelles can be translocated into the inclusion and that vesicle fusion is not the only means to deliver potential nutrients into the inclusion lumen. What role these organelles play in Chlamydia biology is less clear, although pharmacological inhibition of neutral lipid biosynthesis, and thus LD biogenesis, impairs chlamydial development, suggesting a function for these organelles in Chlamydia pathogenesis [55].

Conclusions

In contrast with the numerous trafficking pathways intercepted by Chlamydiae during infection contrasts, little is known about the bacterial factors which mediate these processes. The molecular underpinnings of these complex host-pathogen interactions may be difficult to decipher without the development of a robust system for mutational analysis in Chlamydiae. Here we reviewed how some of the most recent cell biological findings present new potential mechanisms for Chlamydia acquisition of nutrients while sequestered in an intracellular vacuole (Figure 1). Like many highly adapted and successful pathogens we predict that Chlamydia has evolved to tap several redundant mechanisms to obtain nutrients. However, it should also be noted that the host cell is not a passive partner in this interaction and actively seeks to destroy the microbial invader. Therefore, the rerouting of membrane and protein transport events cannot be solely viewed as a “feeding” mechanism but also as bacterial countermeasure against innate immune functions.

Figure 1.

Figure 1

Schematic representation of host cell pathways exploited by Chlamydiae to acquire nutrients. (A) Golgin-84 is proteolytically processed during infection through activation of host caspases, calpains and potentially by bacterial protease(s). This leads to Golgi fragmentation into ribbon-like structures in a process mediated by Rab-6 and Rab-11. These Golgi ministacks are recruited around the inclusion. Formation of Golgi-ministacks is necessary for chlamydial uptake of host sphingolipids and replication. Golgi-derived exocytic vesicles containing sphingolipids and cholesterol fuse with the inclusion. (B) GPI-anchored plasma membrane protein CD59 is trafficked to the inner face of the inclusion in a Golgi-independent manner. (C) Infection leads to increased phosphorylation of extracellular signal-regulated map kinase (ERK) which, in turn, activates calcium-dependent cytosolic phospholipase A2 (cPLA2). cPLA2 removes straight chain fatty acids from the sn2 position of host glycerophospholipids to generate lyso-phospholipids (Lyso-PL). Chlamydia-derived branched fatty acids (b-FA) are incorporated into Lyso-PL to generate Chlamydia-modified phospholipids (Chl-PL). (D) Chlamydia targets multivesicular bodies (MVB) to sequester required nutrients like sphingolipids and cholesterol. (E) Lipid droplets, the main store of neutral lipids, are targeted by Chlamydia lipid droplet-associated proteins (Lda) and translocated into the inclusion. (F) Pan-species recruitment of Rab-1, Rab-4, and Rab-11, whereas Rab-6 and Rab-10 are recruited to C. trachomatis and C. pneumoniae inclusions, respectively. Candidate bacterial recruitment factors have been identified (Cpn0585 and CT229). (G) SNARE proteins Vamp-3, 7 and 8 are recruited to the inclusion, possibly by chlamydial proteins containing SNARE-like motifs (like incA and CT813) in the inclusion membrane.

Acknowledgements

We thank J. Cocchiaro and J.D. Dunn for comments and feedback on this manuscript. Due to space limitations, not all contributions to this field could be included in this review and we apologize to those investigators whose work was not cited. Work in our laboratory is supported by grants from the NIH and the Burroughs Wellcome Trust Program in Infectious Diseases. H.A. Saka is supported by a fellowship from the Pew Latin American Fellows Program.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References and recommended reading

Papers of special (•) and outstanding (••) interest

  • 1.Horn M, Collingro A, Schmitz-Esser S, Beier CL, Purkhold U, Fartmann B, Brandt P, Nyakatura GJ, Droege M, Frishman D, et al. Illuminating the evolutionary history of chlamydiae. Science. 2004;304:728–730. doi: 10.1126/science.1096330. [DOI] [PubMed] [Google Scholar]
  • 2.Bebear C, de Barbeyrac B. Genital Chlamydia trachomatis infections. Clin Microbiol Infect. 2009;15:4–10. doi: 10.1111/j.1469-0691.2008.02647.x. [DOI] [PubMed] [Google Scholar]
  • 3.Blasi F, Tarsia P, Aliberti S. Chlamydophila pneumoniae. Clin Microbiol Infect. 2009;15:29–35. doi: 10.1111/j.1469-0691.2008.02130.x. [DOI] [PubMed] [Google Scholar]
  • 4.Beeckman DS, Vanrompay DC. Zoonotic Chlamydophila psittaci infections from a clinical perspective. Clin Microbiol Infect. 2009;15:11–17. doi: 10.1111/j.1469-0691.2008.02669.x. [DOI] [PubMed] [Google Scholar]
  • 5.Stephens RS, Myers G, Eppinger M, Bavoil PM. Divergence without difference: phylogenetics and taxonomy of Chlamydia resolved. FEMS Immunol Med Microbiol. 2009;55:115–119. doi: 10.1111/j.1574-695X.2008.00516.x. [DOI] [PubMed] [Google Scholar]
  • 6.Schachter J. Chlamydia: Intracellular Biology, Pathogenesis and Immunity. In: Stephens RS, editor. Infection and Disease Epidemiology. A.S.M.; 1999. pp. 139–169. [Google Scholar]
  • 7.Valdivia RH. Chlamydia effector proteins and new insights into chlamydial cellular microbiology. Curr Opin Microbiol. 2008;11:53–59. doi: 10.1016/j.mib.2008.01.003. [DOI] [PubMed] [Google Scholar]
  • 8.Kalman S, Mitchell W, Marathe R, Lammel C, Fan J, Hyman RW, Olinger L, Grimwood J, Davis RW, Stephens RS. Comparative genomes of Chlamydia pneumoniae and C. trachomatis. Nat Genet. 1999;21:385–389. doi: 10.1038/7716. [DOI] [PubMed] [Google Scholar]
  • 9.Moran NA. Microbial minimalism: genome reduction in bacterial pathogens. Cell. 2002;108:583–586. doi: 10.1016/s0092-8674(02)00665-7. [DOI] [PubMed] [Google Scholar]
  • 10.Stephens RS, Kalman S, Lammel C, Fan J, Marathe R, Aravind L, Mitchell W, Olinger L, Tatusov RL, Zhao Q, et al. Genome sequence of an obligate intracellular pathogen of humans: Chlamydia trachomatis. Science. 1998;282:754–759. doi: 10.1126/science.282.5389.754. [DOI] [PubMed] [Google Scholar]
  • 11.Read TD, Myers GS, Brunham RC, Nelson WC, Paulsen IT, Heidelberg J, Holtzapple E, Khouri H, Federova NB, Carty HA, et al. Genome sequence of Chlamydophila caviae (Chlamydia psittaci GPIC): examining the role of niche-specific genes in the evolution of the Chlamydiaceae. Nucleic Acids Res. 2003;31:2134–2147. doi: 10.1093/nar/gkg321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Read TD, Brunham RC, Shen C, Gill SR, Heidelberg JF, White O, Hickey EK, Peterson J, Utterback T, Berry K, et al. Genome sequences of Chlamydia trachomatis MoPn and Chlamydia pneumoniae AR39. Nucleic Acids Res. 2000;28:1397–1406. doi: 10.1093/nar/28.6.1397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Moulder JW. Interaction of chlamydiae and host cells in vitro. Microbiol Rev. 1991;55:143–190. doi: 10.1128/mr.55.1.143-190.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.McClarty G, Fan H. Purine metabolism by intracellular Chlamydia psittaci. J Bacteriol. 1993;175:4662–4669. doi: 10.1128/jb.175.15.4662-4669.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.McClarty G, Qin B. Pyrimidine metabolism by intracellular Chlamydia psittaci. J Bacteriol. 1993;175:4652–4661. doi: 10.1128/jb.175.15.4652-4661.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.McClarty G. Chlamydiae and the biochemistry of intracellular parasitism. Trends Microbiol. 1994;2:157–164. doi: 10.1016/0966-842x(94)90665-3. [DOI] [PubMed] [Google Scholar]
  • 17.Wylie JL, Hatch GM, McClarty G. Host cell phospholipids are trafficked to and then modified by Chlamydia trachomatis. J Bacteriol. 1997;179:7233–7242. doi: 10.1128/jb.179.23.7233-7242.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Cocchiaro JL, Valdivia RH. New insights into Chlamydia intracellular survival mechanisms. Cell Microbiol. 2009;11:1571–1578. doi: 10.1111/j.1462-5822.2009.01364.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Zalkin H, Nygaard P. Biosynthesis of purine nucleotides. In: Neidhardt FCCR III, Ingraham JL, Lin ECC, Low KB, Magasanik B, Reznikoff WS, Riley M, Schaechter M, Umbarger HE, editors. Escherichia coli and Salmonella: Cellular and Molecular Biology. edn 2nd. A. S. M.; 1996. pp. 561–579. [Google Scholar]
  • 20.Neuhard J, Kelln RA. Biosynthesis and conversion of pyrimidines. In: Neidhardt FCCR III, Ingraham JL, Lin ECC, Low KB, Magasanik B, Reznikoff WS, Riley M, Schaechter M, Umbarger HE, editors. Escherichia coli and Salmonella: Cellular and Molecular Biology. edn 2nd. A. S. M.; 1996. pp. 580–599. vol.] [Google Scholar]
  • 21.Tipples G, McClarty G. The obligate intracellular bacterium Chlamydia trachomatis is auxotrophic for three of the four ribonucleoside triphosphates. Mol Microbiol. 1993;8:1105–1114. doi: 10.1111/j.1365-2958.1993.tb01655.x. [DOI] [PubMed] [Google Scholar]
  • 22.Wylie JL, Berry JD, McClarty G. Chlamydia trachomatis CTP synthetase: molecular characterization and developmental regulation of expression. Mol Microbiol. 1996;22:631–642. doi: 10.1046/j.1365-2958.1996.d01-1717.x. [DOI] [PubMed] [Google Scholar]
  • 23.Haferkamp I, Schmitz-Esser S, Linka N, Urbany C, Collingro A, Wagner M, Horn M, Neuhaus HE. A candidate NAD+ transporter in an intracellular bacterial symbiont related to Chlamydiae. Nature. 2004;432:622–625. doi: 10.1038/nature03131. [DOI] [PubMed] [Google Scholar]
  • 24.Linka N, Hurka H, Lang BF, Burger G, Winkler HH, Stamme C, Urbany C, Seil I, Kusch J, Neuhaus HE. Phylogenetic relationships of non-mitochondrial nucleotide transport proteins in bacteria and eukaryotes. Gene. 2003;306:27–35. doi: 10.1016/s0378-1119(03)00429-3. [DOI] [PubMed] [Google Scholar]
  • 25.Winkler HH, Neuhaus HE. Non-mitochondrial ATP transport. Trends Biochem Sci. 1999;24:64–68. doi: 10.1016/s0968-0004(98)01334-6. [DOI] [PubMed] [Google Scholar]
  • 26.Tjaden J, Winkler HH, Schwoppe C, Van der Laan M, Mohlmann T, Neuhaus HE. Two nucleotide transport proteins in Chlamydia trachomatis, one for net nucleoside triphosphate uptake and the other for transport of energy. Journal of Bacteriology. 1999;181:1196–1202. doi: 10.1128/jb.181.4.1196-1202.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27. Trentmann O, Horn M, van Scheltinga AC, Neuhaus HE, Haferkamp I. Enlightening energy parasitism by analysis of an ATP/ADP transporter from chlamydiae. PLoS Biol. 2007;5:e231. doi: 10.1371/journal.pbio.0050231. In this study, a detailed biochemical analysis of an ATP/ADP transporter (PamNTT1), from Protochlamydia amoebophila (UWE25) was performed. This paper contains the first successful purification and functional reconstitution of a bacterial nucleotide transporter and provides the functional basis of how an obligate intracellular parasite exploits the energy pool of its host cell.
  • 28.Schmitz-Esser S, Linka N, Collingro A, Beier CL, Neuhaus HE, Wagner M, Horn M. ATP/ADP translocases: a common feature of obligate intracellular amoebal symbionts related to Chlamydiae and Rickettsiae. J Bacteriol. 2004;186:683–691. doi: 10.1128/JB.186.3.683-691.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Haferkamp I, Schmitz-Esser S, Wagner M, Neigel N, Horn M, Neuhaus HE. Tapping the nucleotide pool of the host: novel nucleotide carrier proteins of Protochlamydia amoebophila. Mol Microbiol. 2006;60:1534–1545. doi: 10.1111/j.1365-2958.2006.05193.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Heizer EM, Jr, Raiford DW, Raymer ML, Doom TE, Miller RV, Krane DE. Amino acid cost and codon-usage biases in 6 prokaryotic genomes: a whole-genome analysis. Mol Biol Evol. 2006;23:1670–1680. doi: 10.1093/molbev/msl029. [DOI] [PubMed] [Google Scholar]
  • 31.Braun PR, Al-Younes H, Gussmann J, Klein J, Schneider E, Meyer TF. Competitive inhibition of amino acid uptake suppresses chlamydial growth: involvement of the chlamydial amino acid transporter BrnQ. J Bacteriol. 2008;190:1822–1830. doi: 10.1128/JB.01240-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Rottenberg ME, Gigliotti-Rothfuchs A, Wigzell H. The role of IFN-gamma in the outcome of chlamydial infection. Curr Opin Immunol. 2002;14:444–451. doi: 10.1016/s0952-7915(02)00361-8. [DOI] [PubMed] [Google Scholar]
  • 33.Fehlner-Gardiner C, Roshick C, Carlson JH, Hughes S, Belland RJ, Caldwell HD, McClarty G. Molecular basis defining human Chlamydia trachomatis tissue tropism. A possible role for tryptophan synthase. J Biol Chem. 2002;277:26893–26903. doi: 10.1074/jbc.M203937200. [DOI] [PubMed] [Google Scholar]
  • 34.McClarty G, Caldwell HD, Nelson DE. Chlamydial interferon gamma immune evasion influences infection tropism. Curr Opin Microbiol. 2007;10:47–51. doi: 10.1016/j.mib.2006.12.003. [DOI] [PubMed] [Google Scholar]
  • 35.Caldwell HD, Wood H, Crane D, Bailey R, Jones RB, Mabey D, Maclean I, Mohammed Z, Peeling R, Roshick C, et al. Polymorphisms in Chlamydia trachomatis tryptophan synthase genes differentiate between genital and ocular isolates. J Clin Invest. 2003;111:1757–1769. doi: 10.1172/JCI17993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Heinzen RA, Hackstadt T. The Chlamydia trachomatis parasitophorous vacuolar membrane is not passively permeable to low-molecular-weight compounds. Infect Immun. 1997;65:1088–1094. doi: 10.1128/iai.65.3.1088-1094.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Taraska T, Ward DM, Ajioka RS, Wyrick PB, Davis-Kaplan SR, Davis CH, Kaplan J. The late chlamydial inclusion membrane is not derived from the endocytic pathway and is relatively deficient in host proteins. Infect Immun. 1996;64:3713–3727. doi: 10.1128/iai.64.9.3713-3727.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Heinzen RA, Scidmore MA, Rockey DD, Hackstadt T. Differential interaction with endocytic and exocytic pathways distinguish parasitophorous vacuoles of Coxiella burnetii and Chlamydia trachomatis. Infect Immun. 1996;64:796–809. doi: 10.1128/iai.64.3.796-809.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Scidmore MA, Fischer ER, Hackstadt T. Restricted fusion of Chlamydia trachomatis vesicles with endocytic compartments during the initial stages of infection. Infect Immun. 2003;71:973–984. doi: 10.1128/IAI.71.2.973-984.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40. Moore ER, Fischer ER, Mead DJ, Hackstadt T. The chlamydial inclusion preferentially intercepts basolaterally directed sphingomyelin-containing exocytic vacuoles. Traffic. 2008;9:2130–2140. doi: 10.1111/j.1600-0854.2008.00828.x. In this paper, the authors developed a polarized epithelial cell model to study how Chlamydia affects vectoral trafficking of lipids and proteins to the inclusion. C2BBe1 polarized colonic cells were infected with C. trachomatis and the traffic of ceramide and its metabolic derivatives, glucosylceramide and sphingomyelin, was evaluated. Results support that, as apposed to glucosylceramide, sphingomyelin was retained in the chlamydial inclusion and incorporated into elementary bodies. This retention of sphingomyelin correlated with a disruption of basolateral trafficking.
  • 41.van Ooij C, Apodaca G, Engel J. Characterization of the Chlamydia trachomatis vacuole and its interaction with the host endocytic pathway in HeLa cells. Infect Immun. 1997;65:758–766. doi: 10.1128/iai.65.2.758-766.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Wolf K, Hackstadt T. Sphingomyelin trafficking in Chlamydia pneumoniae-infected cells. Cell Microbiol. 2001;3:145–152. doi: 10.1046/j.1462-5822.2001.00098.x. [DOI] [PubMed] [Google Scholar]
  • 43.Hackstadt T, Rockey DD, Heinzen RA, Scidmore MA. Chlamydia trachomatis interrupts an exocytic pathway to acquire endogenously synthesized sphingomyelin in transit from the Golgi apparatus to the plasma membrane. EMBO J. 1996;15:964–977. [PMC free article] [PubMed] [Google Scholar]
  • 44.Rockey DD, Fischer ER, Hackstadt T. Temporal analysis of the developing Chlamydia psittaci inclusion by use of fluorescence and electron microscopy. Infect Immun. 1996;64:4269–4278. doi: 10.1128/iai.64.10.4269-4278.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Su H, McClarty G, Dong F, Hatch GM, Pan ZK, Zhong G. Activation of Raf/MEK/ERK/cPLA2 signaling pathway is essential for chlamydial acquisition of host glycerophospholipids. J Biol Chem. 2004;279:9409–9416. doi: 10.1074/jbc.M312008200. [DOI] [PubMed] [Google Scholar]
  • 46.Rzomp KA, Scholtes LD, Briggs BJ, Whittaker GR, Scidmore MA. Rab GTPases are recruited to chlamydial inclusions in both a species-dependent and species-independent manner. Infect Immun. 2003;71:5855–5870. doi: 10.1128/IAI.71.10.5855-5870.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Rzomp KA, Moorhead AR, Scidmore MA. The GTPase Rab4 interacts with Chlamydia trachomatis inclusion membrane protein CT229. Infect Immun. 2006;74:5362–5373. doi: 10.1128/IAI.00539-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Cortes C, Rzomp KA, Tvinnereim A, Scidmore MA, Wizel B. Chlamydia pneumoniae inclusion membrane protein Cpn0585 interacts with multiple Rab GTPases. Infect Immun. 2007;75:5586–5596. doi: 10.1128/IAI.01020-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49. Delevoye C, Nilges M, Dehoux P, Paumet F, Perrinet S, Dautry-Varsat A, Subtil A. SNARE protein mimicry by an intracellular bacterium. PLoS Pathog. 2008;4 doi: 10.1371/journal.ppat.1000022. e1000022. The authors show that host SNARE proteins Vamp-3, Vamp-7 and Vamp-8 are recruited to the immediate vicinity of the inclusion membrane and that chlamydial inclusion membrane protein IncA has a dominant role in this recruitment. SNARE-like motifs were found in the inclusion membrane proteins incA and CT813 from Chlamydia trachomatis and these two proteins were shown to bind SNAREs. Moreover, SNARE-like motif present in incA was shown to be required for incA-SNARE binding.
  • 50. Paumet F, Wesolowski J, Garcia-Diaz A, Delevoye C, Aulner N, Shuman HA, Subtil A, Rothman JE. Intracellular bacteria encode inhibitory SNARE-like proteins. PLoS One. 2009;4:e7375. doi: 10.1371/journal.pone.0007375. In this paper, the investigators assess how different bacteria can influence eukaryotic membrane fusion. The results show that both, IncA from Chlamydia trachomatis and IcmG/DotF from Legionella pneuomphila, can inhibit inhibit the endocytic SNARE machinery. The inhibitory function was found to reside in the SNARE-like motifs present in these bacterial proteins.
  • 51.Beatty WL. Trafficking from CD63-positive late endocytic multivesicular bodies is essential for intracellular development of Chlamydia trachomatis. J Cell Sci. 2006;119:350–359. doi: 10.1242/jcs.02733. [DOI] [PubMed] [Google Scholar]
  • 52.Beatty WL. Late endocytic multivesicular bodies intersect the chlamydial inclusion in the absence of CD63. Infect Immun. 2008;76:2872–2881. doi: 10.1128/IAI.00129-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53. Heuer D, Lipinski AR, Machuy N, Karlas A, Wehrens A, Siedler F, Brinkmann V, Meyer TF. Chlamydia causes fragmentation of the Golgi compartment to ensure reproduction. Nature. 2009;457:731–735. doi: 10.1038/nature07578. This study shows that in Chlamydia-infected cells, the Golgi-apparatus is fragmented into ministacks and recruited to the parasitopherous vacuole. Golgi-fragmentation was associated with host caspases and calpains-mediated processing of the Golgi protein golgin-84. This phenomenon was required for lipid acquisition and replication of C. trachomatis.
  • 54. Lipinski AR, Heymann J, Meissner C, Karlas A, Brinkmann V, Meyer TF, Heuer D. Rab6 and Rab11 regulate Chlamydia trachomatis development and golgin-84-dependent Golgi fragmentation. PLoS Pathog. 2009;5 doi: 10.1371/journal.ppat.1000615. e1000615. By using RNA interference, these investigators show that depletion of Rab-6 or Rab-11, negatively impacted Chlamydia trachomatis replication. These Rab proteins were found to be required for Chlamydia-induced Golgi-fragmentation downstream of golgin-84 processing. Moreover, knocking down of Rab-6 and Rab-11 blocked sphingolipid transport to the chlamydial inclusion.
  • 55.Kumar Y, Cocchiaro J, Valdivia RH. The obligate intracellular pathogen Chlamydia trachomatis targets host lipid droplets. Curr Biol. 2006;16:1646–1651. doi: 10.1016/j.cub.2006.06.060. [DOI] [PubMed] [Google Scholar]
  • 56. Cocchiaro JL, Kumar Y, Fischer ER, Hackstadt T, Valdivia RH. Cytoplasmic lipid droplets are translocated into the lumen of the Chlamydia trachomatis parasitophorous vacuole. Proc Natl Acad Sci U S A. 2008;105:9379–9384. doi: 10.1073/pnas.0712241105. By means of live cell, confocal and electron microscopy, this paper provides evidence showing that host cell lipid droplets are recruited around and translocated int the chlamydial inclusion. The fact that intact organelles can be translocated into the inclusion demonstrate that besides vesicle fusion, another means of nutrients delivery into the inclusion lumen can occur.
  • 57. Derre I, Pypaert M, Dautry-Varsat A, Agaisse H. RNAi screen in Drosophila cells reveals the involvement of the Tom complex in Chlamydia infection. PLoS Pathog. 2007;3:1446–1458. doi: 10.1371/journal.ppat.0030155. This paper and [62] describe genome wide RNAi screens in Drosophila S2 cells for host factors required for chlamydial replication. This study reveals a role for mitochondria for efficient C. caviae infection.
  • 58.van Ooij C, Kalman L, van I, Nishijima M, Hanada K, Mostov K, Engel JN. Host cell-derived sphingolipids are required for the intracellular growth of Chlamydia trachomatis. Cell Microbiol. 2000;2:627–637. doi: 10.1046/j.1462-5822.2000.00077.x. [DOI] [PubMed] [Google Scholar]
  • 59.Carabeo RA, Mead DJ, Hackstadt T. Golgi-dependent transport of cholesterol to the Chlamydia trachomatis inclusion. Proc Natl Acad Sci U S A. 2003;100:6771–6776. doi: 10.1073/pnas.1131289100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Hackstadt T, Scidmore MA, Rockey DD. Lipid metabolism in Chlamydia trachomatis-infected cells: directed trafficking of Golgi-derived sphingolipids to the chlamydial inclusion. Proc Natl Acad Sci U S A. 1995;92:4877–4881. doi: 10.1073/pnas.92.11.4877. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Elwell CA, Ceesay A, Kim JH, Kalman D, Engel JN. RNA interference screen identifies Abl kinase and PDGFR signaling in Chlamydia trachomatis entry. PLoS Pathog. 2008;4:e1000021. doi: 10.1371/journal.ppat.1000021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62. Hasegawa A, Sogo LF, Tan M, Sutterlin C. Host complement regulatory protein CD59 is transported to the chlamydial inclusion by a Golgi apparatus-independent pathway. Infect Immun. 2009;77:1285–1292. doi: 10.1128/IAI.01062-08. This paper and [57] describe genome wide RNAi screens in Drosophila S2 cells for host factors required for chlamydial replication. This study reveals a role for c-Abl in chlamydial entry.
  • 63.De Matteis MA, Luini A. Exiting the Golgi complex. Nat Rev Mol Cell Biol. 2008;9:273–284. doi: 10.1038/nrm2378. [DOI] [PubMed] [Google Scholar]
  • 64.Stenmark H. Rab GTPases as coordinators of vesicle traffic. Nat Rev Mol Cell Biol. 2009;10:513–525. doi: 10.1038/nrm2728. [DOI] [PubMed] [Google Scholar]
  • 65.Sudhof TC, Rothman JE. Membrane fusion: grappling with SNARE and SM proteins. Science. 2009;323:474–477. doi: 10.1126/science.1161748. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Li Z, Chen C, Chen D, Wu Y, Zhong Y, Zhong G. Characterization of fifty putative inclusion membrane proteins encoded in the Chlamydia trachomatis genome. Infect Immun. 2008;76:2746–2757. doi: 10.1128/IAI.00010-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Hatch GM, McClarty G. Phospholipid composition of purified Chlamydia trachomatis mimics that of the eucaryotic host cell. Infect Immun. 1998;66:3727–3735. doi: 10.1128/iai.66.8.3727-3735.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Fader CM, Colombo MI. Autophagy and multivesicular bodies: two closely related partners. Cell Death Differ. 2009;16:70–78. doi: 10.1038/cdd.2008.168. [DOI] [PubMed] [Google Scholar]
  • 69.Russell MR, Nickerson DP, Odorizzi G. Molecular mechanisms of late endosome morphology, identity and sorting. Curr Opin Cell Biol. 2006;18:422–428. doi: 10.1016/j.ceb.2006.06.002. [DOI] [PubMed] [Google Scholar]
  • 70.Guo Y, Cordes KR, Farese RV, Jr, Walther TC. Lipid droplets at a glance. J Cell Sci. 2009;122:749–752. doi: 10.1242/jcs.037630. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Walther TC, Farese RV., Jr The life of lipid droplets. Biochim Biophys Acta. 2009;1791:459–466. doi: 10.1016/j.bbalip.2008.10.009. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES