Skip directly to site content Skip directly to page options Skip directly to A-Z link Skip directly to A-Z link Skip directly to A-Z link
Volume 27, Number 3—March 2021
Research

Foodborne Origin and Local and Global Spread of Staphylococcus saprophyticus Causing Human Urinary Tract Infections

Opeyemi U. Lawal, Maria J. Fraqueza, Ons Bouchami, Peder Worning, Mette D. Bartels, Maria L. Gonçalves, Paulo Paixão, Elsa Gonçalves, Cristina Toscano, Joanna Empel, Małgorzata Urbaś, M. Angeles Domínguez, Henrik Westh, Hermínia de Lencastre, and Maria MiragaiaComments to Author 
Author affiliations: Universidade Nova de Lisboa, Oeiras, Portugal (O.U. Lawal, O. Bouchami, H. de Lencastre, M. Miragaia); Centre for Interdisciplinary Research in Animal Health (CIISA), Universidade de Lisboa, Lisbon, Portugal (M.J., Fraqueza); Hvidovre University Hospital, Hvidovre, Denmark (P. Worning, M.D. Bartels, H. Westh); SAMS Hospital, Lisbon (M.L. Gonçalves); Hospital da Luz, Lisbon (P. Paixão); Hospital Egas Moniz, Lisbon (E. Gonçalves, C. Toscano); Narodowy Instytut Leków, Warsaw, Poland (J. Empel, M. Urbaś); Hospital Universitari de Bellvitge, Barcelona, Spain (M.A. Domínguez); University of Copenhagen, Copenhagen, Denmark (H. Westh); The Rockefeller University, New York, New York, USA (H. de Lencastre)

Main Article

Figure 1

Maximum-likelihood tree of Staphylococcus saprophyticus isolates recovered from human infections and colonization globally, 1997–2017. The tree was constructed by using 9,134 SNPs without recombination. Among analyzed isolates, 321 were recovered from UTIs, 12 from blood, and 4 from colonization. Each node represents a strain; nodes with identical color belong to the same lineage. The assembled contigs were mapped to the reference genome S. saprophyticus ATCC 15305 (GenBank accession no. AP008934.1; black star). Polymorphic sites resulting from recombination events in the single-nucleotide polymorphism (SNP) alignments were filtered out by using Gubbins version 2.3.4 (12). Maximum likelihood tree was reconstructed by using RAxML version 8.2.4 (https://github.com/stamatak/standard-RAxML). We performed generalized time-reversible nucleotide substitution model with gamma correction with 100 bootstraps random resampling for support. We visualized the tree by using Interactive Tree of Life (iTOL; https://itol.embl.de). Black triangles represent isolates fully sequenced by using the long-read nanopore technologies and used as reference to estimate r/m in the respective lineage. Cream color represents clusters G1, G2, G3, G4, and S1, which had dissemination and transmission in same country and in different countries. The outer ring represents isolates’ country of origin; blocks with identical color represent isolates from the same country. Of note, cluster G4 contains a pair of isolates collected in 2016 that had only 10 SNPs difference; one is a blood isolate from Barcelona, Spain (KS266) and the other is a UTI isolate recovered in Lisbon, Portugal (KS135). Scale bar indicates number of substitutions per site. UTI, urinary tract infection; r/m, recombination to mutation ratio.

Figure 1. Maximum-likelihood tree of Staphylococcus saprophyticus isolates recovered from human infections and colonization globally, 1997–2017. The tree was constructed by using 9,134 SNPs without recombination. Among analyzed isolates, 321 were recovered from UTIs, 12 from blood, and 4 from colonization. Each node represents a strain; nodes with identical color belong to the same lineage. The assembled contigs were mapped to the reference genome S. saprophyticus ATCC 15305 (GenBank accession no. AP008934.1; black star). Polymorphic sites resulting from recombination events in the single-nucleotide polymorphism (SNP) alignments were filtered out by using Gubbins version 2.3.4 (12). Maximum likelihood tree was reconstructed by using RAxML version 8.2.4 (https://github.com/stamatak/standard-RAxML). We performed generalized time-reversible nucleotide substitution model with gamma correction with 100 bootstraps random resampling for support. We visualized the tree by using Interactive Tree of Life (iTOL; https://itol.embl.de). Black triangles represent isolates fully sequenced by using the long-read nanopore technologies and used as reference to estimate r/m in the respective lineage. Cream color represents clusters G1, G2, G3, G4, and S1, which had dissemination and transmission in same country and in different countries. The outer ring represents isolates’ country of origin; blocks with identical color represent isolates from the same country. Of note, cluster G4 contains a pair of isolates collected in 2016 that had only 10 SNPs difference; one is a blood isolate from Barcelona, Spain (KS266) and the other is a UTI isolate recovered in Lisbon, Portugal (KS135). Scale bar indicates number of substitutions per site. UTI, urinary tract infection; r/m, recombination to mutation ratio.

Main Article

References
  1. Becker  K, Heilmann  C, Peters  G. Coagulase-negative staphylococci. Clin Microbiol Rev. 2014;27:870926. DOIPubMedGoogle Scholar
  2. Latham  RH, Running  K, Stamm  WE. Urinary tract infections in young adult women caused by Staphylococcus saprophyticus. JAMA. 1983;250:30636. DOIPubMedGoogle Scholar
  3. Garduño  E, Márquez  I, Beteta  A, Said  I, Blanco  J, Pineda  T. Staphylococcus saprophyticus causing native valve endocarditis. Scand J Infect Dis. 2005;37:6901. DOIPubMedGoogle Scholar
  4. Rupp  ME, Soper  DE, Archer  GL. Colonization of the female genital tract with Staphylococcus saprophyticus. J Clin Microbiol. 1992;30:29759. DOIPubMedGoogle Scholar
  5. de Sousa  VS, da-Silva  APS, Sorenson  L, Paschoal  RP, Rabello  RF, Campana  EH, et al. Staphylococcus saprophyticus recovered from humans, food, and recreational waters in Rio de Janeiro, Brazil. Int J Microbiol. 2017;2017:4287547. DOIPubMedGoogle Scholar
  6. Hedman  P, Ringertz  O, Eriksson  B, Kvarnfors  P, Andersson  M, Bengtsson  L, et al. Staphylococcus saprophyticus found to be a common contaminant of food. J Infect. 1990;21:119. DOIPubMedGoogle Scholar
  7. Lee  B, Jeong  D-W, Lee  J-H. Genetic diversity and antibiotic resistance of Staphylococcus saprophyticus isolates from fermented foods and clinical samples. J Korean Soc Appl Biol Chem. 2015;58:65968. DOIGoogle Scholar
  8. Mortimer  TD, Annis  DS, O’Neill  MB, Bohr  LL, Smith  TM, Poinar  HN, et al. Adaptation in a fibronectin binding autolysin of Staphylococcus saprophyticus. MSphere. 2017;2:e005117. DOIPubMedGoogle Scholar
  9. Kuroda  M, Yamashita  A, Hirakawa  H, Kumano  M, Morikawa  K, Higashide  M, et al. Whole genome sequence of Staphylococcus saprophyticus reveals the pathogenesis of uncomplicated urinary tract infection. Proc Natl Acad Sci U S A. 2005;102:132727 . DOIPubMedGoogle Scholar
  10. Hansen  KH, Andreasen  MR, Pedersen  MS, Westh  H, Jelsbak  L, Schønning  K. Resistance to piperacillin/tazobactam in Escherichia coli resulting from extensive IS26-associated gene amplification of blaTEM-1. J Antimicrob Chemother. 2019;74:317983. DOIPubMedGoogle Scholar
  11. Kaas  RS, Leekitcharoenphon  P, Aarestrup  FM, Lund  O. Solving the problem of comparing whole bacterial genomes across different sequencing platforms. PLoS One. 2014;9:e104984. DOIPubMedGoogle Scholar
  12. Croucher  NJ, Page  AJ, Connor  TR, Delaney  AJ, Keane  JA, Bentley  SD, et al. Rapid phylogenetic analysis of large samples of recombinant bacterial whole genome sequences using Gubbins. Nucleic Acids Res. 2015;43:e15. DOIPubMedGoogle Scholar
  13. Brynildsrud  O, Bohlin  J, Scheffer  L, Eldholm  V. Erratum to: Rapid scoring of genes in microbial pan-genome-wide association studies with Scoary. Genome Biol. 2016;17:262. DOIPubMedGoogle Scholar
  14. Chen  L, Zheng  D, Liu  B, Yang  J, Jin  Q. VFDB 2016: hierarchical and refined dataset for big data analysis—10 years on. Nucleic Acids Res. 2016;44(D1):D6947. DOIPubMedGoogle Scholar
  15. Monk  IR, Foster  TJ. Genetic manipulation of Staphylococci-breaking through the barrier. Front Cell Infect Microbiol. 2012;2:49. DOIPubMedGoogle Scholar
  16. Tomita  K, Nagura  T, Okuhara  Y, Nakajima-Adachi  H, Shigematsu  N, Aritsuka  T, et al. Dietary melibiose regulates th cell response and enhances the induction of oral tolerance. Biosci Biotechnol Biochem. 2007;71:277480. DOIPubMedGoogle Scholar
  17. Dinicola  S, Minini  M, Unfer  V, Verna  R, Cucina  A, Bizzarri  M. Nutritional and acquired deficiencies in inositol bioavailability. Correlations with metabolic disorders. Int J Mol Sci. 2017;18:E2187. DOIPubMedGoogle Scholar
  18. Gómez  FA, Cárdenas  C, Henríquez  V, Marshall  SH. Characterization of a functional toxin-antitoxin module in the genome of the fish pathogen Piscirickettsia salmonis. FEMS Microbiol Lett. 2011;317:8392. DOIPubMedGoogle Scholar
  19. Conceição  T, Coelho  C, de Lencastre  H, Aires-de-Sousa  M. High prevalence of biocide resistance determinants in Staphylococcus aureus isolates from three African countries. Antimicrob Agents Chemother. 2015;60:67881. DOIPubMedGoogle Scholar
  20. Costliow  ZA, Degnan  PH. Thiamine acquisition strategies impact metabolism and competition in the gut microbe Bacteroides thetaiotaomicron. mSystems. 2017;2:117. DOIPubMedGoogle Scholar
  21. Chekabab  SM, Harel  J, Dozois  CM. Interplay between genetic regulation of phosphate homeostasis and bacterial virulence. Virulence. 2014;5:78693. DOIPubMedGoogle Scholar
  22. Sharp  JA, Echague  CG, Hair  PS, Ward  MD, Nyalwidhe  JO, Geoghegan  JA, et al. Staphylococcus aureus surface protein SdrE binds complement regulator factor H as an immune evasion tactic. PLoS One. 2012;7:e38407. DOIPubMedGoogle Scholar
  23. Ghali  I, Sofyan  A, Ohmori  H, Shinkai  T, Mitsumori  M. Diauxic growth of Fibrobacter succinogenes S85 on cellobiose and lactose. FEMS Microbiol Lett. 2017;364:19. DOIPubMedGoogle Scholar
  24. Frankel  MB, Hendrickx  APA, Missiakas  DM, Schneewind  O. LytN, a murein hydrolase in the cross-wall compartment of Staphylococcus aureus, is involved in proper bacterial growth and envelope assembly. J Biol Chem. 2011;286:32593605. DOIPubMedGoogle Scholar
  25. Hallet  B, Sherratt  DJ. Transposition and site-specific recombination: adapting DNA cut-and-paste mechanisms to a variety of genetic rearrangements. FEMS Microbiol Rev. 1997;21:15778. DOIPubMedGoogle Scholar
  26. García-Gómez  E, González-Pedrajo  B, Camacho-Arroyo  I. Role of sex steroid hormones in bacterial-host interactions. BioMed Res Int. 2013;2013:928290. DOIPubMedGoogle Scholar
  27. Datta  S, Costantino  N, Zhou  X, Court  DL. Identification and analysis of recombineering functions from Gram-negative and Gram-positive bacteria and their phages. Proc Natl Acad Sci U S A. 2008;105:162631. DOIPubMedGoogle Scholar
  28. Schürch  AC, Arredondo-Alonso  S, Willems  RJL, Goering  RV. Whole genome sequencing options for bacterial strain typing and epidemiologic analysis based on single nucleotide polymorphism versus gene-by-gene-based approaches. Clin Microbiol Infect. 2018;24:3504. DOIPubMedGoogle Scholar
  29. Ronald  A. The etiology of urinary tract infection: traditional and emerging pathogens. Dis Mon. 2003;49:7182. DOIPubMedGoogle Scholar
  30. González-García  S, Belo  S, Dias  AC, Rodrigues  JV, Da Costa  RR, Ferreira  A, et al. Life cycle assessment of pigmeat production: Portuguese case study and proposal of improvement options. J Clean Prod. 2015;100:12639. DOIGoogle Scholar
  31. Chopra  I, Roberts  M. Tetracycline antibiotics: mode of action, applications, molecular biology, and epidemiology of bacterial resistance. Microbiol Mol Biol Rev. 2001;65:23260. DOIPubMedGoogle Scholar
  32. Walker  SW. Laboratory reference ranges. In: Endocrine self-assessment program. Washington, D.C.: Endocrine Society; 2015. p. 1–5 [cited 2019 Apr 29]. https://education.endocrine.org/system/files/ESAP%202015%20Laboratory%20Reference%20Ranges.pdf
  33. Linhares  IM, Minis  E, Robial  R, Witkin  SS. The human vaginal microbiome. In: Faintuch J, Faintuch S, editors. Microbiome and metabolome in diagnosis, therapy, and other strategic applications. London: Elsevier, Inc.; 2019. p. 109–14.
  34. Clarkson  MR, Magee  CN, Brenner  BM. Chapter 2: Laboratory assessment of kidney disease. In: Clarkson MR, Magee CN, Brenner BM eds. Pocket companion to Brenner and Rector’s the Kidney 8th edition. London: Elsevier; 2011. p. 21–41 [cited 2019 Apr 29]. https://www.sciencedirect.com/science/article/pii/B9781416066408000026
  35. Frank  LA, Mullins  R, Rohrbach  BW. Variability of estradiol concentration in normal dogs. Vet Dermatol. 2010;21:4903. DOIPubMedGoogle Scholar
  36. Flores-Mireles  AL, Walker  JN, Caparon  M, Hultgren  SJ. Urinary tract infections: epidemiology, mechanisms of infection and treatment options. Nat Rev Microbiol. 2015;13:26984. DOIPubMedGoogle Scholar
  37. Siboo  IR, Chaffin  DO, Rubens  CE, Sullam  PM. Characterization of the accessory Sec system of Staphylococcus aureus. J Bacteriol. 2008;190:618896. DOIPubMedGoogle Scholar
  38. King  NP, Beatson  SA, Totsika  M, Ulett  GC, Alm  RA, Manning  PA, et al. UafB is a serine-rich repeat adhesin of Staphylococcus saprophyticus that mediates binding to fibronectin, fibrinogen and human uroepithelial cells. Microbiology (Reading). 2011;157:116175. DOIPubMedGoogle Scholar
  39. Arndt  D, Grant  JR, Marcu  A, Sajed  T, Pon  A, Liang  Y, et al. PHASTER: a better, faster version of the PHAST phage search tool. Nucleic Acids Res. 2016;44(W1):W16-21. DOIPubMedGoogle Scholar
  40. Yahara  K, Didelot  X, Jolley  KA, Kobayashi  I, Maiden  MCJ, Sheppard  SK, et al. The landscape of realized homologous recombination in pathogenic bacteria. Mol Biol Evol. 2016;33:45671. DOIPubMedGoogle Scholar
  41. Tristan  A, Bes  M, Meugnier  H, Lina  G, Bozdogan  B, Courvalin  P, et al. Global distribution of Panton-Valentine leukocidin—positive methicillin-resistant Staphylococcus aureus, 2006. Emerg Infect Dis. 2007;13:594600. DOIPubMedGoogle Scholar
  42. Vincent  C, Boerlin  P, Daignault  D, Dozois  CM, Dutil  L, Galanakis  C, et al. Food reservoir for Escherichia coli causing urinary tract infections. Emerg Infect Dis. 2010;16:8895. DOIPubMedGoogle Scholar
  43. Hedman  P, Ringertz  O, Lindström  M, Olsson  K. The origin of Staphylococcus saprophyticus from cattle and pigs. Scand J Infect Dis. 1993;25:5760. DOIPubMedGoogle Scholar
  44. Verstappen  KM, Willems  E, Fluit  AC, Duim  B, Martens  M, Wagenaar  JA. Staphylococcus aureus nasal colonization differs among pig lineages and is associated with the presence of other staphylococcal species. Front Vet Sci. 2017;4:97. DOIPubMedGoogle Scholar
  45. Imperi  F, Leoni  L, Visca  P. Antivirulence activity of azithromycin in Pseudomonas aeruginosa. Front Microbiol. 2014;5:178. DOIPubMedGoogle Scholar
  46. Goerke  C, Köller  J, Wolz  C. Ciprofloxacin and trimethoprim cause phage induction and virulence modulation in Staphylococcus aureus. Antimicrob Agents Chemother. 2006;50:1717. DOIPubMedGoogle Scholar
  47. Davies  MR, McIntyre  L, Mutreja  A, Lacey  JA, Lees  JA, Towers  RJ, et al. Atlas of group A streptococcal vaccine candidates compiled using large-scale comparative genomics. Nat Genet. 2019;51:103543. DOIPubMedGoogle Scholar
  48. Méric  G, Mageiros  L, Pensar  J, Laabei  M, Yahara  K, Pascoe  B, et al. Disease-associated genotypes of the commensal skin bacterium Staphylococcus epidermidis. Nat Commun. 2018;9:5034. DOIPubMedGoogle Scholar
  49. Dearborn  AD, Dokland  T. Mobilization of pathogenicity islands by Staphylococcus aureus strain Newman bacteriophages. Bacteriophage. 2012;2:708. DOIPubMedGoogle Scholar

Main Article

Page created: December 29, 2020
Page updated: February 21, 2021
Page reviewed: February 21, 2021
The conclusions, findings, and opinions expressed by authors contributing to this journal do not necessarily reflect the official position of the U.S. Department of Health and Human Services, the Public Health Service, the Centers for Disease Control and Prevention, or the authors' affiliated institutions. Use of trade names is for identification only and does not imply endorsement by any of the groups named above.
file_external