Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Longitudinal dynamics of farmer and livestock nasal and faecal microbiomes and resistomes

Abstract

Globally, half a billion people are employed in animal agriculture and are directly exposed to the associated microorganisms. However, the extent to which such exposures affect resident human microbiomes is unclear. Here we conducted a longitudinal profiling of the nasal and faecal microbiomes of 66 dairy farmers and 166 dairy cows over a year-long period. We compare farmer microbiomes to those of 60 age-, sex- and ZIP code-matched people with no occupational exposures to farm animals (non-farmers). We show that farming is associated with microbiomes containing livestock-associated microbes; this is most apparent in the nasal bacterial community, with farmers harbouring a richer and more diverse nasal community than non-farmers. Similarly, in the gut microbial communities, we identify more shared microbial lineages between cows and farmers from the same farms. Additionally, we find that shared microbes are associated with antibiotic resistance genes. Overall, our study demonstrates the interconnectedness of human and animal microbiomes.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Farmer nasal microbiome comprises cow-associated microbes.
Fig. 2: Occupational impacts on the gut microbiome are apparent at higher resolutions.
Fig. 3: Functional metagenomics enables cow gut resistome profiling.
Fig. 4: Spread of cow-associated ARGs to farmers.

Similar content being viewed by others

Data availability

All 16S and shotgun sequencing data pertaining to this study are available from the NCBI SRA under BioProject ID PRJNA964705. The databases used in this study include the SILVA database (v.138.1) (https://www.arb-silva.de/documentation/release-1381/), DeconSeq (v.4.3; -dbs hsref38,cow) (https://deconseq.sourceforge.net/), shortBRED (v.0.9.4) (https://github.com/biobakery/biobakery/wiki/shortbred), CARD (v.3.2.2) (https://card.mcmaster.ca/download), NCBI AR gene catalogue (v.2022-04-04.1) (https://www.ncbi.nlm.nih.gov/pathogens/refgene/#), UniRef90 (v.2022-05-29) (https://ftp.uniprot.org/pub/databases/uniprot/uniref/uniref90) and BLASTn (blast-plus) (https://blast.ncbi.nlm.nih.gov/doc/blast-help/downloadblastdata.html#downloadblastdata). Source data are provided with this paper.

Code availability

The code for all computational analyses is available at https://github.com/dantaslab/DOME/tree/main/Scripts.

References

  1. Moving Towards Sustainability: The Livestock Sector and the World Bank (The World Bank, 2020); https://www.worldbank.org/en/topic/agriculture/brief/moving-towards-sustainability-the-livestock-sector-and-the-world-bank

  2. Farming and Farm Income (US Department of Agriculture, 2023); https://www.ers.usda.gov/data-products/ag-and-food-statistics-charting-the-essentials/farming-and-farm-income/

  3. Sowiak, M. et al. An assessment of potential exposure to bioaerosols among swine farm workers with particular reference to airborne microorganisms in the respirable fraction under various breeding conditions. Aerobiologia 28, 121–133 (2012).

    Article  Google Scholar 

  4. Stein, M. M. et al. Innate immunity and asthma risk in Amish and Hutterite farm children. N. Engl. J. Med. 375, 411–421 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Levy, S. B., FitzGerald, G. B. & Macone, A. B. Spread of antibiotic-resistant plasmids from chicken to chicken and from chicken to man. Nature 260, 40–42 (1976).

    Article  CAS  PubMed  Google Scholar 

  6. Angen, O. et al. Transmission of methicillin-resistant Staphylococcus aureus to human volunteers visiting a swine farm. Appl. Environ. Microbiol. 83, e01489-17 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  7. Depner, M. et al. Maturation of the gut microbiome during the first year of life contributes to the protective farm effect on childhood asthma. Nat. Med. 26, 1766–1775 (2020).

    Article  CAS  PubMed  Google Scholar 

  8. Ege, M. J. et al. Exposure to environmental microorganisms and childhood asthma. N. Engl. J. Med. 364, 701–709 (2011).

    Article  CAS  PubMed  Google Scholar 

  9. Illi, S. et al. Protection from childhood asthma and allergy in Alpine farm environments—the GABRIEL advanced studies. J. Allergy Clin. Immunol. 129, 1470–1477 (2012).

    Article  PubMed  Google Scholar 

  10. Steiman, C. A. et al. Patterns of farm exposure are associated with reduced incidence of atopic dermatitis in early life. J. Allergy Clin. Immunol. 146, 1379–1386 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  11. Chen, D. et al. Campylobacter colonization, environmental enteric dysfunction, stunting and associated risk factors among young children in rural Ethiopia: a cross-sectional study from the Campylobacter Genomics and Environmental Enteric Dysfunction (CAGED) Project. Front. Public Health 8, 615793 (2020).

    Article  PubMed  Google Scholar 

  12. Jenkins, P. L., Earle-Richardson, G., Bell, E. M., May, J. J. & Green, A. Chronic disease risk in central New York dairy farmers: results from a large health survey 1989-1999. Am. J. Ind. Med. 47, 20–26 (2005).

    Article  PubMed  Google Scholar 

  13. Eduard, W., Douwes, J., Omenaas, E. & Heederik, D. Do farming exposures cause or prevent asthma? Results from a study of adult Norwegian farmers. Thorax 59, 381–386 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Carnes, M. U. et al. House dust endotoxin levels are associated with adult asthma in a US farming population. Ann. Am. Thorac. Soc. 14, 324–331 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  15. Omland, O., Hjort, C., Pedersen, O. F., Miller, M. R. & Sigsgaard, T. New-onset asthma and the effect of environment and occupation among farming and nonfarming rural subjects. J. Allergy Clin. Immunol. 128, 761–765 (2011).

    Article  PubMed  Google Scholar 

  16. Radon, K., Schulze, A. & Nowak, D. Inverse association between farm animal contact and respiratory allergies in adulthood: protection, underreporting or selection? Allergy 61, 443–446 (2006).

    Article  CAS  PubMed  Google Scholar 

  17. Portengen, L., Preller, L., Tielen, M., Doekes, G. & Heederik, D. Endotoxin exposure and atopic sensitization in adult pig farmers. J. Allergy Clin. Immunol. 115, 797–802 (2005).

    Article  CAS  PubMed  Google Scholar 

  18. Smit, L. A. et al. Exposure–response analysis of allergy and respiratory symptoms in endotoxin-exposed adults. Eur. Respir. J. 31, 1241–1248 (2008).

    Article  CAS  PubMed  Google Scholar 

  19. Fisher, J. A. et al. Residential proximity to intensive animal agriculture and risk of lymphohematopoietic cancers in the agricultural health study. Epidemiology 31, 478–489 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  20. t′Mannetje, A., Eng, A. & Pearce, N. Farming, growing up on a farm and haematological cancer mortality. Occup. Environ. Med. 69, 126–132 (2012).

    Article  PubMed  Google Scholar 

  21. Van Boeckel, T. P. et al. Global trends in antimicrobial use in food animals. Proc. Natl Acad. Sci. USA 112, 5649–5654 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  22. Van Boeckel, T. P. et al. Reducing antimicrobial use in food animals. Science 357, 1350–1352 (2017).

    Article  PubMed  Google Scholar 

  23. Van Boeckel, T. P. et al. Global trends in antimicrobial resistance in animals in low- and middle-income countries. Science 365, eaaw1944 (2019).

    Article  PubMed  Google Scholar 

  24. Pitta, D. W. et al. The distribution of microbiomes and resistomes across farm environments in conventional and organic dairy herds in Pennsylvania. Environ. Microbiome 15, 21 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Mulchandani, R., Wang, Y., Gilbert, M. & Van Boeckel, T. P. Global trends in antimicrobial use in food-producing animals: 2020 to 2030. PLoS Glob. Public Health 3, e0001305 (2023).

    Article  PubMed  PubMed Central  Google Scholar 

  26. Van Den Broek, I. V. F. et al. Methicillin-resistant Staphylococcus aureus in people living and working in pig farms. Epidemiol. Infect. 137, 700–708 (2009).

    Article  PubMed  Google Scholar 

  27. Pirolo, M. et al. Unidirectional animal-to-human transmission of methicillin-resistant Staphylococcus aureus ST398 in pig farming; evidence from a surveillance study in southern Italy. Antimicrob. Resist. Infect. Control 8, 187 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Garcia-Graells, C. et al. Dynamic of livestock-associated methicillin-resistant Staphylococcus aureus CC398 in pig farm households: a pilot study. PLoS ONE 8, e65512 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Cuny, C. et al. Nasal colonization of humans with methicillin-resistant Staphylococcus aureus (MRSA) CC398 with and without exposure to pigs. PLoS ONE 4, e6800 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  30. Larsen, J. et al. Meticillin-resistant Staphylococcus aureus CC398 is an increasing cause of disease in people with no livestock contact in Denmark, 1999 to 2011. Euro Surveill. https://doi.org/10.2807/1560-7917.ES.2015.20.37.30021 (2015).

  31. Liu, C. M. et al. Escherichia coli ST131-H22 as a foodborne uropathogen. mBio 9, e00470-18 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Kuthyar, S. & Reese, A. T. Variation in microbial exposure at the human-animal interface and the implications for microbiome-mediated health outcome. mSystems 6, e0056721 (2021).

    Article  PubMed  Google Scholar 

  33. Sun, J. et al. Environmental remodeling of human gut microbiota and antibiotic resistome in livestock farms. Nat. Commun. 11, 1427 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Kraemer, J. G., Aebi, S., Hilty, M. & Oppliger, A. Nasal microbiota composition dynamics after occupational change in animal farmers suggest major shifts. Sci. Total Environ. 782, 146842 (2021).

    Article  CAS  PubMed  Google Scholar 

  35. Kraemer, J. G., Ramette, A., Aebi, S., Oppliger, A. & Hilty, M. Influence of pig farming on the human nasal microbiota: key role of airborne microbial communities. Appl. Environ. Microbiol. 84, e02470-17 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Abreu, N. A. et al. Sinus microbiome diversity depletion and Corynebacterium tuberculostearicum enrichment mediates rhinosinusitis. Sci. Transl. Med. 4, 151ra124 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  37. VanWormer, J. J., Bendixsen, C. G. & Shukla, S. K. Dairy farm work and protection from gastrointestinal illness. J. Agromed 28, 640–646 (2023).

    Article  Google Scholar 

  38. Blanco-Miguez, A. et al. Extending and improving metagenomic taxonomic profiling with uncharacterized species using MetaPhlAn 4. Nat. Biotechnol. 41, 1633–1644 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Beghini, F. et al. Integrating taxonomic, functional and strain-level profiling of diverse microbial communities with bioBakery 3. eLife 10, e65088 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Antharam, V. C. et al. Intestinal dysbiosis and depletion of butyrogenic bacteria in Clostridium difficile infection and nosocomial diarrhea. J. Clin. Microbiol. 51, 2884–2892 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  41. Olm, M. R. et al. inStrain profiles population microdiversity from metagenomic data and sensitively detects shared microbial strains. Nat. Biotechnol. 39, 727–736 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Manara, S. et al. Microbial genomes from non-human primate gut metagenomes expand the primate-associated bacterial tree of life with over 1,000 novel species. Genome Biol. 20, 299 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Mahmud, B., Boolchandani, M., Patel, S. & Dantas, G. Functional metagenomics to study antibiotic resistance. Methods Mol. Biol. 2601, 379–401 (2023).

    Article  CAS  PubMed  Google Scholar 

  44. Forsberg, K. J. et al. The shared antibiotic resistome of soil bacteria and human pathogens. Science 337, 1107–1111 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Gibson, M. K. et al. Developmental dynamics of the preterm infant gut microbiota and antibiotic resistome. Nat. Microbiol. 1, 16024 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Gasparrini, A. J. et al. Persistent metagenomic signatures of early-life hospitalization and antibiotic treatment in the infant gut microbiota and resistome. Nat. Microbiol. 4, 2285–2297 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  47. Campbell, T. P. et al. The microbiome and resistome of chimpanzees, gorillas, and humans across host lifestyle and geography. ISME J. 14, 1584–1599 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  48. 2021 Summary Report on Antimicrobials Sold or Distributed for Use in Food-producing Animals (US Food and Drug Administration, 2022); https://www.fda.gov/media/163739/download

  49. Alcock, B. P. et al. CARD 2023: expanded curation, support for machine learning and resistome prediction at the Comprehensive Antibiotic Resistance Database. Nucleic Acids Res. 51, D690–D699 (2023).

    Article  CAS  PubMed  Google Scholar 

  50. Feldgarden, M. et al. AMRFinderPlus and the Reference Gene Catalog facilitate examination of the genomic links among antimicrobial resistance, stress response and virulence. Sci. Rep. 11, 12728 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Anthony, W. E. et al. Acute and persistent effects of commonly used antibiotics on the gut microbiome and resistome in healthy adults. Cell Rep. 39, 110649 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. de Nies, L. et al. Evolution of the murine gut resistome following broad-spectrum antibiotic treatment. Nat. Commun. 13, 2296 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

  53. Johnson, T. A. et al. Clusters of antibiotic resistance genes enriched together stay together in swine agriculture. mBio 7, e02214–e02215 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Xihui, Z. et al. Antibiotic resistance of Riemerella anatipestifer and comparative analysis of antibiotic-resistance gene detection methods. Poult. Sci. 102, 102405 (2023).

    Article  PubMed  Google Scholar 

  55. Fenske, G. J., Ghimire, S., Antony, L., Christopher-Hennings, J. & Scaria, J. Integration of culture-dependent and independent methods provides a more coherent picture of the pig gut microbiome. FEMS Microbiol. Ecol. 96, fiaa022 (2020).

    Article  CAS  PubMed  Google Scholar 

  56. Cusco, A., Perez, D., Vines, J., Fabregas, N. & Francino, O. Novel canine high-quality metagenome-assembled genomes, prophages and host-associated plasmids provided by long-read metagenomics together with Hi-C proximity ligation. Micro. Genom. 8, 000802 (2022).

    CAS  Google Scholar 

  57. Brito, I. L. et al. Mobile genes in the human microbiome are structured from global to individual scales. Nature 535, 435–439 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Wallace, M. J., Jean, S., Wallace, M. A., Burnham, C. D. & Dantas, G. Comparative genomics of Bacteroides fragilis group isolates reveals species-dependent resistance mechanisms and validates clinical tools for resistance prediction. mBio 13, e0360321 (2022).

    Article  PubMed  Google Scholar 

  59. Modi, S. R., Lee, H. H., Spina, C. S. & Collins, J. J. Antibiotic treatment expands the resistance reservoir and ecological network of the phage metagenome. Nature 499, 219–222 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Hernando-Amado, S., Coque, T. M., Baquero, F. & Martinez, J. L. Defining and combating antibiotic resistance from One Health and Global Health perspectives. Nat. Microbiol. 4, 1432–1442 (2019).

    Article  CAS  PubMed  Google Scholar 

  61. Kieke, A. L. et al. Validation of health event capture in the Marshfield Epidemiologic Study Area. Clin. Med. Res. 13, 103–111 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  62. Sukhum, K. V. et al. Manure microbial communities and resistance profiles reconfigure after transition to manure pits and differ from those in fertilized field soil. mBio 12, e00798-21 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. D’Souza, A. W. et al. Cotrimoxazole prophylaxis increases resistance gene prevalence and α-diversity but decreases β-diversity in the gut microbiome of Human Immunodeficiency Virus-exposed, uninfected infants. Clin. Infect. Dis. 71, 2858–2868 (2020).

    Article  PubMed  Google Scholar 

  64. Caporaso, J. G. et al. Ultra-high-throughput microbial community analysis on the Illumina HiSeq and MiSeq platforms. ISME J. 6, 1621–1624 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Callahan, B. J. et al. DADA2: high-resolution sample inference from Illumina amplicon data. Nat. Methods 13, 581–583 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Quast, C. et al. The SILVA ribosomal RNA gene database project: improved data processing and web-based tools. Nucleic Acids Res. 41, D590–D596 (2013).

    Article  CAS  PubMed  Google Scholar 

  67. Davis, N. M., Proctor, D. M., Holmes, S. P., Relman, D. A. & Callahan, B. J. Simple statistical identification and removal of contaminant sequences in marker-gene and metagenomics data. Microbiome 6, 226 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  68. Dixon, P. VEGAN, a package of R functions for community ecology. J. Veg. Sci. 14, 927–930 (2003).

    Article  Google Scholar 

  69. Lloyd-Price, J. et al. Erratum: Strains, functions and dynamics in the expanded Human Microbiome Project. Nature 551, 256 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Staff, P. O. Correction: Inexpensive multiplexed library preparation for megabase-sized genomes. PLoS ONE 10, e0131262 (2015).

    Article  Google Scholar 

  71. Bolger, A. M., Lohse, M. & Usadel, B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Schmieder, R. & Edwards, R. Fast identification and removal of sequence contamination from genomic and metagenomic datasets. PLoS ONE 6, e17288 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Li, D., Liu, C. M., Luo, R., Sadakane, K. & Lam, T. W. MEGAHIT: an ultra-fast single-node solution for large and complex metagenomics assembly via succinct de Bruijn graph. Bioinformatics 31, 1674–1676 (2015).

    Article  CAS  PubMed  Google Scholar 

  74. Wu, Y. W., Simmons, B. A. & Singer, S. W. MaxBin 2.0: an automated binning algorithm to recover genomes from multiple metagenomic datasets. Bioinformatics 32, 605–607 (2016).

    Article  CAS  PubMed  Google Scholar 

  75. Kang, D. D. et al. MetaBAT 2: an adaptive binning algorithm for robust and efficient genome reconstruction from metagenome assemblies. PeerJ 7, e7359 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  76. Alneberg, J. et al. Binning metagenomic contigs by coverage and composition. Nat. Methods 11, 1144–1146 (2014).

    Article  CAS  PubMed  Google Scholar 

  77. Sieber, C. M. K. et al. Recovery of genomes from metagenomes via a dereplication, aggregation and scoring strategy. Nat. Microbiol. 3, 836–843 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Parks, D. H., Imelfort, M., Skennerton, C. T., Hugenholtz, P. & Tyson, G. W. CheckM: assessing the quality of microbial genomes recovered from isolates, single cells and metagenomes. Genome Res. 25, 1043–1055 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Chaumeil, P. A., Mussig, A. J., Hugenholtz, P. & Parks, D. H. GTDB-Tk: a toolkit to classify genomes with the Genome Taxonomy Database. Bioinformatics 36, 1925–1927 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  80. Schwengers, O. et al. Bakta: rapid and standardized annotation of bacterial genomes via alignment-free sequence identification. Micro. Genom. 7, 000685 (2021).

    CAS  Google Scholar 

  81. Olm, M. R., Brown, C. T., Brooks, B. & Banfield, J. F. dRep: a tool for fast and accurate genomic comparisons that enables improved genome recovery from metagenomes through de-replication. ISME J. 11, 2864–2868 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Hyatt, D. et al. Prodigal: prokaryotic gene recognition and translation initiation site identification. BMC Bioinformatics 11, 119 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  83. Zankari, E. et al. Identification of acquired antimicrobial resistance genes. J. Antimicrob. Chemother. 67, 2640–2644 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Feldgarden, M. et al. Validating the AMRFinder Tool and Resistance Gene Database by using antimicrobial resistance genotype-phenotype correlations in a collection of isolates. Antimicrob. Agents Chemother. 63, e00483-19 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Gibson, M. K., Forsberg, K. J. & Dantas, G. Improved annotation of antibiotic resistance determinants reveals microbial resistomes cluster by ecology. ISME J. 9, 207–216 (2015).

    Article  CAS  PubMed  Google Scholar 

  86. Kaminski, J. et al. High-specificity targeted functional profiling in microbial communities with ShortBRED. PLoS Comput. Biol. 11, e1004557 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  87. Moore, A. M. et al. Pediatric fecal microbiota harbor diverse and novel antibiotic resistance genes. PLoS ONE 8, e78822 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  88. Moore, A. M. et al. Gut resistome development in healthy twin pairs in the first year of life. Microbiome 3, 27 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  89. Pehrsson, E. C. et al. Interconnected microbiomes and resistomes in low-income human habitats. Nature 533, 212–216 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Forsberg, K. J. et al. Bacterial phylogeny structures soil resistomes across habitats. Nature 509, 612–616 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Clemente, J. C. et al. The microbiome of uncontacted Amerindians. Sci. Adv. 1, e1500183 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  92. Tsukayama, P. et al. Characterization of wild and captive baboon gut microbiota and their antibiotic resistomes. mSystems 3, e00016–e00018 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Suzek, B. E., Huang, H., McGarvey, P., Mazumder, R. & Wu, C. H. UniRef: comprehensive and non-redundant UniProt reference clusters. Bioinformatics 23, 1282–1288 (2007).

    Article  CAS  PubMed  Google Scholar 

  94. Jain, C., Rodriguez, R, L., Phillippy, A. M., Konstantinidis, K. T. & Aluru, S. High throughput ANI analysis of 90K prokaryotic genomes reveals clear species boundaries. Nat. Commun. 9, 5114 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  95. Mallick, H. et al. Multivariable association discovery in population-scale meta-omics studies. PLoS Comput. Biol. 17, e1009442 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Virtanen, P. et al. SciPy 1.0: fundamental algorithms for scientific computing in Python. Nat. Methods 17, 261–272 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank all study participants, particularly the dairy farmers and dairy-farm owners, who, in spite of their busy schedules and harvesting season, participated in the study and also allowed their cows to be enrolled in the study. We thank and acknowledge R. Pilsner, M. Presson and N. Esser of Marshfield Agricultural Station for helping with biospecimen collection from dairy farms. We would like to acknowledge the IACUC of the University of Wisconsin-Madison for reviewing our animal protocol and guidance. We also thank the staff at The Edison Family Center for Genome Sciences & Systems Biology at the Washington University School of Medicine in St Louis, including E. Martin and B. Koebbe for computational support, J. Hoisington-López and M. Crosby for managing the high-throughput sequencing core, and B. Dee, K. Matheny, J. Theodore and K. Page for administrative support. Finally, we would like to thank the members of the Dantas laboratory for helpful general discussions and comments on the manuscript. We thank the National Institute for Occupational Safety and Health of the US Centers for Disease Control and Prevention (grant no. R01OH011578 to G.D. and S.K.S.) for funding support, as well as Marshfield Clinic Research Institute and Weber Endowment Fund (to S.K.S.), the Society for Healthcare Epidemiology of America Research Scholar Award (to K.V.S.), the Initiative for Maximizing Student Development R25 (grant no. GM103757 to R.C.V.) and the Genome Analysis Training Program T32 (grant no. HG000045 to R.C.V.).

Author information

Authors and Affiliations

Authors

Contributions

S.K.S., G.D., C.G.B. and J.J.V. designed the study and secured the funding. S.K.S., G.D., C.G.B., J.J.V. and K.K. strategized the sample collection method. T.K., T.L. and E.K. performed the logistics of sample collection, processing and management of the sample databases. B.M., R.C.V., K.V.S., S.P. and J.L. processed the samples and prepared the sequencing libraries. B.M., R.C.V., L.R.H. and A.K. conducted the computational analysis. B.M. wrote the initial draft of the manuscript, with subsequent review and editing by R.C.V., L.R.H., C.G.B., J.J.V., S.K.S. and G.D. All authors reviewed and approved the final manuscript.

Corresponding authors

Correspondence to Sanjay K. Shukla or Gautam Dantas.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Microbiology thanks Timothy Walsh and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Genus-level richness of nasal and faecal samples.

Genus-level richness of the cow, farmer, and non-farmer nasal (a) and faecal (b) microbiomes across seasons. Boxplots show median (center line), quartiles (box limits), and 1.5x interquartile range (whiskers). Dots correspond to individual samples. Half-violins show the data distribution. P values were calculated using the two-tailed Wilcoxon rank-sum test, with subsequent Benjamini-Hochberg correction for multiple hypotheses.

Source data

Extended Data Fig. 2 Microbial families differentially abundant in nasal samples.

Microbial families with significant (P < 0.05) differential abundances between farmer (n = 145) and non-farmer (n = 103) nasal microbiomes are indicated. For each family, when significant, the corresponding coefficient in cow samples (n = 363) relative to those of non-farmers is also shown. Points denote mean coefficients; whiskers correspond to standard error. Enrichment tested using MaAsLin 2 (see Methods), Benjamini-Hochberg correction for multiple hypotheses.

Source data

Extended Data Fig. 3 Principal coordinate analysis of Bray-Curtis dissimilarities.

Analysis consists of genus (a–h) and species (i-l) compositions of human fecal samples across seasons. Taxonomic profiling is based on 16S (a-d) and shotgun metagenomic (e-l) sequencing data P values were calculated using PERMANOVA and adjusted for multiple hypotheses using the Benjamini-Hochberg method.

Source data

Extended Data Fig. 4 Average fecal Bray-Curtis distance of farmers and non-farmers to cows residing in the same or different collection site.

Beta diversities are based on genus (a) or resistome (b) compositions. Boxplots show median (center line), quartiles (box limits), and 1.5x interquartile range (whiskers). Dots correspond to individual samples. Half-violins show the data distribution. P values were calculated using the two-tailed Wilcoxon rank-sum test, with subsequent Benjamini-Hochberg correction for multiple hypotheses. No significant differences (P < 0.05) were identified.

Source data

Extended Data Fig. 5 Differential abundances of microbiome characteristics across farmers and non-farmers.

Species (a), genera (b), and microbial pathways (c) with significant (P < 0.05) differential abundances between farmer (n = 134) and non-farmer (n = 95) fecal microbiomes are indicated. Points denote mean coefficients; whiskers correspond to standard error. Enrichment tested using MaAsLin 2 (see Methods), Benjamini-Hochberg correction for multiple hypotheses.

Source data

Extended Data Fig. 6 Total ARG relative abundances of cow, farmer, and non-farmer fecal microbiomes across seasons.

Boxplots show median (center line), quartiles (box limits), and 1.5x interquartile range (whiskers). Dots correspond to individual samples. Half-violins show the data distribution. P values were calculated using the two-tailed Wilcoxon rank-sum test, with subsequent Benjamini-Hochberg correction for multiple hypotheses.

Source data

Extended Data Fig. 7 ARGs overrepresented in the cow gut.

ARGs enriched (P < 0.05) in the cow gut resistome (n = 330) relative to that of humans (n = 229). The enrichment coefficients were determined through MaAsLin2 (see Methods), with Benjamini-Hochberg correction for multiple hypotheses. The ARGs are colored according to the corresponding antibiotic classes.

Source data

Extended Data Fig. 8 Farmer and non-farmer gut ARG richness in spring.

The analysis of richness involved only ARGs correlated with genera that i) are enriched in the farmer and cow nasal microbiomes relative to that of non-farmers, and ii) represent the microbial lineages cooccurring the farmer and cow guts. Boxplots show median (center line), quartiles (box limits), and 1.5x interquartile range (whiskers). Dots correspond to individual samples. Half-violins show the data distribution. The P value was calculated using the two-tailed Wilcoxon rank-sum test.

Source data

Extended Data Fig. 9 Rarefaction analysis for 16S rRNA sequencing.

Analysis consisted of fecal (a,c,e) and nasal (b,d,f) samples of cows (a, b), farmers (c, d), and non-farmers (e, f). The analysis was based on genus richness. Boxplots show median (center line), quartiles (box limits), and 1.5x interquartile range (whiskers). Dots correspond to individual subsamples. Half-violins show the data distribution. The differences among subsamples were tested for significance using Dunn’s test, and the P values adjusted for multiple hypotheses using Benjamini-Hochberg. ns, not significant.

Source data

Extended Data Fig. 10 Rarefaction analysis for shotgun sequencing.

Analysis consisted of cow (a), farmer (b), and non-farmer (c) fecal samples. The analysis was based on ARG richness. Boxplots show median (center line), quartiles (box limits), and 1.5x interquartile range (whiskers). Dots correspond to individual subsamples. Half-violins show the data distribution. The differences among subsamples were tested for significance using Dunn’s test, and the P values adjusted for multiple hypotheses using Benjamini-Hochberg. ns, not significant.

Source data

Supplementary information

Supplementary Information

Information guide to Supplementary Tables 1–8.

Reporting Summary

Peer Review File

Supplementary Tables 1–8

Supplementary Tables 1–8.

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 9

Statistical source data.

Source Data Extended Data Fig. 10

Statistical source data.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mahmud, B., Vargas, R.C., Sukhum, K.V. et al. Longitudinal dynamics of farmer and livestock nasal and faecal microbiomes and resistomes. Nat Microbiol 9, 1007–1020 (2024). https://doi.org/10.1038/s41564-024-01639-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41564-024-01639-4

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing