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:

Microbial production and consumption of hydrocarbons in the global ocean

Abstract

Seeps, spills and other oil pollution introduce hydrocarbons into the ocean. Marine cyanobacteria also produce hydrocarbons from fatty acids, but little is known about the size and turnover of this cyanobacterial hydrocarbon cycle. We report that cyanobacteria in an oligotrophic gyre mainly produce n-pentadecane and that microbial hydrocarbon production exhibits stratification and diel cycling in the sunlit surface ocean. Using chemical and isotopic tracing we find that pentadecane production mainly occurs in the lower euphotic zone. Using a multifaceted approach, we estimate that the global flux of cyanobacteria-produced pentadecane exceeds total oil input in the ocean by 100- to 500-fold. We show that rapid pentadecane consumption sustains a population of pentadecane-degrading bacteria, and possibly archaea. Our findings characterize a microbial hydrocarbon cycle in the open ocean that dwarfs oil input. We hypothesize that cyanobacterial hydrocarbon production selectively primes the ocean’s microbiome with long-chain alkanes whereas degradation of other petroleum hydrocarbons is controlled by factors including proximity to petroleum seepage.

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: Pentadecane maps onto trends in ocean fluorescence and cyanobacteria abundance.
Fig. 2: Most pentadecane production in lower euphotic zone.
Fig. 3: Pentadecane rapidly consumed by specialists in waters underlying the oligotrophic ocean.

Similar content being viewed by others

Data availability

All oceanographic, chemical and cell count data are available at the Biological and Chemical Oceanography Data Management Office website under project code NSF OCE-1635562 (https://doi.org/10.26008/1912/bco-dmo.826878.1). Metagenomes are available through the NCBI in BioProject PRJNA657625. Databases accessed were the Genome Taxonomy Database (https://data.ace.uq.edu.au/public/gtdb/data/releases/release89/89.0/, v.r89), the Pfam database (ftp://ftp.ebi.ac.uk/pub/databases/Pfam/releases/Pfam31.0, v.31.0) and the Ocean Microbial Reference Gene Catalogue (http://ocean-microbiome.embl.de/companion.html). Source data are provided with this paper.

References

  1. Oil in the Sea III (National Research Council, 2003).

  2. Han, J., McCarthy, E. D., Hoeven, W. V., Calvin, M. & Bradley, W. H. Organic geochemical studies II. A preliminary report on the distribution of aliphatic hydrocarbons in algae, in bacteria, and in recent lake sediment. Proc. Natl Acad. Sci. USA 59, 29–33 (1968).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Li, X., del Cardayre, S. B., Popova, E., Schirmer, A. & Rude, M. A. Microbial biosynthesis of alkanes. Science 329, 559–562 (2010).

    Article  PubMed  Google Scholar 

  4. Coates, R. C., Podell, S., Korobeynikov, A., Lapidus, A. & Pevzner, P. Characterization of cyanobacterial hydrocarbon composition and distribution of biosynthetic pathways. PLoS ONE 9, 85140 (2014).

    Article  Google Scholar 

  5. White, H. K. et al. Examining inputs of biogenic and oil-derived hydrocarbons in surface waters following the Deepwater Horizon oil spill. ACS Earth Space Chem. 3, 1329–1337 (2019).

    Article  CAS  Google Scholar 

  6. Lea-Smith, D. J. et al. Contribution of cyanobacterial alkane production to the ocean hydrocarbon cycle. Proc. Natl Acad. Sci. USA 112, 13591–13596 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Chisholm, S. W. et al. A novel free-living prochlorophyte abundant in the oceanic euphotic zone. Nature 52, 169–173 (1988).

    Google Scholar 

  8. Field, C. B., Behrenfeld, M. J., Randerson, J. T. & Falkowski, P. Primary production of the biosphere: integrating terrestrial and oceanic components. Science 281, 237–240 (1998).

    Article  CAS  PubMed  Google Scholar 

  9. Flombaum, P. et al. Present and future global distributions of the marine Cyanobacteria Prochlorococcus and Synechococcus. Proc. Natl Acad. Sci. USA 110, 9824–9829 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Karl, D. M. & Church, M. J. Microbial oceanography and the Hawaii Ocean Time-series programme. Nat. Rev. Microbiol. 12, 699–713 (2014).

    Article  CAS  PubMed  Google Scholar 

  11. Polovina, J. J., Howell, E. A. & Abecassis, M. Ocean’s least productive waters are expanding. Geophys. Res. Lett. 35, 2–6 (2008).

    Article  Google Scholar 

  12. Lea-Smith, D. J. et al. Hydrocarbons are essential for optimal cell size, division, and growth of Cyanobacteria. Plant Physiol. 172, 1928–1940 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Cavender-Bares, K. K., Karl, D. M. & Chisholm, S. W. Nutrient gradients in the western North Atlantic Ocean: relationship to microbial community structure and comparison to patterns in the Pacific Ocean. Deep Sea Res. I Oceanogr. Res. Pap. 48, 2373–2395 (2001).

    Article  CAS  Google Scholar 

  14. Johnson, Z. I. et al. Niche partitioning among Prochlorococcus ecotypes along ocean-scale environmental gradients. Science 311, 1737–1740 (2006).

    Article  CAS  PubMed  Google Scholar 

  15. Grande, K. D. et al. Primary production in the North Pacific gyre: a comparison of rates determined by the 14C, O2 concentration and 18O methods. Deep Sea Res. A Oceanogr. Res. Pap. 36, 1621–1634 (1989).

    Article  CAS  Google Scholar 

  16. Karl, D. M. & Church, M. J. Ecosystem structure and dynamics in the North Pacific subtropical gyre: new views of an old ocean. Ecosystems 20, 433–457 (2017).

    Article  Google Scholar 

  17. Knoot, C. J. & Pakrasi, H. B. Diverse hydrocarbon biosynthetic enzymes can substitute for olefin synthase in the cyanobacterium Synechococcus sp. PCC 7002. Sci. Rep. 9, 1360 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  18. Martiny, A. C., Kathuria, S. & Berube, P. M. Widespread metabolic potential for nitrite and nitrate assimilation among Prochlorococcus ecotypes. Proc. Natl Acad. Sci. USA 106, 10787–10792 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Saunois, M. et al. The global methane budget 2000–2017. Earth Syst. Sci. Data 12, 1561–1623 (2020).

    Article  Google Scholar 

  20. Guenther, A. B. et al. The model of emissions of gases and aerosols from nature version 2.1 (MEGAN2.1): an extended and updated framework for modeling biogenic emissions. Geosci. Model Dev. 5, 1471–1492 (2012).

    Article  Google Scholar 

  21. McGenity, T. J., Crombie, A. T. & Murrell, J. C. Microbial cycling of isoprene, the most abundantly produced biological volatile organic compound on Earth. ISME J. 12, 931–941 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Vaulot, D., Marie, D., Olson, R. J. & Chisholm, S. W. Growth of Prochlorococcus, a photosynthetic prokaryote, in the equatorial Pacific. Science 268, 1480–1482 (1995).

    Article  CAS  PubMed  Google Scholar 

  23. Mann, E. L. & Chisholm, S. W. Iron limits the cell division rate of Prochlorococcus in the eastern equatorial Pacific. Limnol. Oceanogr. 45, 1067–1076 (2000).

    Article  CAS  Google Scholar 

  24. Zubkov, M. V. Faster growth of the major prokaryotic versus eukaryotic CO2 fixers in the oligotrophic ocean. Nat. Commun. 5, 3776 (2014).

    Article  PubMed  Google Scholar 

  25. Liu, H. B., Campbell, L. & Landry, M. R. Growth and mortality rates of Prochlorococcus and Synechococcus measured with a selective inhibitor technique. Mar. Ecol. Prog. Ser. 116, 277–288 (1995).

    Article  Google Scholar 

  26. Head, I. M., Jones, D. M. & Larter, S. R. Biological activity in the deep subsurface and the origin of heavy oil. Nature 426, 344–352 (2003).

    Article  CAS  PubMed  Google Scholar 

  27. Reddy, C. M. et al. Composition and fate of gas and oil released to the water column during the Deepwater Horizon oil spill. Proc. Natl Acad. Sci. USA 109, 20229–20234 (2012).

    Article  CAS  PubMed  Google Scholar 

  28. Head, I. M., Jones, D. M. & Röling, W. F. M. Marine microorganisms make a meal of oil. Nat. Rev. Microbiol. 4, 173–182 (2006).

    Article  CAS  PubMed  Google Scholar 

  29. Frysinger, G. S., Gaines, R. B., Xu, L. & Reddy, C. M. Resolving the unresolved complex mixture in petroleum-contaminated sediments. Environ. Sci. Technol. 37, 1653–1662 (2003).

    Article  CAS  PubMed  Google Scholar 

  30. McKenna, A. M. et al. Unprecedented ultrahigh resolution FT-ICR mass spectrometry and parts-per-billion mass accuracy enable direct characterization of nickel and vanadyl porphyrins in petroleum from natural seeps. Energy Fuels 28, 2454–2464 (2014).

    Article  CAS  Google Scholar 

  31. Wardlaw, G. D. et al. Disentangling oil weathering at a marine seep using GCxGC: broad metabolic specificity accompanies subsurface petroleum biodegradation. Environ. Sci. Technol. 42, 7166–7173 (2008).

    Article  CAS  PubMed  Google Scholar 

  32. Wang, W. & Shao, Z. Diversity of flavin-binding monooxygenase genes (almA) in marine bacteria capable of degradation long-chain alkanes. FEMS Microbiol. Ecol. 80, 523–533 (2012).

    Article  CAS  PubMed  Google Scholar 

  33. van Beilen, J. B., Li, Z., Duetz, W. A., Smits, T. H. M. & Witholt, B. Diversity of alkane hydroxylase systems in the environment. Oil Gas Sci. Technol. 58, 427–440 (2003).

    Article  Google Scholar 

  34. Smits, T. H. M., Balada, S. B., Witholt, B. & Van Beilen, J. B. Functional analysis of alkane hydroxylases from Gram-negative and Gram-positive bacteria. J. Bacteriol. 184, 1733–1742 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. González-Gaya, B. et al. Biodegradation as an important sink of aromatic hydrocarbons in the oceans. Nat. Geosci. 12, 119–125 (2019).

    Article  Google Scholar 

  36. Sunagawa, S. et al. Structure and function of the global ocean microbiome. Science 348, 6237 (2015).

    Article  Google Scholar 

  37. Rinke, C. et al. A phylogenomic and ecological analysis of the globally abundant Marine Group II archaea (Ca. Poseidoniales ord. nov.). ISME J. 13, 663–675 (2019).

    Article  CAS  PubMed  Google Scholar 

  38. Tully, B. J. Metabolic diversity within the globally abundant Marine Group II Euryarchaea offers insight into ecological patterns. Nat. Commun. 10, 271 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  39. Hazen, T. C., Prince, R. C. & Mahmoudi, N. Marine oil biodegradation. Environ. Sci. Technol. 50, 2121–2129 (2016).

    Article  CAS  PubMed  Google Scholar 

  40. Prince, R. C., Amande, T. J. & McGenity, T. J. in Taxonomy, Genomics and Ecophysiology of Hydrocarbon-Degrading Microbes (ed. McGenity, T. J.) 1–39 (Springer, 2019).

  41. Valentine, D. L. Adaptations to energy stress dictate the ecology and evolution of the Archaea. Nat. Rev. Microbiol. 5, 316–323 (2007).

    Article  CAS  PubMed  Google Scholar 

  42. MacDonald, I. R. et al. Natural and unnatural oil slicks in the Gulf of Mexico. J. Geophys. Res. Oceans 120, 8364–8380 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Ryerson, T. B. et al. Atmospheric emissions from the Deepwater Horizon spill constrain air-water partitioning, hydrocarbon fate, and leak rate. Geophys. Res. Lett. 38, L07803 (2011).

    Article  Google Scholar 

  44. Ryerson, T. B. et al. Chemical data quantify Deepwater Horizon hydrocarbon flow rate and environmental distribution. Proc. Natl Acad. Sci. USA 109, 20246–20253 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Sorigué, D. et al. Microalgae synthesize hydrocarbons from long-chain fatty acids via a light-dependent pathway. Plant Physiol. 171, 2393–2405 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  46. Sorigué, D. et al. An algal photoenzyme converts fatty acids to hydrocarbons. Science 357, 903–907 (2017).

    Article  PubMed  Google Scholar 

  47. Aleksenko, V. A. et al. Phylogeny and structure of fatty acid photodecarboxylases and glucose-methanol-choline oxidoreductases. Catalysts 10, 1072 (2020).

    Article  CAS  Google Scholar 

  48. Reshkin, S. J. & Knauer, G. A. Light stimulation of phosphate uptake in natural assemblages of phytoplankton. Limnol. Oceanogr. 24, 1121–1124 (1979).

    Article  Google Scholar 

  49. Van Mooy, B. A. S., Moutin, T., Duhamel, S., Rimmelin, P. & Van Wambeke, F. Phospholipid synthesis rates in the eastern subtropical South Pacific Ocean. Biogeosciences 5, 133–139 (2008).

    Article  Google Scholar 

  50. Lomas, M. W. et al. Increased ocean carbon export in the Sargasso Sea linked to climate variability is countered by its enhanced mesopelagic attenuation. Biogeosciences 7, 57–70 (2010).

    Article  CAS  Google Scholar 

  51. Durand, M. D. & Olson, R. J. Contributions of phytoplankton light scattering and cell concentration changes to diel variations in beam attenuation in the equatorial Pacific from flow cytometric measurements of pico-, ultra and nanoplankton. Deep Sea Res. II Top. Stud. Oceanogr. 43, 891–906 (1996).

    Article  Google Scholar 

  52. López-Sandoval, D. C., Delgado-Huertas, A. & Agustí, S. The 13C method as a robust alternative to 14C-based measurements of primary productivity in the Mediterranean Sea. J. Plankton Res. 40, 544–554 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  53. Kozich, J. J., Westcott, S. L., Baxter, N. T., Highlander, S. K. & Schloss, P. D. Development of a dual-index sequencing strategy and curation pipeline for analyzing amplicon sequence data on the MiSeq Illumina sequencing platform. Appl. Environ. Microbiol. 79, 5112–5120 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. 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 

  55. 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 

  56. 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 

  57. Joshi, N. & Fass, J. Sickle: a sliding-window, adaptive, quality-based trimming tool for FastQ files v.1.33 (2011); https://github.com/najoshi/sickle

  58. Nurk, S., Meleshko, D., Korobeynikov, A. & Pevzner, P. A. MetaSPAdes: a new versatile metagenomic assembler. Genome Res. 27, 824–834 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Gurevich, A., Saveliev, V., Vyahhi, N. & Tesler, G. QUAST: quality assessment tool for genome assemblies. Bioinformatics 29, 1072–1075 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  62. Kim, D., Song, L., Breitwieser, F. P. & Salzberg, S. L. Centrifuge: rapid and sensitive classification of metagenomic sequences. Genome Res. 26, 1721–1729 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Eren, A. M. et al. Anvi’o: an advanced analysis and visualization platform for ’omics data. PeerJ 2015, e1319 (2015).

    Article  Google Scholar 

  64. 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 

  65. 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).

    PubMed Central  Google Scholar 

  66. Parks, D. H. et al. A complete domain-to-species taxonomy for Bacteria and Archaea. Nat. Biotechnol. https://doi.org/10.1038/s41587-020-0501-8 (2020).

  67. Tully, B. J., Wheat, C. G., Glazer, B. T. & Huber, J. A. A dynamic microbial community with high functional redundancy inhabits the cold, oxic subseafloor aquifer. ISME J. 12, 1–16 (2018).

    Article  CAS  PubMed  Google Scholar 

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

    Article  PubMed  PubMed Central  Google Scholar 

  69. Eddy, S. R. Accelerated profile HMM searches. PLoS Comput. Biol. 7, 1002195 (2011).

    Article  Google Scholar 

  70. El-Gebali, S. et al. The Pfam protein families database in 2019. Nucleic Acids Res. 47, D427–D432 (2019).

    Article  CAS  PubMed  Google Scholar 

  71. Aramaki, T. et al. KofamKOALA: KEGG Ortholog assignment based on profile HMM and adaptive score threshold. Bioinformatics 36, 2251–2252 (2020).

    Article  CAS  PubMed  Google Scholar 

  72. Shanklin, J., Whittle, E. & Fox, B. G. Eight histidine residues are catalytically essential in a membrane-associated iron enzyme, stearoyl-CoA desaturase, and are conserved in alkane hydroxylase and xylene monooxygenase. Biochemistry 33, 12787–12794 (1994).

    Article  CAS  PubMed  Google Scholar 

  73. Edgar, R. C. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 32, 1792–1797 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Huang, Y., Niu, B., Gao, Y., Fu, L. & Li, W. CD-HIT Suite: a web server for clustering and comparing biological sequences. Bioinformatics 26, 680–682 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Capella-Gutiérrez, S., Silla-Martínez, J. M. & Gabaldón, T. trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25, 1972–1973 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  76. Stamatakis, A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30, 1312–1313 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Rambaut, A. FigTree v.1.4.3 (2012); http://tree.bio.ed.ac.uk/software/figtree/

  78. Sosa, O. A., Repeta, D. J., DeLong, E. F., Ashkezari, M. D. & Karl, D. M. Phosphate-limited ocean regions select for bacterial populations enriched in the carbon–phosphorus lyase pathway for phosphonate degradation. Environ. Microbiol. 21, 2402–2414 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Martinez, A., Tyson, G. W. & Delong, E. F. Widespread known and novel phosphonate utilization pathways in marine bacteria revealed by functional screening and metagenomic analyses. Environ. Microbiol. 12, 222–238 (2010).

    Article  CAS  PubMed  Google Scholar 

  80. Camacho, C. et al. BLAST+: architecture and applications. BMC Bioinformatics 10, 421 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank G. Paradis for analytical support; the R/V Neil Armstrong captain and crew for support at sea; the Hawai’ian Oceanographic Time Series programme and the crew of the R/V Kilo Moana for enabling a preliminary study; J. Hayes and P. Chisholm for their advice on the study; J. Ossolinski, H. Fredricks, B. Jenkins and R. Swarthout for assistance on the R/V Armstrong cruise; T. McKinnon for assistance in dissolved inorganic carbon isotope measurements; A. Ebling for total particulate phosphate measurements; N. Poulton for flow cytometry analysis; and K. Krogslund for nutrient analysis. For bioinformatic analysis, this work used the Extreme Science and Engineering Discovery Environment (XSEDE), which is supported by National Science Foundation (NSF) grant no. ACI-1548562. Specifically, it used the Bridges system, which is supported by NSF award no. ACI-1445606, at the Pittsburgh Supercomputing Center. This project was supported by NSF grant nos. OCE-1635562, OCE-1536346, OCE-1756254 and OCE-1634478.

Author information

Authors and Affiliations

Authors

Contributions

C.R.L. carried out the organization of field experiments and measurements of pentadecane production and concentration, and corresponding data analysis. E.C.A. carried out the biodegradation experiments and bioinformatic work. B.A.S.V.M. contributed nutrient data, cell count data and sediment trap particles for experimentation. K.M.G. carried out the pentadecane concentration measurements. C.M.R. built the methodology for pentadecane quantification. R.K.N. carried out the two-dimensional gas chromatography quality checks. D.L.V. contributed towards experimental design and data interpretation.

Corresponding author

Correspondence to David L. Valentine.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

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

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

Extended data

Extended Data Fig. 1 Pentadecane extraction data quality check.

a, Histogram of percent recovery of the internal standard (DDTP). b, Histogram of standard deviation of replicates of pentadecane concentration measurements; only a few replicates have a standard deviation > 15 ng L−1. c, Pentadecane concentration data vs. percent recovery of DDTP; there is no coherent trend of greater recovery with higher concentration. d, Cyanobacterial cell abundance (Pro. + Syn.) vs. standard deviation of pentadecane concentration between replicates; points with high standard deviation and low cyanobacterial cell abundance were further investigated (see Supplementary Note).

Source data

Extended Data Fig. 2 Pentadecane concentration changes during light incubations.

Concentration of pentadecane at beginning and end of 30-hour light incubations (time = 0 and 30 hours) at three light penetration depths for stations 3, 4, 5, 7, 8, 9 (indicated by number at right of each panel). Water was incubated at the light level from which it was collected (see Methods). Data are plotted as black open circles and represent biologically independent measurements; bar indicates mean of replicates at that light depth, error bars indicate standard deviation of n = 3 replication.

Source data

Extended Data Fig. 3 Pentadecane concentration changes during dark control incubations.

Concentration of pentadecane at beginning and end of 30-hour dark control incubations (time = 0 and 30 hours) at three light penetration depths for stations 3, 4, 5, 7, 8 (indicated by number at right of each panel). Data are plotted as black open circles and represent biologically independent measurements; bar indicates mean of replicates at that light depth, error bars indicate standard deviation of n = 3 replication.

Source data

Extended Data Fig. 4 Station 6 diel patterns across different light depths.

Light depths kept constant through Lagrangian sampling framework whereas the DCM is a depth variable feature throughout the diel cycle (see Methods). The x-axis represents time of day in hours, with gray shading representing night. Diel patterns of a pentadecane, b Prochlorococcus, c Synechococcus, d fluorescence (averaged with 1-meter resolution data with 2 data points above and 2 data points below to smooth signal, n = 5) and e-g selected ratios (see Supplementary Note). a, e, Data are plotted as open circles with n = 2 biologically independent pentadecane measurements, solid circles indicate mean.

Source data

Extended Data Fig. 5 Water mass proxies show consistent diel sampling for lower euphotic zone.

Seawater density plotted against pentadecane concentration colored by light penetration depth and feature (DCM). In this plot, seawater density acts as a proxy for water mass identity in diel sampling. The closer the vertical spread of points of the same color means that samples are more likely to have originated from the same water mass, whereas the further spread means that samples may have originated from different water masses. The horizontal spread of points of the same color represents different concentrations of pentadecane found in the diel cycle. 3% PAR, 1% PAR and particularly the DCM, have pronounced changes in pentadecane over the diel cycle with minimal shifts in seawater density. We conclude this to mean that pentadecane patterns at these depths can be attributed to biological origin, rather than sampling of different water masses. Further information on sampling and data in Methods and Supplementary Note.

Source data

Extended Data Fig. 6 Community composition and diversity in pentadecane incubations.

a-b, Microbial community composition within pentadecane incubations informed via the V4 region of the 16 S rRNA gene for initial samples and those harvested at 27 days (station 3) and 29 days (station 6). Labels on x axis are sample IDs of biologically independent DNA samples with the following abbreviations (T0: time point 0, T0part: initial sediment trap particle community, T0SW (a and b): initial seawater community, T0sw + part (a, b, and c): initial seawater community immediately after particles added, #: pentadecane enrichment). Nucleotide variants are grouped by genus and are listed under associated family and genus; if genus is unclassified than it is listed as NA. All taxa less than 5% are aggregated and shaded gray. c, Shannon diversity index (see Supplementary Note) for each biologically independent DNA sample. Shannon indices for pentadecane (n = 2 at station 3, n = 3 at station 6) and pentadecane + particles (n = 4 at station 3, n = 3 at station 6).

Source data

Extended Data Fig. 7 AlkB diversity underlying the North Atlantic subtropical gyre.

Phylogenetic analysis of genes closely related to alkB from Tara Ocean dataset reveal bacterial and archaeal clades distinct from xylene monooxygenase and fatty acid desaturases. a, Protein domain architecture across select representatives of xylene monooxygenase, alkane-1 monooxygenase, fatty acid desaturase, and related proteins from the Tara Oceans dataset which share a core fatty acid desaturase-like region (blue) expanded on in panel b. b, Abbreviated protein alignment for phylogenetic analyses (for details see Methods). Each column of alignment figure represents a sliding window of 5 bp with the following identity to consensus sequence coloration: green (100%), mustard (80-99% similar), yellow (60-79% similar), gray (<60% similar). The black box represents the region containing the eight histidine residues considered catalytically essential which were used for phylogenetic analyses in panels c-d. c, Maximum-likelihood phylogenetic tree with scale bar of substitutions per site. For clarity, bootstrap values are not shown for the full tree. Δ-X indicates activity X carbons from the carboxylic end of the fatty acid and ω-X indicates activity X carbons from the methyl end of the fatty acid. d, Expanded subtree of membrane monooxygenases and delta-9 fatty acid desaturases (outgroup). Clade coloration in panel d is according to position in panel c. NCBI accession codes are given for functional representatives in the subtree (accession_ORF#).

Source data

Extended Data Fig. 8 Putative archaeal alkB consistently abundant at surface and DCM stations.

a, Maximum-likelihood phylogenetic analysis for each station with scale bar of substitutions per site. Clade designations as follows: green (alkane-1-monoxygenase representatives and related Tara hits), blue (xylene monooxygenase representative and related Tara hits), yellow (putative Marine Group II/III archaeal monooxygenase). See Supplementary Data 2 for homology search results for putative MG II/III monooxygenase hits. Trees <27 unknown Tara sequences are out-grouped with delta-9 fatty acid desaturases, whereas trees with >27 unknown Tara sequences are left unrooted. b, Meta-data for each Tara station and abundance of unique hits derived from read-mapping. % Coverage indicates the fraction of reads that map to genes within each clade (xylene monooxygenase, alkB, or archaeal monooxygenase) over the total reads mapped to all alkB-like, xylene-like, and archaeal monooxygenases found at each station.

Source data

Extended Data Fig. 9 Phylogenetic confirmation of AlkB presence in MAGs from pentadecane incubations.

a, Maximum likelihood tree of alkB hits within metagenomes compared to fatty acid desaturase, xylene monooxygenase, and alkB functionally expressed/characterized representatives (See Supplementary Note for identification details). Δ-X indicates activity X carbons from the carboxylic end of the fatty acid and ω-X indicates activity X carbons from the methyl end of the fatty acid. b, Expanded view of the alkane-1-monooxygenase, xylene monooxygenase, and related hits from metagenomes. Coloration in panel b is according to position in panel a. Gene copies for alkB in MAGS (in green) used in Fig. 3h.

Source data

Extended Data Fig. 10 Proportional prokaryote contribution to phytoplankton community compared across time-series stations.

Depth profiles of ~20 years of data from the Bermuda Atlantic Time-series (BATS, at top, data obtained from Bermuda Atlantic Time-series Study http://bats.bios.edu/bats-data/), the Hawaii Ocean Time-series (HOT, in middle, data obtained from Hawaii Ocean Time-series HOT-DOGS application; University of Hawai’i at Mānoa, National Science Foundation Award #1756517), and this study (at bottom). Data points are colored on a gradient by the proportional contribution to the phytoplankton community by Prochlorococcus and Synechococcus (total phytoplankton community is calculated as Pro. + Syn. + pico- + nano-Eukaryotes for BATS and this study, and Pro. + Syn. + pico-Eukaryotes for HOT). BATS and HOT data are each from a single station measured nearly monthly for ~20 years whereas measurements from this study incorporate spatial variability (see Fig. 1) with minimal temporal variability (all measurements taken in May 2017). The proportional contribution of Prochlorococcus and Synechococcus is >90 % of the phytoplankton community at BATS 84% of the time. At HOT, Pro. + Syn. is > 90 % of phytoplankton community ~100% of the time. For this study, Pro. + Syn. is >90% of the phytoplankton community (Pro. + Syn. + pico + nano-Eukaryotes) for 80% of the measurements, with most cases of lower proportional prokaryote abundance due to an anomalous nutrient pulse observed at station 9 (a Synechococcus bloom) or at low absolute abundance of Pro. + Syn.

Source data

Supplementary information

Supplementary Information

Supplementary Tables 1–4 and Supplementary Note.

Reporting Summary

Supplementary Data 1

Tara Oceans hydrocarbon gene abundance.

Supplementary Data 2

Blast results for Tara Oceans alkB-related hits.

Peer Review File

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 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

16S rRNA community analysis of pentadecane incubations.

Source Data Extended Data Fig. 7

Protein alignment of Tara Oceans alkB-like hits across all North Atlantic subtropical gyre stations, in fasta format.

Source Data Extended Data Fig. 8

Zip file of 8 protein alignments for each individual Tara Ocean station analysed in this study, in fasta format.

Source Data Extended Data Fig. 9

Protein alignment of alkB hits within MAGs, in fasta format.

Source Data Extended Data Fig. 10

Statistical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Love, C.R., Arrington, E.C., Gosselin, K.M. et al. Microbial production and consumption of hydrocarbons in the global ocean. Nat Microbiol 6, 489–498 (2021). https://doi.org/10.1038/s41564-020-00859-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41564-020-00859-8

This article is cited by

Search

Quick links

Nature Briefing Microbiology

Sign up for the Nature Briefing: Microbiology newsletter — what matters in microbiology research, free to your inbox weekly.

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