Skip to main content
Advertisement
  • Loading metrics

"With a Little Help from My Friends"—Social Motility in Trypanosoma brucei

  • Edwin A. Saada,

    Current address: Sandia National Laboratories, Livermore, California, United States of America

    Affiliation Department of Microbiology, Immunology and Molecular Genetics, University of California, Los Angeles, Los Angeles, California, United States of America

  • Stephanie F. DeMarco,

    Affiliations Department of Microbiology, Immunology and Molecular Genetics, University of California, Los Angeles, Los Angeles, California, United States of America, Molecular Biology Institute, University of California, Los Angeles, Los Angeles, California, United States of America

  • Michelle M. Shimogawa,

    Affiliation Department of Microbiology, Immunology and Molecular Genetics, University of California, Los Angeles, Los Angeles, California, United States of America

  • Kent L. Hill

    kenthill@microbio.ucla.edu

    Affiliations Department of Microbiology, Immunology and Molecular Genetics, University of California, Los Angeles, Los Angeles, California, United States of America, Molecular Biology Institute, University of California, Los Angeles, Los Angeles, California, United States of America

Microbial Social Behavior: “The Whole Is Greater Than the Sum of Its Parts”

In their natural environments, microbes are not found in isolation, but live in groups where the ability to communicate and cooperate with friends, while thwarting activities of enemies, is of paramount importance [1,2]. The capacity for interaction among cells in a group allows for social behaviors, which present as emergent properties of the group as a whole that are not predicted from the sum of activities of individual cells. One example includes quorum sensing (QS), which enables cells in a population to coordinate gene expression and limit premature expenditure of resources. Other social behaviors include those occurring in the context of surfaces, such as biofilm formation and various forms of swarming motility across surfaces, as seen in bacteria and slime molds [35].

Recognizing cell–cell communication and social behaviors as ubiquitous among bacteria transformed our views of microbiology and microbial pathogenesis [1,2]. Parasitic protozoa cause tremendous human suffering worldwide and, although they engage in cell–cell interactions, the paradigm of social behavior is not commonly applied in studies of these organisms. One long-known social activity in protozoan parasites is QS-driven differentiation of the African trypanosome, Trypanosoma brucei, in the bloodstream of a mammalian host. During infection, proliferating T. brucei cells undergo cell-density dependent differentiation into quiescent forms that are then competent to be transmitted by a tsetse fly vector (Fig 1) [6,7]. More recently, the discovery of social motility (SoMo) in insect-stage T. brucei (Fig 1) [8] has highlighted the capacity of these organisms for group-level behavior, and several recent studies of SoMo emphasize the potential for concepts underlying social behavior to provide insight into parasite biology.

thumbnail
Fig 1. Trypanosomes are social.

Trypanosome cell–cell interactions operate in bloodstream and insect stage parasites. In the bloodstream, “long slender form” parasites (red) differentiate into growth-arrested “short stumpy forms” (purple) through a quorum sensing-mediated mechanism. Stumpy parasites are pre-adapted for the tsetse fly environment and the transition thus establishes transmission competence, while also limiting bloodstream parasitemia [52,53]. Procyclic T. brucei (insect midgut stage, blue) undergo social motility when cultivated on semi-solid surfaces, using cell–cell signaling to promote collective motility across the surface and coordinating their movements in response to extracellular signals from nearby parasites. This leads to formation of radial projections that extend outward from the initial site of inoculation [8]. These activities are hypothesized to support colonization and/or transit of tissue surfaces in the fly. Beyond their direct impact on understanding parasite development, recent studies of stumpy formation and social motility have provided insight into parasite signal transduction. See text for details.

https://doi.org/10.1371/journal.ppat.1005272.g001

Discovery of Social Motility in Trypanosoma brucei

African trypanosomes cause sleeping sickness in humans and nagana in livestock. They impose a substantial medical and economic burden across nearly 30 countries of sub-Saharan Africa, where approximately 70 million people live at risk of infection [911]. The parasites are transmitted between mammalian hosts by blood-feeding tsetse flies. Most studies and animal infection models consider T. brucei as individual cells in suspension, yet in their natural environments these parasites frequently live in contact with tissue surfaces. This is most evident in the tsetse fly, where trypanosomes undergo extensive movements across fly tissues, culminating in colonization of the salivary gland epithelium and subsequent differentiation into human-infectious forms [12].

To investigate the influence of surface interactions on trypanosome biology, Oberholzer and Lopez et al. cultivated procyclic form (insect midgut stage) T. brucei on semisolid agarose [8]. They discovered that parasites assemble into groups of cells that undergo collective movements, forming multicellular projections that radiate outward from the site of inoculation. Although individual parasites can move freely within a group, movement of the group is polarized such that it advances at a single, leading edge. Moreover, groups alter their movements when they sense other parasites nearby. The combined data reveal that procyclic trypanosomes can sense extracellular signals and coordinate their activities in response to these signals. This behavior was termed “social motility” based on analogies to social motility and other surface-associated motilities in bacteria.

Why Be Social?

In bacteria, social behaviors provide numerous advantages, including enhanced ability to colonize, transit, and penetrate surfaces; enhanced protection from host defenses; improved accessibility to nutrients; and opportunities for genetic exchange [1,2,5,13,14]. Although in vivo manifestations and ramifications of T. brucei social motility are not yet known, the parasites can foreseeably benefit from the same advantages afforded to bacteria.

To complete their transmission through the tsetse fly, trypanosomes must undertake an epic journey fraught with hazards [1517]. Under optimized laboratory conditions, less than 20% of fly infections yield mammalian-infectious parasites, and rates in the field are even lower [17,18]. The journey begins when ingested bloodstream parasites differentiate into procyclic-stage parasites that colonize the midgut lumen and must then get through or around the peritrophic matrix (PM) to establish infection on the midgut epithelium. The PM is a chitinous lining that protects the gut epithelium from digestive enzymes and presents a formidable barrier to T. brucei infection [19,20]. Once beyond the PM, parasites must penetrate the proventriculus and move along the alimentary tract to the mouth parts. From there, they move up the salivary duct to the salivary gland, where they colonize the gland epithelium and complete differentiation into human-infectious forms. Along the way, T. brucei must compete for resources while overcoming harsh conditions and fly defenses, including antimicrobial peptides and lectins [16,21].

Extensive surface colonization and the arduous nature of parasite movement through the fly present bottlenecks to infection [16,22] and led Oberholzer and Lopez et al. to propose that SoMo supports one or more of these in vivo activities [8]. Recent work from the Roditi group has strengthened this hypothesis [23,24]. Prior work defined parasites appearing within the first few days of tsetse infection as “early” procyclics, while “late” procyclics are those that appear at later time points [25], after parasites have persistently colonized the ectoperitrophic space. Early and late procyclics can be distinguished by differences in surface protein expression [23]. Imhof and colleagues discovered that SoMo initiation is cell density-dependent and that early procyclics cultured in vitro undergo social motility, while late procyclics do not. Importantly, they showed that SoMo competence is linked to the early developmental stage, rather than to the presence of specific surface proteins that distinguish between early and late developmental stages. The precise point at which the switch from early to late procyclics occurs is not yet clear but is hypothesized to correlate with infiltration of the ectoperitrophic space, suggesting that SoMo may be a property of cells that colonize and penetrate the peritrophic matrix. Further evidence linking SoMo to successful midgut infection comes from studies of a null mutant of the Requires Fifty Three (RFT1) gene, which is required for N-glycosylation of parasite proteins [26]. RFT1 null mutants show defective SoMo in vitro and exhibit reduced and delayed establishment of midgut infections in the fly [24]. Thus, these combined studies establish SoMo as a property of a specific parasite developmental stage and correlate the in vitro behavior to a critical step of the in vivo transmission cycle.

More Than an Oar: The Trypanosome Flagellum Contains cAMP Signaling Systems That Control Social Motility

A universal requirement of microbial social behavior is the ability of cells to sense and respond to extracellular signals, and SoMo therefore provides a novel opportunity for dissecting trypanosome signal transduction. This presents an important advance because perception and transduction of extracellular signals are important features of parasite biology [27], but underlying mechanisms are poorly understood. In addition to a well-known role in motility, the conserved role of the eukaryotic flagellum (also known as cilium) as a signaling center has emerged as a unifying theme in vertebrate development and pathophysiology of human genetic diseases [28]. This concept, together with the prominent role of cyclic nucleotides in cilium-dependent signaling and in control of microbial social behaviors [2932], made flagellar cyclic-AMP (cAMP) a focal point for interrogation of SoMo signaling mechanisms.

Cellular cAMP levels are controlled through opposing activities of adenylate cyclases (ACs) and cAMP-specific phosphodiesterases (PDEs). Through proteomic analysis of flagellar membranes, Saada et al. identified T. brucei ACs that are localized to specific flagellum subdomains and upregulated in the procyclic life cycle stage [33], making them good candidates to test for a role in SoMo. Utilizing gene-specific RNAi, Lopez et al. found that knockdown of certain procyclic-specific ACs enhanced social motility, while knockdown of others had no effect [34]. The findings reveal that individual AC isoforms have specialized functions and indicate that a subset of ACs coordinates the response to signals governing SoMo. Importantly, AC point mutants that disrupt catalytic activity phenocopy RNAi knockdown, supporting the idea that attenuating cAMP production, as opposed to total loss of the protein, activates SoMo.

In complementary studies, Oberholzer and colleagues found that pharmacological or genetic inhibition of cAMP-specific phosphodiesterase B1 (PDEB1) blocks SoMo, indicating that elevated cAMP levels are inhibitory [35]. Supporting this, cAMP levels in live trypanosomes were monitored using a FRET sensor, demonstrating that SoMo inhibition directly correlates with a rise in intracellular cAMP. Additionally, membrane-permeant cAMP or non-hydrolyzable cAMP analogues were found to inhibit SoMo. Surprisingly, the authors also found that the SoMo defect of PDEB1 knockdown cells is rescued when these mutants are co-cultured with wild type (WT) cells. To do this, they employed WT and PDEB1 knockdown cell lines, each expressing a different fluorescent protein, making it possible to independently monitor movement of WT and PDEB1 knockdown cells in a mixed culture. Despite being incapable of forming projections on their own, PDEB1 knockdown cells formed projections when mixed with WT cells. The result is the first demonstration of trans-complementation in trypanosomes and suggests that PDEB1 knockdown cells are competent for SoMo, but fail to engage in SoMo because they lack a factor that can be provided by WT cells.

Notably, PDEB1 localizes exclusively to the flagellum [36], as do all ACs analyzed thus far, with at least one AC involved in SoMo that is further restricted to the flagellum tip [33,34,37]. The combined data thus reveal that the T. brucei flagellum is specialized for cAMP signaling and suggest a model in which localized changes in cAMP levels within specific flagellum subdomains control SoMo [3335,38]. In this model, PDEB1 acts as a diffusion barrier along the length of the flagellum to confine cAMP signaling output to the site of its production by specific ACs. In the absence of PDEB1, cAMP becomes elevated and diffuses throughout the flagellum, inhibiting activities that would otherwise drive SoMo [35]. Interestingly, regulation of SoMo by cAMP in T. brucei has parallels to cyclic-nucleotide regulation of swarming motility in bacterial pathogens. In Pseudomonas spp., for example, loss of a diguanylate cyclase that produces cyclic-di-GMP results in hyper-swarming behavior, whereas loss of the cyclic-di-GMP phosphodiesterase results in the inability to swarm at all [39,40].

Trypanosomal ACs have attracted interest because their domain structure suggests they may act as receptors for extracellular ligands [41]. However, the unusually large size of the AC gene family in T. brucei (>65 genes) and lack of assays for cAMP responses have hindered studies of these intriguing proteins. Likewise, although PDEs have been a focus of attention as a drug target in bloodstream-stage T. brucei, their function in fly life cycle stages has received little attention. Indeed, although important advances have been made recently [34,35,4244], cAMP signaling in general remains poorly understood in trypanosomes. In this context, SoMo signaling studies have broad impact because, in addition to identifying genes governing social behavior, they advance understanding of cAMP signaling systems in trypanosomes and demonstrate that the concept of the flagellum as a signaling organelle extends to a group of important human and animal pathogens.

Summary and Outlook

Studies of social motility in African trypanosomes have revealed new conceptual frameworks and produced novel approaches for considering these pathogens. For example, it was not previously possible to discern phenotypic differences between early and late procyclic developmental stages beyond the presence of a single surface protein. SoMo has been instrumental in establishing these as specific life cycle stages and defining a distinct developmental transition [23]. Likewise, although functional analyses of receptor-type adenylate cyclases previously provided insight into host-parasite interaction [44], such studies have been limited to a few members of this large and enigmatic protein family. SoMo now provides ready avenues for structure-function analyses of these understudied proteins [34,45]. More broadly, the ease of visualizing individual trypanosomes in a mixed community and the genetic tractability of T. brucei mean that SoMo can be exploited for elucidating principles of self-assembly and cell–cell interactions in microbial systems.

With foundations established, the stage is now set to tackle several key questions. An obvious area of interest will be to evaluate the role of SoMo and underlying signaling events during fly infection. With mutants now available, it will be possible to address this directly, and these efforts would be aided by development of methodologies for visualizing parasites in live tsetse flies. With regard to signaling mechanisms, it will be informative to elucidate the relationship between elevated cAMP and the late procyclic developmental stage, both of which inhibit SoMo. At least three models can be envisioned to explain these results (Fig 2A). For example, cAMP and the late procyclic developmental stage might act independently. Alternatively, elevated cAMP at the flagellum tip might promote the early to late developmental transition, which then inhibits SoMo. However, the presence of GPEET procyclin, a marker of early procyclics, in PDEB1 knockdowns [35] argues against this model. Finally, development into late procyclics might promote elevation of cAMP at the flagellum tip, which would be the inhibitory trigger. Further studies are needed to distinguish between these models. Additional important topics include investigation of cellular signaling within early versus late-stage procyclic cells, determining the role of AC extracellular domains and location, as well as identification of AC ligands.

thumbnail
Fig 2. Regulation of social motility.

(A) Three alternate models for regulation of social motility. (Left) Elevated cAMP at the flagellum tip (red) in response to regulation of tip-localized adenylate cyclase and the transition from early to late procyclics independently inhibit SoMo. (Middle) Elevated cAMP at the flagellum tip triggers the transition from early to late procyclics, which then inhibits SoMo. (Right) Development into late procyclics triggers elevated cAMP at the flagellum tip, which is then the inhibitory signal. (B) In addition to the known cAMP signaling systems that control SoMo (red), the trypanosome flagellum harbors several predicted signaling systems, e.g., ion transporters, kinases, additional ACs, and other receptor-like proteins [50,51] whose functions await discovery.

https://doi.org/10.1371/journal.ppat.1005272.g002

Beyond cAMP, SoMo undoubtedly depends on additional cell-derived as well as non-cell-derived elements. Of particular interest are factors that drive assembly of parasites into groups, those that promote outward movement of these groups, and those that govern repulsion of groups from one another. Also important will be to define factors responsible for trans-complementation. In bacteria, trans-complementation can be achieved through transfer of outer membrane proteins from one cell to another, and multiple signaling systems have been implicated in coordinating swarming [4649]. The trypanosome flagellum houses several signaling systems with the capacity for metabotropic and ionotropic responses to external signals—including ion transporters, kinases, ACs, and other putative receptors (Fig 2B) [50,51]—whose functions await discovery.

In closing, studies of social motility and quorum sensing have provided insight into trypanosome developmental biology and signal transduction, illustrating the value of considering these parasites in the context of microbial social behavior concepts. It will be informative to apply this paradigm more broadly among protozoan pathogens.

Acknowledgments

We thank members of the Hill laboratory as well as Isabel Roditi for many stimulating discussions about social motility.

References

  1. 1. Shapiro JA. 1998. Thinking about bacterial populations as multicellular organisms. Annu Rev Microbiol 52:81–104. pmid:9891794
  2. 2. Bassler BL, Losick R. 2006. Bacterially speaking. Cell 125:237–246. pmid:16630813
  3. 3. Kaiser D. 2003. Coupling cell movement to multicellular development in myxobacteria. Nat Rev Microbiol 1:45–54. pmid:15040179
  4. 4. Shaulsky G, Kessin RH. 2007. The cold war of the social amoebae. Current Biology 17:R684–692. pmid:17714662
  5. 5. Harshey RM. 2003. Bacterial motility on a surface: many ways to a common goal. Annu Rev Microbiol 57:249–273. pmid:14527279
  6. 6. Reuner B, Vassella E, Yutzy B, Boshart M. 1997. Cell density triggers slender to stumpy differentiation of Trypanosoma brucei bloodstream forms in culture. Mol Biochem Parasitol 90:269–280. pmid:9497048
  7. 7. Mony BM, MacGregor P, Ivens A, Rojas F, Cowton A, Young J, Horn D, Matthews K. 2014. Genome-wide dissection of the quorum sensing signalling pathway in Trypanosoma brucei. Nature 505:681–685. pmid:24336212
  8. 8. Oberholzer M, Lopez MA, McLelland BT, Hill KL. 2010. Social motility in african trypanosomes. PLoS Pathog 6:e1000739. pmid:20126443
  9. 9. Cecchi G, Mattioli RC, Slingenbergh J, de la Rocque S. 2008. Land cover and tsetse fly distributions in sub-Saharan Africa. Med Vet Entomol 22:364–373. pmid:18785934
  10. 10. Simarro PP, Cecchi G, Paone M, Franco JR, Diarra A, Ruiz JA, Fevre EM, Courtin F, Mattioli RC, Jannin JG. 2010. The Atlas of human African trypanosomiasis: a contribution to global mapping of neglected tropical diseases. Int J Health Geogr 9:57. pmid:21040555
  11. 11. WHO. 2012. Mapping the risk of Human African Trypanosomiasis. http://www.who.int/trypanosomiasis_african/country/risk_AFRO/en/
  12. 12. Vickerman K, Tetley L, Hendry KA, Turner CM. 1988. Biology of African trypanosomes in the tsetse fly. Biol Cell 64:109–119. pmid:3067793
  13. 13. Fraser GM, Hughes C. 1999. Swarming motility. Curr Opin Microbiol 2:630–635. pmid:10607626
  14. 14. Butler MT, Wang Q, Harshey RM. 2010. Cell density and mobility protect swarming bacteria against antibiotics. Proc Natl Acad Sci U S A 107:3776–3781. pmid:20133590
  15. 15. Roditi I, Lehane MJ. 2008. Interactions between trypanosomes and tsetse flies. Curr Opin Microbiol 11:345–351. pmid:18621142
  16. 16. Pan J, Snell WJ. 2000. Signal transduction during fertilization in the unicellular green alga, Chlamydomonas. Curr Opin Microbiol 3:596–602. pmid:11121779
  17. 17. Peacock L, Ferris V, Bailey M, Gibson W. 2012. The influence of sex and fly species on the development of trypanosomes in tsetse flies. PLoS Negl Trop Dis 6:e1515. pmid:22348165
  18. 18. Haines LR. 2013. Examining the tsetse teneral phenomenon and permissiveness to trypanosome infection. Front Cell Infect Microbiol 3:84. pmid:24312903
  19. 19. Gibson W, Bailey M. 2003. The development of Trypanosoma brucei within the tsetse fly midgut observed using green fluorescent trypanosomes. Kinetoplastid Biol Dis 2:1. pmid:12769824
  20. 20. Lehane MJ. 1997. Peritrophic matrix structure and function. Annu Rev Entomol 42:525–550. pmid:15012322
  21. 21. Dyer NA, Rose C, Ejeh NO, Acosta-Serrano A. 2013. Flying tryps: survival and maturation of trypanosomes in tsetse flies. Trends Parasitol 29:188–196. pmid:23507033
  22. 22. Oberle M, Balmer O, Brun R, Roditi I. 2010. Bottlenecks and the maintenance of minor genotypes during the life cycle of Trypanosoma brucei. PLoS Pathog 6:e1001023. pmid:20686656
  23. 23. Imhof S, Knusel S, Gunasekera K, Vu XL, Roditi I. 2014. Social motility of African trypanosomes is a property of a distinct life-cycle stage that occurs early in tsetse fly transmission. PLoS Pathog 10:e1004493. pmid:25357194
  24. 24. Imhof S, Vu XL, Butikofer P, Roditi I. 2015. A glycosylation mutant of Trypanosoma brucei links social motility defects in vitro to impaired colonisation of tsetse in vivo. Eukaryot Cell 14:588–92. pmid:25862152
  25. 25. Vassella E, Den Abbeele JV, Butikofer P, Renggli CK, Furger A, Brun R, Roditi I. 2000. A major surface glycoprotein of trypanosoma brucei is expressed transiently during development and can be regulated post-transcriptionally by glycerol or hypoxia. Genes Dev 14:615–626. pmid:10716949
  26. 26. Jelk J, Gao N, Serricchio M, Signorell A, Schmidt RS, Bangs JD, Acosta-Serrano A, Lehrman MA, Butikofer P, Menon AK. 2013. Glycoprotein biosynthesis in a eukaryote lacking the membrane protein Rft1. J Biol Chem 288:20616–20623. pmid:23720757
  27. 27. Matthews KR. 2011. Controlling and coordinating development in vector-transmitted parasites. Science 331:1149–1153. pmid:21385707
  28. 28. Singla V, Reiter JF. 2006. The primary cilium as the cell's antenna: signaling at a sensory organelle. Science 313:629–633. pmid:16888132
  29. 29. Johnson JL, Leroux MR. 2010. cAMP and cGMP signaling: sensory systems with prokaryotic roots adopted by eukaryotic cilia. Trends Cell Biol 20:435–444. pmid:20541938
  30. 30. Ferreira RB, Chodur DM, Antunes LC, Trimble MJ, McCarter LL. 2012. Output targets and transcriptional regulation by a cyclic dimeric GMP-responsive circuit in the Vibrio parahaemolyticus Scr network. J Bacteriol 194:914–924. pmid:22194449
  31. 31. Meili R, Firtel RA. 2003. Follow the leader. Dev Cell 4:291–293. pmid:12636911
  32. 32. Merritt JH, Ha DG, Cowles KN, Lu W, Morales DK, Rabinowitz J, Gitai Z, O'Toole GA. 2010. Specific control of Pseudomonas aeruginosa surface-associated behaviors by two c-di-GMP diguanylate cyclases. MBio 1: e00183–10. pmid:20978535
  33. 33. Bloodgood RA. 2012. The future of ciliary and flagellar membrane research. Mol Biol Cell 23:2407–2411. pmid:22745341
  34. 34. Lopez MA, Saada EA, Hill KL. 2015. Insect stage-specific adenylate cyclases regulate social motility in African trypanosomes. Eukaryot Cell 14:104–112. pmid:25416239
  35. 35. Oberholzer M, Saada EA, Hill KL. 2015. Cyclic AMP Regulates Social Behavior in African Trypanosomes. MBio 6:e01954–14. pmid:25922395
  36. 36. Oberholzer M, Marti G, Baresic M, Kunz S, Hemphill A, Seebeck T. 2007. The Trypanosoma brucei cAMP phosphodiesterases TbrPDEB1 and TbrPDEB2: flagellar enzymes that are essential for parasite virulence. Faseb J 21:720–731. pmid:17167070
  37. 37. Paindavoine P, Rolin S, Van Assel S, Geuskens M, Jauniaux JC, Dinsart C, Huet G, Pays E. 1992. A gene from the variant surface glycoprotein expression site encodes one of several transmembrane adenylate cyclases located on the flagellum of Trypanosoma brucei. Mol Cell Biol 12:1218–1225. pmid:1545803
  38. 38. Oberholzer M, Bregy P, Marti G, Minca M, Peier M, Seebeck T. 2007. Trypanosomes and mammalian sperm: one of a kind? Trends Parasitol 23:71–77. pmid:17174157
  39. 39. Merritt JH, Brothers KM, Kuchma SL, O'Toole GA. 2007. SadC reciprocally influences biofilm formation and swarming motility via modulation of exopolysaccharide production and flagellar function. J Bacteriol 189:8154–8164. pmid:17586642
  40. 40. Kuchma SL, Brothers KM, Merritt JH, Liberati NT, Ausubel FM, O'Toole GA. 2007. BifA, a cyclic-Di-GMP phosphodiesterase, inversely regulates biofilm formation and swarming motility by Pseudomonas aeruginosa PA14. J Bacteriol 189:8165–8178. pmid:17586641
  41. 41. Pays E, Nolan DP. 1998. Expression and function of surface proteins in Trypanosoma brucei. Mol Biochem Parasitol 91:3–36. pmid:9574923
  42. 42. Gould MK, Bachmaier S, Ali JA, Alsford S, Tagoe DN, Munday JC, Schnaufer AC, Horn D, Boshart M, de Koning HP. 2013. Cyclic AMP Effectors in African Trypanosomes Revealed by Genome-Scale RNA Interference Library Screening for Resistance to the Phosphodiesterase Inhibitor CpdA. Antimicrob Agents Chemother 57:4882–4893. pmid:23877697
  43. 43. de Koning HP, Gould MK, Sterk GJ, Tenor H, Kunz S, Luginbuehl E, Seebeck T. 2012. Pharmacological validation of Trypanosoma brucei phosphodiesterases as novel drug targets. J Infect Dis 206:229–237. pmid:22291195
  44. 44. Salmon D, Vanwalleghem G, Morias Y, Denoeud J, Krumbholz C, Lhomme F, Bachmaier S, Kador M, Gossmann J, Dias FB, De Muylder G, Uzureau P, Magez S, Moser M, De Baetselier P, Van Den Abbeele J, Beschin A, Boshart M, Pays E. 2012. Adenylate cyclases of Trypanosoma brucei inhibit the innate immune response of the host. Science 337:463–466. pmid:22700656
  45. 45. Saada EA, Kabututu ZP, Lopez M, Shimogawa MM, Langousis G, Oberholzer M, Riestra A, Jonsson ZO, Wohlschlegel JA, Hill KL. 2014. Insect stage-specific receptor adenylate cyclases are localized to distinct subdomains of the Trypanosoma brucei Flagellar membrane. Eukaryot Cell 13:1064–1076. pmid:24879126
  46. 46. Nickzad A, Lepine F, Deziel E. 2015. Quorum Sensing Controls Swarming Motility of Burkholderia glumae through Regulation of Rhamnolipids. PLoS One 10:e0128509. pmid:26047513
  47. 47. Shrout JD, Tolker-Nielsen T, Givskov M, Parsek MR. 2011. The contribution of cell-cell signaling and motility to bacterial biofilm formation. MRS Bull 36:367–373. pmid:22053126
  48. 48. Pathak DT, Wei X, Wall D. 2012. Myxobacterial tools for social interactions. Res Microbiol 163:579–591. pmid:23123306
  49. 49. Pathak DT, Wei X, Bucuvalas A, Haft DH, Gerloff DL, Wall D. 2012. Cell contact-dependent outer membrane exchange in myxobacteria: genetic determinants and mechanism. PLoS Genet 8:e1002626. pmid:22511878
  50. 50. Oberholzer M, Langousis G, Nguyen HT, Saada EA, Shimogawa MM, Jonsson ZO, Nguyen SM, Wohlschlegel JA, Hill KL. 2011. Independent analysis of the flagellum surface and matrix proteomes provides insight into flagellum signaling in mammalian-infectious Trypanosoma brucei. Mol Cell Proteomics 10:M111 010538.
  51. 51. Subota I, Julkowska D, Vincensini L, Reeg N, Buisson J, Blisnick T, Huet D, Perrot S, Santi-Rocca J, Duchateau M, Hourdel V, Rousselle JC, Cayet N, Namane A, Chamot-Rooke J, Bastin P. 2014. Proteomic analysis of intact flagella of procyclic Trypanosoma brucei cells identifies novel flagellar proteins with unique sub-localisation and dynamics. Mol Cell Proteomics.
  52. 52. Turner CM, Barry JD, Vickerman K. 1988. Loss of variable antigen during transformation of Trypanosoma brucei rhodesiense from bloodstream to procyclic forms in the tsetse fly. Parasitol Res 74:507–511. pmid:3194363
  53. 53. MacGregor P, Szoor B, Savill NJ, Matthews KR. 2012. Trypanosomal immune evasion, chronicity and transmission: an elegant balancing act. Nat Rev Microbiol 10:431–438. pmid:22543519