1887

Abstract

-glycolylneuraminic acid (Neu5Gc), and its precursor -acetylneuraminic acid (Neu5Ac), commonly referred to as sialic acids, are two of the most common glycans found in mammals. Humans carry a mutation in the enzyme that converts Neu5Ac into Neu5Gc, and as such, expression of Neu5Ac can be thought of as a ‘human specific’ trait. Bacteria can utilize sialic acids as a carbon and energy source and have evolved multiple ways to take up sialic acids. In order to generate free sialic acid, many bacteria produce sialidases that cleave sialic acid residues from complex glycan structures. In addition, sialidases allow escape from innate immune mechanisms, and can synergize with other virulence factors such as toxins. Human-adapted pathogens have evolved a preference for Neu5Ac, with many bacterial adhesins, and major classes of toxin, specifically recognizing Neu5Ac containing glycans as receptors. The preference of human-adapted pathogens for Neu5Ac also occurs during biosynthesis of surface structures such as lipo-oligosaccharide (LOS), lipo-polysaccharide (LPS) and polysaccharide capsules, subverting the human host immune system by mimicking the host. This review aims to provide an update on the advances made in understanding the role of sialic acid in bacteria-host interactions made in the last 5–10 years, and put these findings into context by highlighting key historical discoveries. We provide a particular focus on ‘molecular mimicry’ and incorporation of sialic acid onto the bacterial outer-surface, and the role of sialic acid as a receptor for bacterial adhesins and toxins.

  • This is an open-access article distributed under the terms of the Creative Commons Attribution NonCommercial License. This article was made open access via a Publish and Read agreement between the Microbiology Society and the corresponding author’s institution.
Loading

Article metrics loading...

/content/journal/micro/10.1099/mic.0.001157
2022-03-22
2024-04-26
Loading full text...

Full text loading...

/deliver/fulltext/micro/168/3/mic001157.html?itemId=/content/journal/micro/10.1099/mic.0.001157&mimeType=html&fmt=ahah

References

  1. Angata T, Varki A. Chemical diversity in the sialic acids and related alpha-keto acids: an evolutionary perspective. Chem Rev 2002; 102:439–469 [View Article] [PubMed]
    [Google Scholar]
  2. McDonald ND, Boyd EF. Structural and biosynthetic diversity of nonulosonic acids (NulOs) that decorate surface structures in bacteria. Trends Microbiol 2021; 29:142–157 [View Article] [PubMed]
    [Google Scholar]
  3. Irie A, Suzuki A. CMP-N-Acetylneuraminic acid hydroxylase is exclusively inactive in humans. Biochem Biophys Res Commun 1998; 248:330–333 [View Article] [PubMed]
    [Google Scholar]
  4. Irie A, Koyama S, Kozutsumi Y, Kawasaki T, Suzuki A et al. The molecular basis for the absence of N-glycolylneuraminic acid in humans. J Biol Chem 1998; 273:15866–15871 [View Article] [PubMed]
    [Google Scholar]
  5. Muchmore EA, Diaz S, Varki A et al. A structural difference between the cell surfaces of humans and the great apes. Am J Phys Anthropol 1998; 107:187–198 [View Article] [PubMed]
    [Google Scholar]
  6. Chou HH, Takematsu H, Diaz S, Iber J, Nickerson E et al. A mutation in human CMP-sialic acid hydroxylase occurred after the Homo-Pan divergence. Proc Natl Acad Sci U S A 1998; 95:11751–11756 [View Article] [PubMed]
    [Google Scholar]
  7. Pham T, Gregg CJ, Karp F, Chow R, Padler-Karavani V et al. Evidence for a novel human-specific xeno-auto-antibody response against vascular endothelium. Blood 2009; 114:5225–5235 [View Article] [PubMed]
    [Google Scholar]
  8. Hedlund M, Padler-Karavani V, Varki NM, Varki A et al. Evidence for a human-specific mechanism for diet and antibody-mediated inflammation in carcinoma progression. Proc Natl Acad Sci U S A 2008; 105:18936–18941 [View Article] [PubMed]
    [Google Scholar]
  9. Diaz SL, Padler-Karavani V, Ghaderi D, Hurtado-Ziola N, Yu H et al. Sensitive and specific detection of the non-human sialic Acid N-glycolylneuraminic acid in human tissues and biotherapeutic products. PLoS One 2009; 4:e4241 [View Article] [PubMed]
    [Google Scholar]
  10. Tangvoranuntakul P, Gagneux P, Diaz S, Bardor M, Varki N et al. Human uptake and incorporation of an immunogenic nonhuman dietary sialic acid. Proc Natl Acad Sci U S A 2003; 100:12045–12050 [View Article] [PubMed]
    [Google Scholar]
  11. Bergfeld AK, Pearce OMT, Diaz SL, Pham T, Varki A et al. Metabolism of vertebrate amino sugars with N-glycolyl groups: elucidating the intracellular fate of the non-human sialic acid N-glycolylneuraminic acid. J Biol Chem 2012; 287:28865–28881 [View Article] [PubMed]
    [Google Scholar]
  12. Bousquet PA, Sandvik JA, Jeppesen Edin NF, Krengel U et al. Hypothesis: Hypoxia induces de novo synthesis of NeuGc gangliosides in humans through CMAH domain substitute. Biochem Biophys Res Commun 2018; 495:1562–1566 [View Article] [PubMed]
    [Google Scholar]
  13. Blaum BS, Hannan JP, Herbert AP, Kavanagh D, Uhrín D et al. Structural basis for sialic acid-mediated self-recognition by complement factor H. Nat Chem Biol 2015; 11:77–82 [View Article] [PubMed]
    [Google Scholar]
  14. Lewis AL, Lewis WG. Host sialoglycans and bacterial sialidases: a mucosal perspective. Cell Microbiol 2012; 14:1174–1182 [View Article] [PubMed]
    [Google Scholar]
  15. Barry GT, Goebel WF. Colominic acid, a substance of bacterial origin related to sialic acid. Nature 1957; 179:206 [View Article] [PubMed]
    [Google Scholar]
  16. Rohr TE, Troy FA. Structure and biosynthesis of surface polymers containing polysialic acid in Escherichia coli . J Biol Chem 1980; 255:2332–2342 [PubMed]
    [Google Scholar]
  17. Troy FA, McCloskey MA. Role of a membranous sialyltransferase complex in the synthesis of surface polymers containing polysialic acid in Escherichia coli. Temperature-induced alteration in the assembly process. J Biol Chem 1979; 254:7377–7387 [PubMed]
    [Google Scholar]
  18. Linton D, Karlyshev AV, Hitchen PG, Morris HR, Dell A et al. Multiple N-acetyl neuraminic acid synthetase (neuB) genes in Campylobacter jejuni: identification and characterization of the gene involved in sialylation of lipo-oligosaccharide. Mol Microbiol 2000; 35:1120–1134 [View Article] [PubMed]
    [Google Scholar]
  19. Gunawan J, Simard D, Gilbert M, Lovering AL, Wakarchuk WW et al. Structural and mechanistic analysis of sialic acid synthase NeuB from Neisseria meningitidis in complex with Mn2+, phosphoenolpyruvate, and N-acetylmannosaminitol. J Biol Chem 2005; 280:3555–3563 [View Article] [PubMed]
    [Google Scholar]
  20. Wessels MR, Pozsgay V, Kasper DL, Jennings HJ et al. Structure and immunochemistry of an oligosaccharide repeating unit of the capsular polysaccharide of type III group B Streptococcus. A revised structure for the type III group B streptococcal polysaccharide antigen. J Biol Chem 1987; 262:8262–8267 [View Article] [PubMed]
    [Google Scholar]
  21. Lewis AL, Robinson LS, Agarwal K, Lewis WG et al. Discovery and characterization of de novo sialic acid biosynthesis in the phylum Fusobacterium. Glycobiology 2016; 26:1107–1119 [View Article] [PubMed]
    [Google Scholar]
  22. Zapata G, Crowley JM, Vann WF et al. Sequence and expression of the Escherichia coli K1 neuC gene product. J Bacteriol 1992; 174:315–319 [View Article] [PubMed]
    [Google Scholar]
  23. Vimr ER, Aaronson W, Silver RP et al. Genetic analysis of chromosomal mutations in the polysialic acid gene cluster of Escherichia coli K1. J Bacteriol 1989; 171:1106–1117 [View Article] [PubMed]
    [Google Scholar]
  24. Silver RP, Vann WF, Aaronson W et al. Genetic and molecular analyses of Escherichia coli K1 antigen genes. J Bacteriol 1984; 157:568–575 [View Article] [PubMed]
    [Google Scholar]
  25. Vann WF, Tavarez JJ, Crowley J, Vimr E, Silver RP et al. Purification and characterization of the Escherichia coli K1 neuB gene product N-acetylneuraminic acid synthetase. Glycobiology 1997; 7:697–701 [View Article] [PubMed]
    [Google Scholar]
  26. Knirel YA, Rietschel ET, Marre R, Zähringer U et al. The structure of the O-specific chain of Legionella pneumophila serogroup 1 lipopolysaccharide. Eur J Biochem 1994; 221:239–245 [View Article] [PubMed]
    [Google Scholar]
  27. Kenyon JJ, Marzaioli AM, De Castro C, Hall RM et al. 5,7-di-N-acetyl-acinetaminic acid: A novel non-2-ulosonic acid found in the capsule of an Acinetobacter baumannii isolate. Glycobiology 2015; 25:644–654 [View Article] [PubMed]
    [Google Scholar]
  28. Morens DM, Taubenberger JK, Fauci AS et al. Predominant role of bacterial pneumonia as a cause of death in pandemic influenza: implications for pandemic influenza preparedness. J Infect Dis 2008; 198:962–970 [View Article] [PubMed]
    [Google Scholar]
  29. Siegel SJ, Roche AM, Weiser JN et al. Influenza promotes pneumococcal growth during coinfection by providing host sialylated substrates as a nutrient source. Cell Host Microbe 2014; 16:55–67 [View Article] [PubMed]
    [Google Scholar]
  30. Blanchette KA, Shenoy AT, Milner J 2nd, Gilley RP, McClure E et al. Neuraminidase a-exposed galactose promotes Streptococcus pneumoniae biofilm formation during colonization. Infect Immun 2016; 84:2922–2932 [View Article] [PubMed]
    [Google Scholar]
  31. Coats MT, Murphy T, Paton JC, Gray B, Briles DE et al. Exposure of Thomsen-Friedenreich antigen in Streptococcus pneumoniae infection is dependent on pneumococcal neuraminidase A. Microbial Pathogenesis 2011; 50:343–349 [View Article] [PubMed]
    [Google Scholar]
  32. Kullaya V, de Jonge MI, Langereis JD, van der Gaast-de Jongh CE, Büll C et al. Desialylation of platelets by pneumococcal neuraminidase A induces ADP-dependent platelet hyperreactivity. Infect Immun 2018; 86:10 [View Article] [PubMed]
    [Google Scholar]
  33. Gratz N, Loh LN, Mann B, Gao G, Carter R et al. Pneumococcal neuraminidase activates TGF-β signalling. Microbiology (Reading) 2017; 163:1198–1207 [View Article]
    [Google Scholar]
  34. Song Y, Pan Q, Xiao J, Li W, Ma H et al. Sialidase of Glaesserella parasuis augments inflammatory response via desialylation and abrogation of negative regulation of siglec-5. Infect Immun 2021; 89: [View Article] [PubMed]
    [Google Scholar]
  35. Hammond AJ, Binsker U, Aggarwal SD, Ortigoza MB, Loomis C et al. Neuraminidase B controls neuraminidase A-dependent mucus production and evasion. PLoS Pathog 2021; 17:e1009158 [View Article] [PubMed]
    [Google Scholar]
  36. King CA, Van Heyningen WE. Deactivation of cholera toxin by a sialidase-resistant monosialosylganglioside. J Infect Dis 1973; 127:639–647 [View Article] [PubMed]
    [Google Scholar]
  37. Holmgren J, Lönnroth I, Månsson J, Svennerholm L et al. Interaction of cholera toxin and membrane GM1 ganglioside of small intestine. Proc Natl Acad Sci U S A 1975; 72:2520–2524 [View Article] [PubMed]
    [Google Scholar]
  38. Robinson LS, Schwebke J, Lewis WG, Lewis AL et al. Identification and characterization of NanH2 and NanH3, enzymes responsible for sialidase activity in the vaginal bacterium Gardnerella vaginalis . J Biol Chem 2019; 294:5230–5245 [View Article] [PubMed]
    [Google Scholar]
  39. Hardy L, Jespers V, Van den Bulck M, Buyze J, Mwambarangwe L et al. The presence of the putative Gardnerella vaginalis sialidase A gene in vaginal specimens is associated with bacterial vaginosis biofilm. PLoS One 2017; 12:e0172522 [View Article] [PubMed]
    [Google Scholar]
  40. Govinden G, Parker JL, Naylor KL, Frey AM, Anumba DOC et al. Inhibition of sialidase activity and cellular invasion by the bacterial vaginosis pathogen Gardnerella vaginalis . Arch Microbiol 2018; 200:1129–1133 [View Article] [PubMed]
    [Google Scholar]
  41. Frey AM, Satur MJ, Phansopa C, Honma K, Urbanowicz PA et al. Characterization of Porphyromonas gingivalis sialidase and disruption of its role in host–pathogen interactions. Microbiology 2019; 165:1181–1197 [View Article] [PubMed]
    [Google Scholar]
  42. Navarro MA, Li J, McClane BA, Morrell E, Beingesser J et al. NanI sialidase is an important contributor to Clostridium perfringens type F strain F4969 intestinal colonization in mice. Infect Immun 2018; 86:12 [View Article] [PubMed]
    [Google Scholar]
  43. Li J, McClane BA, Young VB. NanI sialidase can support the growth and survival of Clostridium perfringens strain F4969 in the presence of sialyated host macromolecules (Mucin) or Caco-2 cells. Infect Immun 2018; 86: [View Article] [PubMed]
    [Google Scholar]
  44. MacMillan JL, Vicaretti SD, Noyovitz B, Xing X, Low KE et al. Structural analysis of broiler chicken small intestinal mucin O-glycan modification by Clostridium perfringens . Poult Sci 2019; 98:5074–5088 [View Article]
    [Google Scholar]
  45. Michaels DL, Moneypenny CG, Shama SM, Leibowitz JA, May MA et al. Sialidase and N-acetylneuraminate catabolism in nutrition of Mycoplasma alligatoris . Microbiology 2019; 165:662–667 [View Article] [PubMed]
    [Google Scholar]
  46. May M, Brown DR. Genetic variation in sialidase and linkage to N-acetylneuraminate catabolism in Mycoplasma synoviae . Microb Pathog 2008; 45:38–44 [View Article] [PubMed]
    [Google Scholar]
  47. Thomas GH. Sialic acid acquisition in bacteria-one substrate, many transporters. Biochem Soc Trans 2016; 44:760–765 [View Article] [PubMed]
    [Google Scholar]
  48. Haines-Menges BL, Whitaker WB, Lubin JB, Boyd EF et al. Host sialic acids: a delicacy for the pathogen with discerning taste. Microbiol Spectr 2015; 3: [View Article] [PubMed]
    [Google Scholar]
  49. Martinez J, Steenbergen S, Vimr E et al. Derived structure of the putative sialic acid transporter from Escherichia coli predicts a novel sugar permease domain. J Bacteriol 1995; 177:6005–6010 [View Article] [PubMed]
    [Google Scholar]
  50. Walters DM, Stirewalt VL, Melville SB et al. Cloning, sequence, and transcriptional regulation of the operon encoding a putative N-acetylmannosamine-6-phosphate epimerase (nanE) and sialic acid lyase (nanA) in Clostridium perfringens . J Bacteriol 1999; 181:4526–4532 [View Article] [PubMed]
    [Google Scholar]
  51. Olson ME, King JM, Yahr TL, Horswill AR. Sialic acid catabolism in Staphylococcus aureus . J Bacteriol 2013; 195:1779–1788 [View Article] [PubMed]
    [Google Scholar]
  52. Rosa LT, Bianconi ME, Thomas GH, Kelly DJ et al. Tripartite ATP-Independent Periplasmic (TRAP) Transporters and Tripartite Tricarboxylate Transporters (TTT): from uptake to pathogenicity. Front Cell Infect Microbiol 2018; 8:33 [View Article] [PubMed]
    [Google Scholar]
  53. Severi E, Randle G, Kivlin P, Whitfield K, Young R et al. Sialic acid transport in Haemophilus influenzae is essential for lipopolysaccharide sialylation and serum resistance and is dependent on a novel tripartite ATP-independent periplasmic transporter. Mol Microbiol 2005; 58:1173–1185 [View Article] [PubMed]
    [Google Scholar]
  54. Allen S, Zaleski A, Johnston JW, Gibson BW, Apicella MA et al. Novel sialic acid transporter of Haemophilus influenzae . Infect Immun 2005; 73:5291–5300 [View Article] [PubMed]
    [Google Scholar]
  55. Johnston JW, Zaleski A, Allen S, Mootz JM, Armbruster D et al. Regulation of sialic acid transport and catabolism in Haemophilus influenzae . Mol Microbiol 2007; 66:26–39 [View Article] [PubMed]
    [Google Scholar]
  56. Jenkins GA, Figueira M, Kumar GA, Sweetman WA, Makepeace K et al. Sialic acid mediated transcriptional modulation of a highly conserved sialometabolism gene cluster in Haemophilus influenzae and its effect on virulence. BMC Microbiol 2010; 10:48 [View Article] [PubMed]
    [Google Scholar]
  57. Gangi Setty T, Cho C, Govindappa S, Apicella MA, Ramaswamy S et al. Bacterial periplasmic sialic acid-binding proteins exhibit a conserved binding site. Acta Crystallogr D Biol Crystallogr 2014; 70:1801–1811 [View Article] [PubMed]
    [Google Scholar]
  58. Tatum FM, Tabatabai LB, Briggs RE. Sialic acid uptake is necessary for virulence of Pasteurella multocida in turkeys. Microb Pathog 2009; 46:337–344 [View Article] [PubMed]
    [Google Scholar]
  59. North RA, Horne CR, Davies JS, Remus DM, Muscroft-Taylor AC et al. “Just a spoonful of sugar”: import of sialic acid across bacterial cell membranes. Biophys Rev 2018; 10:219–227 [View Article] [PubMed]
    [Google Scholar]
  60. Vimr ER, Kalivoda KA, Deszo EL, Steenbergen SM et al. Diversity of microbial sialic acid metabolism. Microbiol Mol Biol Rev 2004; 68:132–153 [View Article] [PubMed]
    [Google Scholar]
  61. Reddi G, Pruss K, Cottingham KL, Taylor RK, Almagro-Moreno S et al. Catabolism of mucus components influences motility of Vibrio cholerae in the presence of environmental reservoirs. PLoS ONE 2018; 13:e0201383 [View Article] [PubMed]
    [Google Scholar]
  62. Crost EH, Tailford LE, Monestier M, Swarbreck D, Henrissat B et al. The mucin-degradation strategy of Ruminococcus gnavus: The importance of intramolecular trans-sialidases. Gut Microbes 2016; 7:302–312 [View Article]
    [Google Scholar]
  63. Uhde A, Brühl N, Goldbeck O, Matano C, Gurow O et al. Transcription of sialic acid catabolism genes in Corynebacterium glutamicum is subject to catabolite repression and control by the transcriptional repressor NanR. J Bacteriol 2016; 198:2204–2218 [View Article] [PubMed]
    [Google Scholar]
  64. Ward ME, Watt PJ, Glynn AA. Gonococci in urethral exudates possess a virulence factor lost on subculture. Nature 1970; 227:382–384 [View Article] [PubMed]
    [Google Scholar]
  65. Nairn CA, Cole JA, Patel PV, Parsons NJ, Fox JE et al. Cytidine 5’-monophospho-N-acetylneuraminic acid or a related compound is the low mr factor from human red blood cells which induces gonococcal resistance to killing by human serum. Microbiology 1988; 134:3295–3306 [View Article] [PubMed]
    [Google Scholar]
  66. Wetzler LM, Barry K, Blake MS, Gotschlich EC. Gonococcal lipooligosaccharide sialylation prevents complement-dependent killing by immune sera. Infect Immun 1992; 60:39–43 [View Article]
    [Google Scholar]
  67. Gulati S, Schoenhofen IC, Whitfield DM, Cox AD, Li J et al. Utilizing CMP-sialic acid analogs to unravel Neisseria gonorrhoeae lipooligosaccharide-mediated complement resistance and design novel therapeutics. PLoS Pathog 2015; 11:e1005290 [View Article] [PubMed]
    [Google Scholar]
  68. Ram S, Gulati S, Lewis LA, Chakraborti S, Zheng B et al. A novel sialylation site on Neisseria gonorrhoeae lipooligosaccharide links heptose II lactose expression with pathogenicity. Infect Immun 2018; 86: [View Article] [PubMed]
    [Google Scholar]
  69. Lewis LA, Gulati S, Burrowes E, Zheng B, Ram S et al. α-2,3-sialyltransferase expression level impacts the kinetics of lipooligosaccharide sialylation, complement resistance, and the ability of Neisseria gonorrhoeae to colonize the murine genital tract. mBio 2015; 6:e02465-14 [View Article]
    [Google Scholar]
  70. Apicella MA, Mandrell RE. Molecular mimicry as a factor in the pathogenesis of human neisserial infections: in vitro and in vivo modification of the lipooligosaccharide of Neisseria gonorrhoeae by N-acetylneuraminic acid. Pediatr Infect Dis J 1989; 8:901–902 [View Article]
    [Google Scholar]
  71. Jackson MD, Wong SM, Akerley BJ et al. Sialic Acid Protects Nontypeable Haemophilus influenzae from Natural IgM and Promotes Survival in Murine Respiratory Tract. Infect Immun 2021; 89:e00676-20 [View Article]
    [Google Scholar]
  72. Oerlemans MMP, Moons SJ, Heming JJA, Boltje TJ, de Jonge MI et al. Uptake of sialic acid by nontypeable Haemophilus influenzae increases complement resistance through decreasing IgM-dependent complement activation. Infect Immun 2019; 87:e00077-19 [View Article]
    [Google Scholar]
  73. Apicella MA, Coffin J, Ketterer M, Post DMB, Day CJ et al. Nontypeable Haemophilus influenzae Lipooligosaccharide Expresses a Terminal Ketodeoxyoctanoate In Vivo, Which Can Be Used as a Target for Bactericidal Antibody. mBio 2018; 9:e01401-18 [View Article]
    [Google Scholar]
  74. Jackson MD, Wong SM, Akerley BJ, Roy CR et al. Underlying glycans determine the ability of sialylated lipooligosaccharide to protect nontypeable Haemophilus influenzae from serum IgM and complement. Infect Immun 2019; 87:11 [View Article]
    [Google Scholar]
  75. Heise T, Langereis JD, Rossing E, de Jonge MI, Adema GJ et al. Selective inhibition of sialic acid-based molecular mimicry in Haemophilus influenzae abrogates serum resistance. Cell Chemical Biology 2018; 25:1279–1285 [View Article]
    [Google Scholar]
  76. Ng PSK, Day CJ, Atack JM, Hartley-Tassell LE, Winter LE et al. Nontypeable Haemophilus influenzae has evolved preferential use of N-acetylneuraminic acid as a host adaptation. mBio 2019; 10:e00422-19 [View Article] [PubMed]
    [Google Scholar]
  77. Lopes GV, Ramires T, Kleinubing NR, Scheik LK, Fiorentini ÂM et al. Virulence factors of foodborne pathogen Campylobacter jejuni . Microb Pathog 2021; 161:105265 [View Article]
    [Google Scholar]
  78. Culebro A, Machado MP, Carriço JA, Rossi M et al. Origin, evolution, and distribution of the molecular machinery for biosynthesis of sialylated lipooligosaccharide structures in Campylobacter coli . Sci Rep 2018; 8:3028 [View Article] [PubMed]
    [Google Scholar]
  79. Silhavy TJ, Kahne D, Walker S. The bacterial cell envelope. Cold Spring Harb Perspect Biol 2010; 2:a000414 [View Article] [PubMed]
    [Google Scholar]
  80. Pier GB. Pseudomonas aeruginosa lipopolysaccharide: a major virulence factor, initiator of inflammation and target for effective immunity. Int J Med Microbiol 2007; 297:277–295 [View Article] [PubMed]
    [Google Scholar]
  81. Slauch JM, Lee AA, Mahan MJ, Mekalanos JJ et al. Molecular characterization of the oafA locus responsible for acetylation of Salmonella typhimurium O-antigen: oafA is a member of a family of integral membrane trans-acylases. J Bacteriol 1996; 178:5904–5909 [View Article] [PubMed]
    [Google Scholar]
  82. Lewis AL, Lubin J-B, Argade S, Naidu N, Choudhury B et al. Genomic and metabolic profiling of nonulosonic acids in Vibrionaceae reveal biochemical phenotypes of allelic divergence in Vibrio vulnificus. Appl Environ Microbiol 2011; 77:5782–5793 [View Article] [PubMed]
    [Google Scholar]
  83. Lubin J-B, Lewis WG, Gilbert NM, Weimer CM, Almagro-Moreno S et al. Host-like carbohydrates promote bloodstream survival of Vibrio vulnificus in vivo . Infect Immun 2015; 83:3126–3136 [View Article] [PubMed]
    [Google Scholar]
  84. Vinogradov E, St Michael F, Homma K, Sharma A, Cox AD et al. Structure of the LPS O-chain from Fusobacterium nucleatum strain 10953, containing sialic acid. Carbohydr Res 2017; 440–441:38–42 [View Article] [PubMed]
    [Google Scholar]
  85. Zaric SS, Lappin MJ, Fulton CR, Lundy FT, Coulter WA et al. Sialylation of Porphyromonas gingivalis LPS and its effect on bacterial-host interactions. Innate Immun 2017; 23:319–326 [View Article] [PubMed]
    [Google Scholar]
  86. Vogel U, Weinberger A, Frank R, Müller A, Köhl J et al. Complement factor C3 deposition and serum resistance in isogenic capsule and lipooligosaccharide sialic acid mutants of serogroup B Neisseria meningitidis . Infect Immun 1997; 65:4022–4029 [View Article] [PubMed]
    [Google Scholar]
  87. Moxon ER, Kroll JS. The role of bacterial polysaccharide capsules as virulence factors. Curr Top Microbiol Immunol 1990; 150:65–85 [View Article] [PubMed]
    [Google Scholar]
  88. Silver R, Vimr E. The Bacteria, vol. 11. Molecular Basis of Bacterial Pathogenesis New York, NY: Academic Press, Inc; 1990
    [Google Scholar]
  89. Hammerschmidt S, Birkholz C, Zähringer U, Robertson BD, van Putten J et al. Contribution of genes from the capsule gene complex (cps) to lipooligosaccharide biosynthesis and serum resistance in Neisseria meningitidis . Mol Microbiol 1994; 11:885–896 [View Article] [PubMed]
    [Google Scholar]
  90. Segura M. Fisher scientific award lecture - the capsular polysaccharides of Group B Streptococcus and Streptococcus suis differently modulate bacterial interactions with dendritic cells. Can J Microbiol 2012; 58:249–260 [View Article] [PubMed]
    [Google Scholar]
  91. Weiman S, Dahesh S, Carlin AF, Varki A, Nizet V et al. Genetic and biochemical modulation of sialic acid O-acetylation on group B Streptococcus: phenotypic and functional impact. Glycobiology 2009; 19:1204–1213 [View Article] [PubMed]
    [Google Scholar]
  92. Weiman S, Uchiyama S, Lin F-YC, Chaffin D, Varki A et al. O-Acetylation of sialic acid on Group B Streptococcus inhibits neutrophil suppression and virulence. Biochem J 2010; 428:163–168 [View Article] [PubMed]
    [Google Scholar]
  93. Uchiyama S, Sun J, Fukahori K, Ando N, Wu M et al. Dual actions of group B Streptococcus capsular sialic acid provide resistance to platelet-mediated antimicrobial killing. Proc Natl Acad Sci U S A 2019; 116:7465–7470 [View Article] [PubMed]
    [Google Scholar]
  94. Auger J-P, Payen S, Roy D, Dumesnil A, Segura M et al. Interactions of Streptococcus suis serotype 9 with host cells and role of the capsular polysaccharide: Comparison with serotypes 2 and 14. PLoS One 2019; 14:e0223864 [View Article] [PubMed]
    [Google Scholar]
  95. Roy D, Takamatsu D, Okura M, Goyette-Desjardins G, Van Calsteren M-R et al. Capsular sialyltransferase specificity mediates different phenotypes in Streptococcus suis and group B Streptococcus . Front Microbiol 2018; 9:545 [View Article] [PubMed]
    [Google Scholar]
  96. Meng F, Wu NH, Nerlich A, Herrler G, Valentin-Weigand P et al. Dynamic virus-bacterium interactions in a porcine precision-cut lung slice coinfection model: swine influenza virus paves the way for Streptococcus suis infection in a two-step process. Infect Immun 2015; 83:2806–2815 [View Article]
    [Google Scholar]
  97. Tong J, Fu Y, Wu N-H, Rohde M, Meng F et al. Sialic acid-dependent interaction of group B streptococci with influenza virus-infected cells reveals a novel adherence and invasion mechanism. Cell Microbiol 2018; 20: [View Article]
    [Google Scholar]
  98. St Geme JW, Kumar VV, Cutter D, Barenkamp SJ. Prevalence and distribution of the hmw and hia genes and the HMW and Hia adhesins among genetically diverse strains of nontypeable Haemophilus influenzae . Infect Immun 1998; 66:364–368 [View Article]
    [Google Scholar]
  99. Barenkamp SJ, St Geme JW. Identification of a second family of high-molecular-weight adhesion proteins expressed by non-typable Haemophilus influenzae . Mol Microbiol 1996; 19:1215–1223 [View Article]
    [Google Scholar]
  100. St Geme JW. The HMW1 adhesin of nontypeable Haemophilus influenzae recognizes sialylated glycoprotein receptors on cultured human epithelial cells. Infect Immun 1994; 62:3881–3889 [View Article]
    [Google Scholar]
  101. Atack JM, Day CJ, Poole J, Brockman KL, Bakaletz LO et al. The HMW2 adhesin of non-typeable Haemophilus influenzae is a human-adapted lectin that mediates high-affinity binding to 2-6 linked N-acetylneuraminic acid glycans. Biochem Biophys Res Commun 2018; 503:1103–1107 [View Article] [PubMed]
    [Google Scholar]
  102. Atack JM, Day CJ, Poole J, Brockman KL, Timms JRL et al. The Nontypeable Haemophilus influenzae major adhesin hia is a dual-function lectin that binds to human-specific respiratory tract sialic acid glycan receptors. mBio 2020; 11:e02714-20 [View Article] [PubMed]
    [Google Scholar]
  103. Shinya K, Ebina M, Yamada S, Ono M, Kasai N et al. Avian flu: influenza virus receptors in the human airway. Nature 2006; 440:435–436 [View Article] [PubMed]
    [Google Scholar]
  104. Padra M, Benktander J, Padra JT, Andersson A, Brundin B et al. Mucin binding to Moraxella catarrhalis during airway inflammation is dependent on sialic acid. Am J Respir Cell Mol Biol 2021; 65:593–602 [View Article] [PubMed]
    [Google Scholar]
  105. Mahdavi J, Sondén B, Hurtig M, Olfat FO, Forsberg L et al. Helicobacter pylori SabA adhesin in persistent infection and chronic inflammation. Science 2002; 297:573–578 [View Article] [PubMed]
    [Google Scholar]
  106. Benktander J, Barone A, Johansson MM, Teneberg S et al. Helicobacter pylori SabA binding gangliosides of human stomach. Virulence 2018; 9:738–751 [View Article] [PubMed]
    [Google Scholar]
  107. Voland P, Weeks DL, Vaira D, Prinz C, Sachs G et al. Specific identification of three low molecular weight membrane-associated antigens of Helicobacter pylori. Aliment Pharmacol Ther 2002; 16:533–544 [View Article] [PubMed]
    [Google Scholar]
  108. Bennett HJ, Roberts IS. Identification of a new sialic acid-binding protein in Helicobacter pylori. FEMS Immunol Med Microbiol 2005; 44:163–169 [View Article] [PubMed]
    [Google Scholar]
  109. Xia B, Sachdev GP, Cummings RD. Pseudomonas aeruginosa mucoid strain 8830 binds glycans containing the sialyl-Lewis x epitope. Glycoconj J 2007; 24:87–95 [View Article] [PubMed]
    [Google Scholar]
  110. Tram G, Poole J, Adams FG, Jennings MP, Eijkelkamp BA et al. The Acinetobacter baumannii autotransporter adhesin Ata recognizes host glycans as high-affinity receptors. ACS Infect Dis 2021; 7:2352–2361 [View Article] [PubMed]
    [Google Scholar]
  111. Ryan PA, Pancholi V, Fischetti VA et al. Group A streptococci bind to mucin and human pharyngeal cells through sialic acid-containing receptors. Infect Immun 2001; 69:7402–7412 [View Article]
    [Google Scholar]
  112. De Oliveira DMP, Hartley-Tassell L, Everest-Dass A, Day CJ, Dabbs RA et al. Blood group antigen recognition via the group A streptococcal M protein mediates host colonization. mBio 2017; 8:e02237-16 [View Article]
    [Google Scholar]
  113. De Oliveira DMP, Everest-Dass A, Hartley-Tassell L, Day CJ, Indraratna A et al. Human glycan expression patterns influence Group A streptococcal colonization of epithelial cells. FASEB J 2019; 33:10808–10818 [View Article]
    [Google Scholar]
  114. Zhang Y, Lu P, Pan Z, Zhu Y, Ma J et al. SssP1, a Streptococcus suis fimbria-like protein transported by the SecY2/A2 system, contributes to bacterial virulence. Appl Environ Microbiol 2018; 84:e01385-18 [View Article]
    [Google Scholar]
  115. Bensing BA, Li Q, Park D, Lebrilla CB, Sullam PM et al. Streptococcal Siglec-like adhesins recognize different subsets of human plasma glycoproteins: implications for infective endocarditis. Glycobiology 2018; 28:601–611 [View Article]
    [Google Scholar]
  116. Di Carluccio C, Forgione RE, Bosso A, Yokoyama S, Manabe Y et al. Molecular recognition of sialoglycans by streptococcal Siglec-like adhesins: toward the shape of specific inhibitors. RSC Chem Biol 2021; 2:1618–1630 [View Article] [PubMed]
    [Google Scholar]
  117. Oguchi R, Takahashi Y, Shimazu K, Urano-Tashiro Y, Kawarai T et al. Contribution of Streptococcus gordonii Hsa adhesin to biofilm formation. Jpn J Infect Dis 2017; 70:399–404 [View Article]
    [Google Scholar]
  118. Singh AK, Woodiga SA, Grau MA, King SJ, Pirofski L et al. Streptococcus oralis neuraminidase modulates adherence to multiple carbohydrates on platelets. Infect Immun 2017; 85: [View Article] [PubMed]
    [Google Scholar]
  119. Gaytán MO, Singh AK, Woodiga SA, Patel SA, An S-S et al. A novel sialic acid-binding adhesin present in multiple species contributes to the pathogenesis of Infective endocarditis. PLoS Pathog 2021; 17:e1009222 [View Article]
    [Google Scholar]
  120. Ronis A, Brockman K, Singh AK, Gaytán MO, Wong A et al. Streptococcus oralis subsp. dentisani produces monolateral serine-rich repeat protein fibrils, one of which contributes to saliva binding via sialic acid. Infect Immun 2019; 87:10 [View Article] [PubMed]
    [Google Scholar]
  121. Aparicio D, Torres-Puig S, Ratera M, Querol E, Piñol J et al. Mycoplasma genitalium adhesin P110 binds sialic-acid human receptors. Nat Commun 2018; 9:4471 [View Article] [PubMed]
    [Google Scholar]
  122. Aparicio D, Scheffer MP, Marcos-Silva M, Vizarraga D, Sprankel L et al. Structure and mechanism of the Nap adhesion complex from the human pathogen Mycoplasma genitalium . Nat Commun 2020; 11:2877 [View Article] [PubMed]
    [Google Scholar]
  123. Williams CR, Chen L, Driver AD, Arnold EA, Sheppard ES et al. Sialylated receptor setting influences Mycoplasma pneumoniae attachment and gliding motility. Mol Microbiol 2018; 109:735–744 [View Article] [PubMed]
    [Google Scholar]
  124. Morio H, Kasai T, Miyata M et al. Gliding direction of Mycoplasma mobile . J Bacteriol 2016; 198:283–290 [View Article] [PubMed]
    [Google Scholar]
  125. Hamaguchi T, Kawakami M, Furukawa H, Miyata M et al. Identification of novel protein domain for sialyloligosaccharide binding essential to Mycoplasma mobile gliding. FEMS Microbiol Lett 2019; 366:fnz016 [View Article] [PubMed]
    [Google Scholar]
  126. Dal Peraro M, van der Goot FG. Pore-forming toxins: ancient, but never really out of fashion. Nat Rev Microbiol 2016; 14:77–92 [View Article] [PubMed]
    [Google Scholar]
  127. Shewell LK, Day CJ, Jen FE-C, Haselhorst T, Atack JM et al. All major cholesterol-dependent cytolysins use glycans as cellular receptors. Sci Adv 2020; 6:eaaz4926 [View Article] [PubMed]
    [Google Scholar]
  128. Beddoe T, Paton AW, Le Nours J, Rossjohn J, Paton JC et al. Structure, biological functions and applications of the AB5 toxins. Trends Biochem Sci 2010; 35:411–418 [View Article] [PubMed]
    [Google Scholar]
  129. Wang H, Paton JC, Paton AW. Pathologic changes in mice induced by subtilase cytotoxin, a potent new Escherichia coli AB5 toxin that targets the endoplasmic reticulum. J Infect Dis 2007; 196:1093–1101 [View Article] [PubMed]
    [Google Scholar]
  130. Seyahian EA, Oltra G, Ochoa F, Melendi S, Hermes R et al. Systemic effects of Subtilase cytotoxin produced by Escherichia coli O113:H21. Toxicon 2017; 127:49–55 [View Article] [PubMed]
    [Google Scholar]
  131. Byres E, Paton AW, Paton JC, Löfling JC, Smith DF et al. Incorporation of a non-human glycan mediates human susceptibility to a bacterial toxin. Nature 2008; 456:648–652 [View Article] [PubMed]
    [Google Scholar]
  132. Lee S, Inzerillo S, Lee GY, Bosire EM, Mahato SK et al. Glycan-mediated molecular interactions in bacterial pathogenesis. Trends Microbiol 2021; 30:254–267 [View Article] [PubMed]
    [Google Scholar]
  133. Song J, Gao X, Galán JE et al. Structure and function of the Salmonella Typhi chimaeric A(2)B(5) typhoid toxin. Nature 2013; 499:350–354 [View Article] [PubMed]
    [Google Scholar]
  134. Deng L, Song J, Gao X, Wang J, Yu H et al. Host adaptation of a bacterial toxin from the human pathogen Salmonella Typhi. Cell 2014; 159:1290–1299 [View Article] [PubMed]
    [Google Scholar]
  135. Nguyen T, Lee S, Yang Y-A, Ahn C, Sim JH et al. The role of 9-O-acetylated glycan receptor moieties in the typhoid toxin binding and intoxication. PLoS Pathog 2020; 16:e1008336 [View Article] [PubMed]
    [Google Scholar]
  136. Lee S, Yang Y-A, Milano SK, Nguyen T, Ahn C et al. Salmonella typhoid toxin PltB subunit and its non-typhoidal salmonella ortholog confer differential host adaptation and virulence. Cell Host & Microbe 2020; 27:937–949 [View Article] [PubMed]
    [Google Scholar]
  137. Almagro-Moreno S, Boyd EF. Sialic acid catabolism confers a competitive advantage to pathogenic Vibrio cholerae in the mouse intestine. Infect Immun 2009; 77:3807–3816 [View Article] [PubMed]
    [Google Scholar]
  138. Alisson-Silva F, Liu JZ, Diaz SL, Deng L, Gareau MG et al. Human evolutionary loss of epithelial Neu5Gc expression and species-specific susceptibility to cholera. PLoS Pathog 2018; 14:e1007133 [View Article] [PubMed]
    [Google Scholar]
  139. Zhang S, Berntsson RP-A, Tepp WH, Tao L, Johnson EA et al. Structural basis for the unique ganglioside and cell membrane recognition mechanism of botulinum neurotoxin DC. Nat Commun 2017; 8: [View Article] [PubMed]
    [Google Scholar]
  140. Zalem D, Ribeiro JP, Varrot A, Lebens M, Imberty A et al. Biochemical and structural characterization of the novel sialic acid-binding site of Escherichia coli heat-labile enterotoxin LT-IIb. Biochem J 2016; 473:3923–3936 [View Article] [PubMed]
    [Google Scholar]
http://instance.metastore.ingenta.com/content/journal/micro/10.1099/mic.0.001157
Loading
/content/journal/micro/10.1099/mic.0.001157
Loading

Data & Media loading...

This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error