More stories

  • in

    Ammonia-oxidizing archaea are integral to nitrogen cycling in a highly fertile agricultural soil

    1.Erisman, J. W. et al. Consequences of human modification of the global nitrogen cycle. Philos. Trans. RSoc. Lond. B Biol. Sci. 368, 20130116 (2013).Article 
    CAS 

    Google Scholar 
    2.Fowler, D. et al. The global nitrogen cycle in the Twenty-First Century. Philos. Trans. RSoc. Lond. B Biol. Sci. 368, 20130164 (2013).Article 
    CAS 

    Google Scholar 
    3.Francis, C. A., Beman, J. M. & Kuypers, M. M. M. New processes and players in the nitrogen cycle: the microbial ecology of anaerobic and archaeal ammonia oxidation. ISME J. 1, 19–27 (2007).CAS 
    PubMed 
    Article 

    Google Scholar 
    4.Kuypers, M. M. M., Marchant, H. K. & Kartal, B. The microbial nitrogen-cycling network. Nat. Rev. Microbiol. 16, 263–276 (2018).CAS 
    PubMed 
    Article 

    Google Scholar 
    5.Canfield, D. E., Glazer, A. N. & Falkowski, P. G. The evolution and future of Earth’s nitrogen cycle. Science 330, 192–196 (2010).CAS 
    PubMed 
    Article 

    Google Scholar 
    6.Diaz, R. J. & Rosenberg, R. Spreading dead zones and consequences for marine ecosystems. Science 321, 926–929 (2008).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    7.Gruber, N. & Galloway, J. N. An Earth-system perspective of the global nitrogen cycle. Nature 451, 293–296 (2008).CAS 
    PubMed 
    Article 

    Google Scholar 
    8.Beeckman, F., Motte, H. & Beeckman, T. Nitrification in agricultural soils: impact, actors and mitigation. Curr.Opin. Biotechnol. 50, 166–173 (2018).CAS 
    PubMed 
    Article 

    Google Scholar 
    9.Könneke, M. et al. Isolation of an autotrophic ammonia-oxidizing marine archaeon. Nature 437, 543–546 (2005).PubMed 
    Article 
    CAS 

    Google Scholar 
    10.Tourna, M. et al. Nitrososphaera viennensis, an ammonia oxidizing archaeon from soil. Proc. Natl Acad. Sci. USA. 108, 8420–8425 (2011).CAS 
    PubMed 
    Article 

    Google Scholar 
    11.Daims, H. et al. Complete nitrification by Nitrospira bacteria. Nature. 528, 504–509 (2015).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    12.van Kessel, M. A. H. J. et al. Complete nitrification by a single microorganism. Nature 528, 555–559 (2015).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    13.Leininger, S. et al. Archaea predominate among ammonia-oxidizing prokaryotes in soils. Nature 442, 806–809 (2006).CAS 
    PubMed 
    Article 

    Google Scholar 
    14.Prosser, J. I. & Nicol, G. W. Relative contributions of archaea and bacteria to aerobic ammonia oxidation in the environment. Environ. Microbiol. 10, 2931–2941 (2008).CAS 
    PubMed 
    Article 

    Google Scholar 
    15.Prosser, J. I. & Nicol, G. W. Archaeal and bacterial ammonia-oxidisers in soil: the quest for niche specialisation and differentiation. Trends Microbiol. 20, 523–531 (2012).CAS 
    PubMed 
    Article 

    Google Scholar 
    16.Meinhardt, K. A. et al. Ammonia-oxidizing bacteria are the primary N2O producers in an ammonia-oxidizing archaea dominated alkaline agricultural soil. Environ. Microbiol. 20, 2195–2206 (2018).CAS 
    PubMed 
    Article 

    Google Scholar 
    17.Di, H. J. et al. Nitrification driven by bacteria and not archaea in nitrogen-rich grassland soils. Nat. Geosci. 2, 621–624 (2009).CAS 
    Article 

    Google Scholar 
    18.Prosser, J. I., Hink, L., Gubry-Rangin, C. & Nicol, G. W. Nitrous oxide production by ammonia oxidizers: physiological diversity, niche differentiation and potential mitigation strategies. Glob. Change Biol. 26, 103–118 (2020).Article 

    Google Scholar 
    19.Norton, J. & Ouyang, Y. Controls and adaptive management of nitrification in agricultural soils. Front. Microbiol. 10, 1931 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    20.Pjevac, P. et al. AmoA-targeted polymerase chain reaction primers for the specific detection and quantification of comammox Nitrospira in the environment. Front. Microbiol. 8, 1508 (2017).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    21.Lawson, C. E. & Lücker, S. Complete ammonia oxidation: an important control on nitrification in engineered ecosystems? Curr. Opin. Biotechnol. 50, 158–165 (2018).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    22.Orellana, L. H., Chee-Sanford, J. C., Sanford, R. A., Löffler, F. E. & Konstantinidis, K. T. Year-round shotgun metagenomes reveal stable microbial communities in agricultural soils and novel ammonia oxidizers responding to fertilization. Appl. Environ. Microbiol. 84, e01646–01617 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    23.Kits, K. D. et al. Low yield and abiotic origin of N2O formed by the complete nitrifier Nitrospira inopinata. Nat. Commun. 10, 1836 (2019).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    24.Kits, K. D. et al. Kinetic analysis of a complete nitrifier reveals an oligotrophic lifestyle. Nature. 549, 269–272 (2017).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    25.Stein, L. Y. Insights into the physiology of ammonia-oxidizing microorganisms. Curr. Opin. Chem. Biol. 49, 9–15 (2019).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    26.Lehtovirta-Morley, L. E. Ammonia oxidation: ecology, physiology, biochemistry and why they must all come together. FEMS Microbiol. Lett. 365, fny058–fny058 (2018).Article 
    CAS 

    Google Scholar 
    27.Lu, X., Taylor, A. E., Myrold, D. D. & Neufeld, J. D. Expanding perspectives of soil nitrification to include ammonia-oxidizing archaea and comammox bacteria. Soil Sci. Soc. Am J. 84, 287–302 (2020).CAS 
    Article 

    Google Scholar 
    28.Taylor, A. E., Zeglin, L. H., Wanzek, T. A., Myrold, D. D. & Bottomley, P. J. Dynamics of ammonia-oxidizing archaea and bacteria populations and contributions to soil nitrification potentials. ISME J. 6, 2024–2032 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    29.Taylor, A. E. et al. Use of aliphatic n-alkynes to discriminate soil nitrification activities of ammonia-oxidizing Thaumarchaea and Bacteria. Appl. Environ. Microbiol. 79, 6544–6551 (2013).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    30.Taylor, A. E., Giguere, A. T., Zoebelein, C. M., Myrold, D. D. & Bottomley, P. J. Modeling of soil nitrification responses to temperature reveals thermodynamic differences between ammonia-oxidizing activity of archaea and bacteria. ISME J. 11, 896–908 (2017).CAS 
    PubMed 
    Article 

    Google Scholar 
    31.Hink, L., Gubry-Rangin, C., Nicol, G. W. & Prosser, J. I. The consequences of niche and physiological differentiation of archaeal and bacterial ammonia oxidisers for nitrous oxide emissions. ISME J. 12, 1084–1093 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    32.Hink, L., Nicol, G. W. & Prosser, J. I. Archaea produce lower yields of N2O than bacteria during aerobic ammonia oxidation in soil. Environ. Microbiol. 19, 4829–4837 (2017).CAS 
    PubMed 
    Article 

    Google Scholar 
    33.Levičnik-Höfferle, Š., Nicol, G. W., Ausec, L., Mandić-Mulec, I. & Prosser, J. I. Stimulation of Thaumarchaeal ammonia oxidation by ammonia derived from organic nitrogen but not added inorganic nitrogen. FEMS Microbiol. Ecol. 80, 114–123 (2012).PubMed 
    Article 
    CAS 

    Google Scholar 
    34.Stopnisek, N. et al. Thaumarchaeal ammonia oxidation in an acidic forest peat soil is not Influenced by ammonium amendment. Appl. Environ. Microbiol. 76, 7626–7634 (2010).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    35.Verhamme, D. T., Prosser, J. I. & Nicol, G. W. Ammonia concentration determines differential growth of ammonia-oxidising archaea and bacteria in soil microcosms. ISME J. 5, 1067–1071 (2011).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    36.Rodriguez, A. F., Gerber, S. & Daroub, S. H. Modeling soil subsidence in a subtropical drained peatland. The case of the everglades agricultural Area. Ecol. Modelling. 415, 108859 (2020).Article 

    Google Scholar 
    37.Terry, R. E. Nitrogen mineralization in Florida histosols. Soil Sci. Soc. Am. J. 44, 747–750 (1980).CAS 
    Article 

    Google Scholar 
    38.Zhalnina, K. et al. Ca. Nitrososphaera and Bradyrhizobium are inversely correlated and related to agricultural practices in long-term field experiments. Front. Microbiol. 4, 104 (2013).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    39.Hart S. C., Stark, J. M., Davidson, E. A., Firestone, M. K. Nitrogen mineralization, immobilization, and nitrification. In Methods of soil analysis (eds, Weaver, R.W., Angle, S., Bottomley, P., Bezdicek, D., Smith, S., Tabatabai, A. et al). pp 985–1018. (Soil Science Society of America, 1994).40.Martens-Habbena, W. et al. The production of nitric oxide by marine ammonia-oxidizing archaea and inhibition of archaeal ammonia oxidation by a nitric oxide scavenger. Environ. Microbiol. 17, 2261–2274 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    41.Tourna, M., Freitag, T. E., Nicol, G. W. & Prosser, J. I. Growth, activity and temperature responses of ammonia-oxidizing archaea and bacteria in soil microcosms. Environ. Microbiol. 10, 1357–1364 (2008).CAS 
    PubMed 
    Article 

    Google Scholar 
    42.Rotthauwe, J. H., Witzel, K. P. & Liesack, W. The ammonia monooxygenase structural gene amoA as a functional marker: molecular fine-scale analysis of natural ammonia-oxidizing populations. Appl. Environ. Microbiol. 63, 4704–4712 (1997).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    43.Hill, J. T. et al. Poly peak parser: Method and software for identification of unknown indels using sanger sequencing of polymerase chain reaction products. Devel Dyn. 243, 1632–1636 (2014).CAS 
    Article 

    Google Scholar 
    44.Ludwig, W. et al. ARB: a software environment for sequence data. Nucleic Acids Res. 32, 1363–1371 (2004).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    45.Caporaso, J. G. et al. QIIME allows analysis of high-throughput community sequencing data. Nat. Methods. 7, 335–336 (2010).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    46.Thompson, L. R. et al. A communal catalogue reveals Earth’s multiscale microbial diversity. Nature 551, 457–463 (2017).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    47.Parada, A. E., Needham, D. M. & Fuhrman, J. A. Every base matters: assessing small subunit rRNA primers for marine microbiomes with mock communities, time series and global field samples. Environ. Microbiol. 18, 1403–1414 (2016).CAS 
    Article 

    Google Scholar 
    48.Apprill, A., McNally, S., Parsons, R. & Weber, L. Minor revision to V4 region SSU rRNA 806R gene primer greatly increases detection of SAR11 bacterioplankton. Aquat. Microb. Ecol. 75, 129–137 (2015).Article 

    Google Scholar 
    49.Callahan, B. J. et al. DADA2: High-resolution sample inference from Illumina amplicon data. Nat. Methods. 13, 581–583 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    50.Bolyen, E. et al. Reproducible, interactive, scalable and extensible microbiome data science using QIIME 2. Nat. Biotechnol. 37, 852–857 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    51.Quast, C. et al. The SILVA ribosomal RNA gene database project: improved data processing and web-based tools. Nucleic Acids Res. 41, D590–D596 (2012).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    52.Oksanen J., et al. vegan: Community ecology package. R package version 2.5-6. https://CRAN.R-project.org/package=vegan. (2019).53.R Core Team. R: A language and environment for statistical computing. (R Foundation for Statistical Computing, 2020).54.Wickham H. ggplot2: Elegant graphics for data analysis. (Springer, 2016).55.Ouyang, Y., Norton, J. M. & Stark, J. M. Ammonium availability and temperature control contributions of ammonia oxidizing bacteria and archaea to nitrification in an agricultural soil. Soil Biol. Biochem. 113, 161–172 (2017).CAS 
    Article 

    Google Scholar 
    56.Ouyang, Y., Evans, S. E., Friesen, M. L. & Tiemann, L. K. Effect of nitrogen fertilization on the abundance of nitrogen cycling genes in agricultural soils: A meta-analysis of field studies. Soil Biol. Biochem. 127, 71–78 (2018).CAS 
    Article 

    Google Scholar 
    57.Ouyang, Y., Norton, J. M., Stark, J. M., Reeve, J. R. & Habteselassie, M. Y. Ammonia-oxidizing bacteria are more responsive than archaea to nitrogen source in an agricultural soil. Soil Biol. Biochem. 96, 4–15 (2016).CAS 
    Article 

    Google Scholar 
    58.Norton, J. M., Alzerreca, J. J., Suwa, Y. & Klotz, M. G. Diversity of ammonia monooxygenase operon in autotrophic ammonia-oxidizing bacteria. Arch. Microbiol. 177, 139–149 (2002).CAS 
    PubMed 
    Article 

    Google Scholar 
    59.Shen, T., Stieglmeier, M., Dai, J., Urich, T. & Schleper, C. Responses of the terrestrial ammonia-oxidizing archaeon Ca. Nitrososphaera viennensis and the ammonia-oxidizing bacterium Nitrosospira multiformis to nitrification inhibitors. FEMS Microbiol. Lett. 344, 121–129 (2013).CAS 
    PubMed 
    Article 

    Google Scholar 
    60.Sauder, L. A., Ross, A. A. & Neufeld, J. D. Nitric oxide scavengers differentially inhibit ammonia oxidation in ammonia-oxidizing archaea and bacteria. FEMS Microbiol. Lett. 363, fnw052 (2016).PubMed 
    Article 
    CAS 
    PubMed Central 

    Google Scholar 
    61.Stieglmeier, M. et al. Aerobic nitrous oxide production through N-nitrosating hybrid formation in ammonia-oxidizing archaea. ISME J. 8, 1135–1146 (2014).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    62.Erguder, T. H., Boon, N., Wittebolle, L., Marzorati, M. & Verstraete, W. Environmental factors shaping the ecological niches of ammonia-oxidizing archaea. FEMS Microbiol. Rev. 33, 855–869 (2009).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    63.Zhalnina, K., Dörr de Quadros, P., Camargo, F. A. O. & Triplett, E. W. Drivers of archaeal ammonia-oxidizing communities in soil. Front Microbiol. 3, 210 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    64.Thion, C. E. et al. Plant nitrogen-use strategy as a driver of rhizosphere archaeal and bacterial ammonia oxidiser abundance. FEMS Microbiol. Ecol. 92, fiw091 (2016).PubMed 
    Article 
    CAS 
    PubMed Central 

    Google Scholar 
    65.Coskun, D., Britto, D. T., Shi, W. & Kronzucker, H. J. How plant root exudates shape the nitrogen cycle. Trends Plant Sci. 22, 661–673 (2017).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    66.Marschner P. Mineral nutrition of higher plants. (Academic Press, 2012).67.Coskun, D., Britto, D. T., Shi, W. & Kronzucker, H. J. Nitrogen transformations in modern agriculture and the role of biological nitrification inhibition. Nat. Plants. 3, 17074 (2017).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    68.Masclaux-Daubresse, C. et al. Nitrogen uptake, assimilation and remobilization in plants: challenges for sustainable and productive agriculture. Ann. Bot. 105, 1141–1157 (2010).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    69.Button, D. K., Robertson, B. R., Lepp, P. W. & Schmidt, T. M. A small, dilute-cytoplasm, high-affinity, novel bacterium isolated by extinction culture and having kinetic constants compatible with growth at ambient concentrations of dissolved nutrients in seawater. Appl. Environ. Microbiol. 64, 4467–4476 (1998).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    70.Martens-Habbena, W., Berube, P. M., Urakawa, H., Torre, J. R. & Stahl, D. A. Ammonia oxidation kinetics determine niche separation of nitrifying archaea and bacteria. Nature. 461, 976–979 (2009).CAS 
    PubMed 
    Article 

    Google Scholar 
    71.Ferreira, D. A. et al. Contribution of N from green harvest residues for sugarcane nutrition in Brazil. GCB Bioenergy. 8, 859–866 (2016).Article 

    Google Scholar 
    72.Li, J., Pei, J., Pendall, E., Fang, C. & Nie, M. Spatial heterogeneity of temperature sensitivity of soil respiration: a global analysis of field observations. Soil Biol. Biochem. 141, 107675 (2020).CAS 
    Article 

    Google Scholar 
    73.Maathuis, F. J. M. Physiological functions of mineral macronutrients. Curr. Opin. Plant Biol. 12, 250–258 (2009).CAS 
    PubMed 
    Article 

    Google Scholar 
    74.Song, G. C. et al. Plant growth-promoting archaea trigger induced systemic resistance in Arabidopsis thaliana against Pectobacterium carotovorum and Pseudomonas syringae. Environ. Microbiol. 21, 940–948 (2019).CAS 
    PubMed 
    Article 

    Google Scholar  More

  • in

    Multi-species and multi-tissue methylation clocks for age estimation in toothed whales and dolphins

    1.Beal, A. P., Kiszka, J. J., Wells, R. S. & Eirin-Lopez, J. M. The Bottlenose dolphin Epigenetic Aging Tool (BEAT): a molecular age estimation tool for small cetaceans. Front. Mar. Sci. https://doi.org/10.3389/fmars.2019.00561 (2019).2.Garde, E., Heide-Jørgensen, M. P., Hansen, S. H., Nachman, G. & Forchhammer, M. C. Age-specific growth and remarkable longevity in narwhals (Monodon monoceros) from West Greenland as estimated by aspartic acid racemization. J. Mammal. 88, 49–58 (2007).Article 

    Google Scholar 
    3.Matkin, C. O., Ward Testa, J., Ellis, G. M. & Saulitis, E. L. Life history and population dynamics of southern Alaska resident killer whales (Orcinus orca). Mar. Mammal. Sci. 30, 460–479 (2014).Article 

    Google Scholar 
    4.Olesiuk, P., Bigg, M. & Ellis, G. Life history and population dynamics of resident killer whales (Orcinus orca) in the coastal waters of British Columbia and Washington State. Report of the International Whaling Commission. Special 12, 209–243 (1990).
    Google Scholar 
    5.Wells, R. S. Primates and Cetaceans: Field Research and Conservation of Complex Mammalian Societies, Primatology Monographs (eds. J. Yamagiwa, & Karczmarski, L.) p. 149–172 (Springer, 2014).6.Robeck, T. R., Willis, K., Scarpuzzi, M. R. & O’Brien, J. K. Survivorship pattern inaccuracies and inappropriate anthropomorphism in scholarly pursuits of killer whale (Orcinus orca) life history: a response to Franks et al.(2016). J. Mammal. 97, 899–905 (2016).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    7.Ellis, S. et al. Analyses of ovarian activity reveal repeated evolution of post-reproductive lifespans in toothed whales. Sci. Rep. 8, 1–10 (2018).CAS 
    Article 

    Google Scholar 
    8.Croft, D. P., Brent, L. J., Franks, D. W. & Cant, M. A. The evolution of prolonged life after reproduction. Trends Ecol. Evol. 30, 407–416 (2015).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    9.Wursig, B. & Jefferson, T. A. Methods of photo-identification for small cetaceans. Rep. Int. Whal. Comm. 12, 43–52 (1990).
    Google Scholar 
    10.Perrin, W. F. & Myrick, A. C. Age Determination Of Toothed Whales And Sirenians (International Whaling Commission, 1980).11.Bryden, M. Research on Dolphins (eds. Bryden, M. M. & Harrison, R. J.) p. 211–224 (Clarendon Press Oxford, 1986).12.Myrick, A. C., Yochem, P. K. & Cornell, L. H. Toward calibrating dentinal layers in captive killer whales by use of tetracycline labels. Rit Fiskid. 11, 285–296 (1988).
    Google Scholar 
    13.Best, P., Meÿer, M. & Lockyer, C. Killer whales in South African waters—a review of their biology. Afr. J. Mar. Sci. 32, 171–186 (2010).Article 

    Google Scholar 
    14.Foote, A. D., Newton, J., Piertney, S. B., Willerslev, E. & Gilbert, M. T. P. Ecological, morphological and genetic divergence of sympatric North Atlantic killer whale populations. Mol. Ecol. 18, 5207–5217 (2009).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    15.Ford, J. K. et al. Shark predation and tooth wear in a population of northeastern Pacific killer whales. Aquat. Biol. 11, 213–224 (2011).Article 

    Google Scholar 
    16.Hohn, A. A. & Fernandez, S. Biases in dolphin age structure due to age estimation technique. Mar. Mammal. Sci. 15, 1124–1132 (1999).Article 

    Google Scholar 
    17.Lockyer, C. A report on patterns of deposition of dentine and cement in teeth of pilot whales, genus Globicephala. Rep. Int. Whal. Comm. 14, 137–161 (1993).
    Google Scholar 
    18.Waugh, D. A., Suydam, R. S., Ortiz, J. D. & Thewissen, J. Validation of Growth Layer Group (GLG) depositional rate using daily incremental growth lines in the dentin of beluga (Delphinapterus leucas (Pallas, 1776)) teeth. PLoS ONE 13, e0190498 (2018).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    19.Sergeant, D. E. Age Determination In Odontocete Whales From Dentinal Growth Layers (Norwegian Whaling Gazette, 1959).20.Brodie, P. F. Mandibular layering in Delphinapterus leucas and age determination. Nature 221, 956–958 (1969).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    21.Goren, A. D. et al. Growth layer groups (GLGs) in the teeth of an adult belukha whale (Delphinapterus leucas) of known age: evidence for two annual layers. Mar. Mammal. Sci. 3, 14–21 (1987).Article 

    Google Scholar 
    22.Brodie, P. & Haulena, M. Dentinal growth layer counts of captive, known-age, mother and daughter belugas (Delphinapterus leucas): confirming two growth layer groups (GLG/2) per year; consequences for recovery and management. J Cetacean. Res Manag. 18, 23–31 (2018).
    Google Scholar 
    23.Brodie, P., Ramirez, K. & Haulena, M. Growth and maturity of belugas (Delphinapterus leucas) in Cumberland Sound, Canada, and in captivity: evidence for two growth layer groups (GLGs) per year in teeth. J. Cetacean Res. Manag. 13, 1–18 (2013).
    Google Scholar 
    24.Lockyer, C., Hohn, A. A., Doidge, D. W., Heide-Jørgensen, M. P. & Suydam, R. Age determination in belugas (Delphinapterus leucas in Belugas): a quest for validation of dentinal layering. Aquat. Mamm. 33, 293–304 (2007).Article 

    Google Scholar 
    25.Stewart, R., Campana, S., Jones, C. & Stewart, B. Bomb radiocarbon dating calibrates beluga (Delphinapterus leucas) age estimates. Can. J. Zool. 84, 1840–1852 (2006).Article 

    Google Scholar 
    26.Brodie, P. A reconsideration of aspects of growth, reproduction, and behavior of the white whale (Delphinapterus leucas), with reference to the Cumberland Sound, Baffin Island, population. J. Fish. Board Can. 28, 1309–1318 (1971).Article 

    Google Scholar 
    27.Brodie, P. F., Parsons, J. L. & Sergeant, D. E. Present status of the white whale (Delphinapterus leucas) in Cumberland Sound, Baffin Island.Rep. Int. Whal. Comm. 31, 579–582 (1981).
    Google Scholar 
    28.Robeck, T. R. et al. Reproduction, growth and development in captive beluga (Delphinapterus leucas). Zoo Biol. 24, 29–49 (2005).Article 

    Google Scholar 
    29.Bada, J., Brown, S. & Masters, P. Age determination of marine mammals based on aspartic acid racemization in the teeth and lens nucleus. Age Determination of Toothed Whales and Sirenians. p. 113–118 (Report of the International Whaling Commission, Special, 1980).30.George, J. C. et al. Age and growth estimates of bowhead whales (Balaena mysticetus) via aspartic acid racemization. Can. J. Zool. 77, 571–580 (1999).Article 

    Google Scholar 
    31.Pleskach, K. et al. Use of mass spectrometry to measure aspartic acid racemization for ageing beluga whales. Mar. Mammal. Sci. 32, 1370–1380 (2016).CAS 
    Article 

    Google Scholar 
    32.Garde, E., Peter Heide-Jørgensen, M., Ditlevsen, S. & Hansen, S. H. Aspartic acid racemization rate in narwhal (Monodon monoceros) eye lens nuclei estimated by counting of growth layers in tusks. Polar Res. https://doi.org/10.3402/polar.v31i0.15865 (2012).33.Herman, D. P. et al. Assessing age distributions of killer whale Orcinus orca populations from the composition of endogenous fatty acids in their outer blubber layers. Mar. Ecol. Prog. Ser. 372, 289–302 (2008).CAS 
    Article 

    Google Scholar 
    34.Herman, D. P. et al. Age determination of humpback whales Megaptera novaeangliae through blubber fatty acid compositions of biopsy samples. Mar. Ecol. Prog. Ser. 392, 277–293 (2009).Article 

    Google Scholar 
    35.Marcoux, M., Lesage, V., Thiemann, G. W., Iverson, S. J. & Ferguson, S. H. Age estimation of belugas, Delphinapterus leucas, using fatty acid composition: a promising method. Mar. Mammal. Sci. 31, 944–962 (2015).Article 

    Google Scholar 
    36.Olsen, M. T., Berube, M., Robbins, J. & Palsboll, P. J. Empirical evaluation of humpback whale telomere length estimates; quality control and factors causing variability in the singleplex and multiplex qPCR methods. BMC Genet. 13, 77 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    37.Broer, L. et al. Meta-analysis of telomere length in 19 713 subjects reveals high heritability, stronger maternal inheritance and a paternal age effect. Eur. J. Hum. Genet. 21, 1163–1168 (2013).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    38.Dunshea, G. et al. Telomeres as age markers in vertebrate molecular ecology. Mol. Ecol. Resour. 11, 225–235 (2011).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    39.Polanowski, A. M., Robbins, J., Chandler, D. & Jarman, S. N. Epigenetic estimation of age in humpback whales. Mol. Ecol. Resour. 14, 976–987 (2014).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    40.Tanabe, A. et al. Age estimation by DNA methylation in the Antarctic minke whale. Fish. Sci. 86, 35–41 (2020).CAS 
    Article 

    Google Scholar 
    41.Smith, Z. D. & Meissner, A. DNA methylation: roles in mammalian development. Nat. Rev. Genet. 14, 204–220 (2013).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    42.Rakyan, V. K. et al. Human aging-associated DNA hypermethylation occurs preferentially at bivalent chromatin domains. Genome Res. 20, 434–439 (2010).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    43.Teschendorff, A. E. et al. Age-dependent DNA methylation of genes that are suppressed in stem cells is a hallmark of cancer. Genome Res. 20, 440–446 (2010).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    44.Horvath, S. & Raj, K. DNA methylation-based biomarkers and the epigenetic clock theory of ageing. Nat. Rev. Genet. 19, 371–384 (2018).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    45.Field, A. E. et al. DNA methylation clocks in aging: categories, causes, and consequences. Mol. Cell 71, 882–895 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    46.Horvath, S. DNA methylation age of human tissues and cell types. Genome Biol. 14, R115 (2013).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    47.Bell, C. G. et al. DNA methylation aging clocks: challenges and recommendations. Genome Biol. 20, 1–24 (2019).Article 

    Google Scholar 
    48.Petkovich, D. A. et al. Using DNA methylation profiling to evaluate biological age and longevity interventions. Cell Metab. 25, 954–960.e956 (2017).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    49.Cole, J. J. et al. Diverse interventions that extend mouse lifespan suppress shared age-associated epigenetic changes at critical gene regulatory regions. Genome Biol. 18, 1–16 (2017).Article 
    CAS 

    Google Scholar 
    50.Wang, T. et al. Epigenetic aging signatures in mice livers are slowed by dwarfism, calorie restriction and rapamycin treatment. Genome Biol. 18, 57 (2017).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    51.Stubbs, T. M. et al. Multi-tissue DNA methylation age predictor in mouse. Genome Biol. 18, 1–14 (2017).Article 
    CAS 

    Google Scholar 
    52.Thompson, M. J. et al. A multi-tissue full lifespan epigenetic clock for mice. Aging 10, 2832 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    53.Meer, M. V., Podolskiy, D. I., Tyshkovskiy, A. & Gladyshev, V. N. A whole lifespan mouse multi-tissue DNA methylation clock. Elife 7, e40675 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    54.Ito, T., Teo, T. V., Evans, S. A., Neretti, N. & Sedivy, J. Regulation of cellular senescence by polycomb chromatin modifiers through distinct DNA damage- and histone methylation-dependent pathways. Cell Rep. 22, 3480–3492 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    55.St Aubin, D., Deguise, S., Richard, P., Smith, T. & Geraci, J. Hematology and plasma chemistry as indicators of health and ecological status in beluga whales, Delphinapterus leucas. Arctic 54, 317–331 (2001).56.Norman, S. A. et al. Seasonal hematology and serum chemistry of wild beluga whales (Delphinapterus leucas) in Bristol Bay, Alaska, USA. J. Wildl. Dis. 48, 21–32 (2012).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    57.Frost, K. J. & Suydam, R. S. Subsistence harvest of beluga or white whales (Delphinapterus leucas) in northern and western Alaska 1987–2006. J. Cetacea. Res. Manag. 11, 293–299 (2010).
    Google Scholar 
    58.Rosen, A. D. et al. DNA methylation age is accelerated in alcohol dependence. Transl. Psychiatry 8, 182 (2018).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    59.Zhang, Q. et al. Improved precision of epigenetic clock estimates across tissues and its implication for biological ageing. Genome Med. 11, 54 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    60.Gronniger, E. et al. Aging and chronic sun exposure cause distinct epigenetic changes in human skin. PLoS Genet. 6, e1000971 (2010).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    61.Doi, A. et al. Differential methylation of tissue-and cancer-specific CpG island shores distinguishes human induced pluripotent stem cells, embryonic stem cells and fibroblasts. Nat. Genet. 41, 1350–1353 (2009).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    62.Vandiver, A. R. et al. Age and sun exposure-related widespread genomic blocks of hypomethylation in nonmalignant skin. Genome Biol. 16, 1–15 (2015).CAS 
    Article 

    Google Scholar 
    63.Li, Q. S., Sun, Y. & Wang, T. Epigenome-wide association study of Alzheimer’s disease replicates 22 differentially methylated positions and 30 differentially methylated regions. Clin. Epigenet. 12, 1–14 (2020).Article 
    CAS 

    Google Scholar 
    64.Sun, L., Zhang, X., Wang, T., Chen, M. & Qiao, H. Association of ANK1 variants with new‑onset type 2 diabetes in a Han Chinese population from northeast China. Exp. Ther. Med. 14, 3184–3190 (2017).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    65.Luoma, L. M. & Berry, F. B. Molecular analysis of NPAS3 functional domains and variants. BMC Mol. Biol. 19, 1–19 (2018).Article 
    CAS 

    Google Scholar 
    66.Cosgrove, D. et al. Genes influenced by MEF2C contribute to neurodevelopmental disease via gene expression changes that affect multiple types of cortical excitatory neurons. bioRxiv https://doi.org/10.1101/2019.12.16.877837 (2019).67.Decourcelle, A. et al. O-GlcNAcylation links nutrition to the epigenetic downregulation of UNC5A during colon carcinogenesis. Cancers 12, 3168 (2020).CAS 
    PubMed Central 
    Article 

    Google Scholar 
    68.Yang, T., Zhang, X.-B., Li, X.-N., Sun, M.-Z. & Gao, P.-Z. Homeobox C4 promotes hepatocellular carcinoma progression by the transactivation of Snail. Neoplasma 68, 23–30 (2020).69.Yeung, B., Law, A. & Wong, C. K. Evolution and roles of stanniocalcin. Mol. Cell. Endocrinol. 349, 272–280 (2012).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    70.Chen, C., Jamaluddin, M. S., Yan, S., Sheikh-Hamad, D. & Yao, Q. Human stanniocalcin-1 blocks TNF-α–induced monolayer permeability in human coronary artery endothelial cells. Arterioscler. Thromb. Vasc. Biol. 28, 906–912 (2008).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    71.Jourdain, E. & Karoliussen, R. Identification catalogue of Norwegian killer whales: 2007–2018. Figshare https://doi.org/10.608/m9.figshare.4205226 (2018).72.Kuningas, S., Similä, T. & Hammond, P. S. Population size, survival and reproductive rates of northern Norwegian killer whales (Orcinus orca) in 1986-2003. J. Mar. Biol. Assoc. UK 94, 1277 (2014).Article 

    Google Scholar 
    73.Christensen, I. Growth and reproduction of killer whales, Orcinus orca, in Norwegian coastal waters. Rep. Int. Whal. Commn 6, 253–258 (1984).
    Google Scholar 
    74.Jourdain, E., Vongraven, D., Bisther, A. & Karoliussen, R. First longitudinal study of seal-feeding killer whales (Orcinus orca) in Norwegian coastal waters. PLoS ONE 12, e0180099 (2017).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    75.Arneson, A. et al. A mammalian methylation array for profiling methylation levels at conserved sequences. bioRxiv https://doi.org/10.1101/2021.01.07.425637 (2021).76.Zhou, W., Triche, T. J. Jr., Laird, P. W. & Shen, H. SeSAMe: reducing artifactual detection of DNA methylation by Infinium BeadChips in genomic deletions. Nucleic Acids Res. 46, e123 (2018).PubMed 
    PubMed Central 

    Google Scholar 
    77.Friedman, J., Hastie, T. & Tibshirani, R. Regularization paths for generalized linear models via coordinate descent. J. Stat. Softw. 33, 1 (2010).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    78.Shao, J. Linear model selection by cross-validation. J. Am. Stat. Assoc. 88, 486–494 (1993).Article 

    Google Scholar 
    79.Zhang, P. Model selection via multifold cross validation. Ann. Statist. 21, 299–313 (1993).80.Team, R. C. R.: A language and environment for statistical computing (2020). More

  • in

    Distinct microbial community along the chronic oil pollution continuum of the Persian Gulf converge with oil spill accidents

    Persian Gulf water and sediment samples along the oil pollution continuumWater and sediment samples were collected along the circulation current of the Persian Gulf from Hormuz Island [HW (SAMN12878178) and HS (SAMN12878113)), Asaluyeh area (AW (SAMN12878179) and AS (SAMN12878114)), and Khark Island (KhW (SAMN12878180) and KhS (SAMN12878115)] (Fig. 1). Physicochemical characteristics and Ionic content of the collected samples are presented in the Supplementary Table S1. The GC-FID analyses showed high TPH and polyaromatic hydrocarbon (PAH) concentrations in the Khark sediment (KhS) (Supplementary Table S2). The GC-SimDis analysis showed that C25–C38 HCs were dominant in the KhS (~ 60%), followed by  > C40 HCs (~ 14%) (Supplementary Fig. S1). Chrysene, fluoranthene, naphthalene, benzo(a)anthracene and phenanthrene were respectively the most abundant PAHs in KhS. This pollution could originate from oil spillage due to Island airstrikes during the imposed war (1980–1988), sub-sea pipeline failures, and discharge of oily wastewater or ballast water of oil tankers (ongoing for ~ 50 years)12. The TPH of other water and sediment samples was below the detection limit of our method ( 2%) representatives were enriched in KhS (Fig. 2B). The co-presence of the orders Methanosarcinals, Alteromanadales, and Thermotogae (Petrotogales) in the KhS hints at potential oil reservoir seepage around the sampling site since these taxa are expected to be present in oil reservoirs38. The main HC pollutants in the Asalouyeh are low molecular weight aromatic compounds that mainly influence the prokaryotic population in the water column and rarely precipitate into sediments hence the similarity of AS to HS microbial composition as they both experience low pollution rates.Apart from oil-degrading Proteobacteria (e.g. Alteromonadales, Rhodobacterales, and Oceanospirillales), a diversity of sulfur/ammonia-oxidizing chemolithoautotrophic Proteobacteria were present in these sediments although at lower abundances e.g., (Acidithiobacillales (KhS 1.8%), Chromatiales (HS 1.5, AS 1.1, KhS 0.85%), Ectothiorhodospirales (HS 3.75, AS 2.3, KhS 1.7%), Halothiobacillales (KhS 2.6%), Thiotrichales (HS 1.5, AS 1.1, KhS 0.3%), Thiohalorhabdales (HS 0.7, AS 1.2, KhS 0.5%), Thiomicrospirales (KhS 1.5%)) (Fig. 2B).Sulfate-reducing bacteria (SRB) in HS comprised up to 16.2% of the community (Desulfobacterales, NB1-j, Myxococcales, Syntrophobacterales, and Thermodesulfovibrionia). Similar groups along with Desulfarculales, comprised the SRB functional guild of the AS (~ 18.9%). In comparison, Desulfuromonadales and Desulfobacterales were the SRB representatives in KhS with a total abundance of only ~ 3.3%. The lower phylogenetic diversity and community contribution of SRBs in KhS hint at the potential susceptibility of some SRBs to oil pollution or that HC degraders might outcompete them (e.g., Deferribacterales). Additionally, KhS was gravel-sized sediment (particles ≥ 4 mm diameter), whereas HS and AS samples were silt and sand-sized sediments39. The higher oxygen penetration in gravel particles of KhS hampers anaerobic metabolism of sulfate/nitrate-reducing bacteria hence their lower relative abundance in this sample (Fig. 2B).Whereas in water samples, sulfur/ammonia-oxidizing chemolithoautotrophs such as Thiomicrospirales and sulfate/nitrate-reducing bacteria such as Desulfobacterales, NB1-j, Deferribacterales, Anaerolineales, Nitrosococcales, Nitrosopumilales, and Pirellulales were present in very small quantities (lower than 0.5% in each sample).Chronic exposure to oil pollution shapes similar prokaryotic communities as oil spill eventsWe analyzed the prokaryotic community composition of 41 oil-polluted marine water metagenomes (different depths in the water column) from Norway (Trondheimsfjord), Deepwater Horizon (Gulf of Mexico), the northern part of the Gulf of Mexico (dead zone) and Coal Oil Point of Santa Barbara; together with 65 oil exposed marine sediment metagenomes (beach sand, surface sediments and deep-sea sediments) originating from DWH Sediment (Barataria Bay), Municipal Pensacola Beach (USA) and a hydrothermal vent in Guaymas Basin (Gulf of California) in comparison with the PG water and sediment samples (in total 112 datasets) (Supplementary Table S3). This extensive analysis allowed us to get a comparative overview of the impact of chronic oil pollution on the prokaryotic community composition.Hydrocarbonoclastic bacteria affiliated to Oceanospirillales, Cellvibrionales (Porticoccaceae family), and Alteromonadales40 comprised a significant proportion of the prokaryotic community in samples with higher aliphatic compounds pollution e.g. DWHW.BD3 (sampled six days after the incubation of unpolluted water with Macondo oil), DWHW.he1, and DWHW.he2 (oil-polluted water samples incubated with hexadecane), DWHW.BM1, DWHW.BM2, DWHW.OV1 and DWHW.OV2 (sampled immediately after the oil spill in the Gulf of Mexico) (Fig. 3). Samples treated with Macondo oil, hexadecane, naphthalene, phenanthrene, and those taken immediately after the oil spill in the Gulf of Mexico had a significantly lower proportion of SAR11 due to the dominance of bloom formers and potential susceptibility of SAR11 to oil pollutants (Fig. 3).Figure 3The abundance of unassembled 16S rDNA reads from unassembled metagenomes of different oil-polluted water samples (41). Row names are microbial taxa at the order level. For taxa with lower frequency, the higher taxonomic level is shown (47 taxa in total). The right-hand dendrogram represents the clustering of rows based on the Pearson correlation. Columns are the name of water samples. Samples are clustered based on Pearson correlation and the color scale on the top left represents the row Z-score. Figure was plotted using “circlize” and “ComplexHeatmap” packages in R.Full size imageFlavobacteriales and Rhodobacterales were present in relatively high abundance in almost all oil-polluted water samples except for those with recent pollution. Samples named NTW5, NTW6, NTW11, NTW12, which were incubated with MC252 oil for 32–64 days, represented similar prokaryotic composition dominating taxa that are reportedly involved in degrading recalcitrant compounds like PAHs in the middle-to-late stages of the oil degradation process (Alteromonadales, Cellvibrionales, Flavobacteriales, and Rhodobacterales). Whereas at the earlier contamination stages, samples represented a different community composition with a higher relative abundance of Oceanospirillales (e.g., NTW8, NTW9, NTW15, NTW16, and NTW17 sampled after 0–8 days incubation) (Fig. 3).The non-metric multidimensional analysis of the prokaryotic community of 106 oil-polluted water and sediment samples, together with the PG samples, is represented in Fig. 4. Water and sediment samples expectedly represented distinct community compositions. The AW sample was placed near samples treated with phenanthrene and naphthalene in the NMDS plot showing the impact of aromatic compounds on its microbial community. The KhW sample was located near NTW13 in the plot, both of which had experienced recent oil pollution.Figure 4Non-metric multidimensional scaling (NMDS) of the Persian Gulf water and sediment metagenomes along with oil-polluted marine water and sediment metagenomes based on Bray–Curtis dissimilarity of the abundance of 16S rDNA reads in unassembled metagenomes at the order level. Samples with different geographical locations are shown in different colors. PG water and sediment samples are shown in red. Water and sediment samples are displayed by triangle and square shapes, respectively. Figure was plotted using “vegan” library in R.Full size imageThe orders Oceanospirillales, Alteromonadales, and Pseudomonadales were present in relatively high abundances in all oil-polluted water samples except for HW (PG input water) and samples collected from the northern Gulf of Mexico dead zone (GOMDZ) (Fig. 3). Persian Gulf was located in the proximity of the developing oxygen minimum zone (OMZ) of the Arabian Sea that is slowly expanding towards the Gulf of Oman41. Potential water exchange with OMZ areas could be the cause of higher similarity to the GOMDZ microbial community42.Our results suggest that water samples with similar contaminants and exposure time to oil pollution enrich for similar phylogenetic diversity in their prokaryotic communities (Fig. 3). Marine prokaryotes represent vertical stratification with discrete community composition across the depth profile. According to our analyses, the prokaryotic communities of the oil-polluted areas are consistently dominated by similar taxa regardless of sampling depth or geographical location. We speculate that the high nutrient input due to crude oil intrusion into the water presumably disturbs this stratification and HC degrading microorganisms are recruited to the polluted sites where their populations flourish.The inherent heterogeneity of the sediment prokaryotic communities is retained even after exposure to oil pollution, reflected in their higher alpha diversity (Supplementary Fig. S3). However, similar taxa dominate the community in response to oil pollution (Fig. 5).Figure 5The abundance of unassembled 16S rDNA reads from unassembled metagenomes of different oil-polluted sediment samples (65). Row names are microbial taxa at the order level. For taxa with lower frequency, the higher taxonomic level is shown (77 taxa in total). The right-hand dendrogram represents the clustering of rows based on the Pearson correlation. Columns are the name of sediment samples. Samples are clustered based on Pearson correlation and the color scale on the top left represents the row Z-score. Figure was plotted using “circlize” and “ComplexHeatmap” packages in R.Full size imageIn sediment samples, Deltaproteobacteria had the highest abundance, followed by Gammaproteobacteria representatives. Ectothiorhodospirales, Rhizobiales, Desulfobacterales, Myxococcales, and Betaproteobacteriales representatives were present in almost all samples at relatively high quantities (Fig. 5). Sulfate/nitrate-reducing bacteria were major HC degraders in sediment, showing substrate specificity for anaerobic HC degradation43. Desulfobacterales and Myxococcales were ubiquitous sulfate-reducers, present in almost all oil-polluted sediment samples44. Sulfate-reducing Deltaproteobacteria play a key role in anaerobic PAH degradation, especially in sediments containing recalcitrant HC types45. Members of Rhizobiales are involved in nitrogen fixation, which accelerates the HC removal process in the sediment samples46, and therefore their abundance increase in response to oil pollution (Fig. 5).Prokaryotes involved in nitrogen/sulfur cycling of sediments are defined by factors such as trace element composition, temperature, pressure, and more importantly, depth and oxygen availability. In oil-polluted sediment samples, the simultaneous reduction of available oxygen with an accumulation of recalcitrant HCs along the depth profile complicates the organic matter removal. However, anaerobic sulfate-reducing HC degrading bacteria will cope with this complexity47. Prokaryotic communities of HS and AS samples represented similar phylogenetic diversity (Figs. 4, 5). Their prokaryotic community involved in the nitrogen and sulfur cycling resembles the community of DWHS samples. The KhS sample had a similar prokaryotic community to deeper sediment samples collected from 30 to 40 cm depth (USFS3, USFS11, and USFS12) which could be due to our sampling method using a grab sampling device.Our results show that the polluted sediments’ sampling depth (surface or subsurface) defines the dominant microbial populations. Hydrocarbon degrading microbes had the ubiquitous distribution in almost all oil-polluted water and sediment samples including Oceanospirillales, Cellvibrionales, Alteromonadales, Flavobacteriales, Pseudomonadales, and Rhodobacterales. Mentioned orders along with Ectothiorhodospirales, Rhizobiales, Desulfobacterales, Myxococcales, and Betaproteobacteriales and also representatives of Deltaproteobacteria phylum dominated in sediment samples. However, their order of frequency varies depending on the type of oil pollution present at the sampling location and the exposure time.Genome-resolved metabolic analysis of the Persian Gulf’s prokaryotic community along the pollution continuumA total of 82 metagenome-assembled genomes (MAGs) were reconstructed from six sequenced metagenomes of the PG (completeness ≥ 40% and contamination ≤ 5%). Amongst them, eight MAGs belonged to domain Archaea and 74 to domain bacteria. According to GTDB-tk assigned taxonomy (release89) (https://data.gtdb.ecogenomic.org/releases/release89/), reconstructed MAGs were affiliated to Gammaproteobacteria (36.6%), Alphaproteobacteria (12.2%), Flavobacteriaceae (9.7%), Thermoplasmatota (5%) together with some representatives of other phyla (MAG stats in Supplementary Table S4).A collection of reported enzymes involved in the degradation of different aromatic and aliphatic HCs under both aerobic and anaerobic conditions was surveyed in the annotated MAGs of this study43,48,49,50. The KEGG orthologous accession numbers (KOs) of genes involved in HC degradation were collected, and the distribution of KEGG orthologues detected at least in one MAG (n = 76 genes) is represented in Fig. 6.Figure 6Hydrocarbon degrading enzymes present in recovered MAGs from the PG water and sediment metagenomes. Row names represent the taxonomy of recovered MAGs and their completeness is provided as a bar plot on the right side. The color indicates the MAG origin. The size of dots indicates the presence or absence of each enzyme in each recovered MAG. Columns indicate the type of hydrocarbon and in the parenthesis is the name of the enzyme hydrolyzing this compound followed by its corresponding KEGG orthologous accession number. Figure was plotted using “reshape2” and “ggplot2” packages in R.Full size imageA combination of different enzymes runs the oil degradation process. Mono- or dioxygenases are the main enzymes triggering the HC degradation process under aerobic conditions. Under anaerobic conditions, degradation is mainly started by the addition of fumarate or in some cases, by carboxylation of the substrate. Therefore, bacteria containing these genes will potentially initiate the degradation process that will be continued by other heterotrophs. Enzymes such as decarboxylase, hydroxylase, dehydrogenase, hydratase, and isomerases act on the products of initiating enzymes mentioned above through a series of oxidation/reduction reactions.Various microorganisms cooperate to cleave HCs into simpler compounds that could enter common metabolic pathways. Mono- or dioxygenases which are involved in the degradation of alkane (alkane 1-monooxygenase, alkB/alkM), cyclododecane (cyclododecanone monooxygenase, cddA), Biphenyl (Biphenyl 2, 3-dioxygenase subunit alpha/beta, bphA1/A2, Biphenyl-2, 3-diol 1, 2-dioxygenase, bphC), phenol (phenol 2-monooxygenase, pheA), toluene (benzene 1, 2-dioxygenase subunit alpha/beta todC1/C2, hydroxylase component of toluene-4-monooxygenase, todE), xylene (toluate/benzoate 1,2-dioxygenase subunit alpha/beta/electron transport component, xylX/Y/Z, hydroxylase component of xylene monooxygenase, xylM) and naphthalene/phenanthrene (catechol 1,2 dioxygenase, catA, a shared enzyme between naphthalene/phenanthrene /phenol degradation) were detected in recovered MAGs of the PG.The key enzymes including Alkylsuccinate synthase (I)/(II) (assA1/A2), benzylsuccinate synthase (BssA)/benzoyl-CoA reductase (BcrA), ethylbenzene dehydrogenase (EbdA), and 6-oxo-cyclohex-1-ene-carbonyl-CoA hydrolase (BamA) that are responsible for initiating the degradation of alkane, toluene, ethylbenzene and benzoate exclusively under anaerobic conditions were not detected in reconstructed MAGs of this study. Consequently, recovered MAGs of this study are not initiating anaerobic degradation via known pathways while they have the necessary genes to continue the degradation process started by other microorganisms.The MAG KhS_63 affiliated to Immundisolibacter contained various types of mono- or dioxygenases and had the potential to degrade a diverse range of HCs such as alkane, cyclododecane, toluene, and xylene (Fig. 6). Members of this genus have been reported to degrade high molecular weight PAHs51.Lutimaribacter representatives have been isolated from seawater and reported to be capable of degrading cyclohexylacetate52. We also detected enzymes responsible for alkane, cycloalkane (even monooxygenase enzymes), and naphthalene degradation under aerobic conditions and alkane, ethylbenzene, toluene, and naphthalene degradation under anaerobic conditions in KhS_39 affiliated to this genus (Fig. 6).The MAGs KhS_15 and KhS_26 affiliated to Roseovarius had the enzymes for degrading alkane (alkane monooxygenase, aldehyde dehydrogenase), cycloalkane, naphthalene, and phenanthrene under aerobic and toluene and naphthalene under anaerobic condition. PAHs degradation has been reported for other representatives of this taxa as well53.The MAGs KhS_11 (a representative of Rhodobacteraceae) and KhS_53 (Marinobacter) had alkB/alkM, KhS_27 (GCA-2701845), KhS_29 (UBA5862) and KhS_40 (from Porticoccaceae family) had cddA, KhS_13 and KhS_21 (UBA5335) and KhS_38 (Oleibacter) had both alkB/alkM and xylM genes. They were among microbes that were initiating the degradation of alkane, cycloalkane and xylene compounds. Other MAGs recovered from Khark sediment were involved in the continuation of the degradation pathway. For example, KhS_1 was affiliated to the genus Halomonas and had different enzymes to degrade intermediate compounds. Halomonas representatives have been frequently isolated from oil-polluted environments54. The phylum Krumholzibacteria has been first introduced in 2019 and reported to contain heterotrophic nitrite reducers55. Two MAGs, KhS_5 and KhS_10, were affiliated to this phylum and contained enzymes involved in the anaerobic degradation of toluene, phenol, and naphthalene (Fig. 6).The MAGs KhS_12 and KhW_31 affiliated to the genus Flexistipes, in Deferribacterales order, were reconstructed from both KhW and KhS samples. Deferribacterales are reported to be present in the medium to high-temperature oil reservoirs with HC degradation activity and also in high-temperature oil-degrading consortia56. The type strain of this species was isolated from environments with a minimum salinity of 3% and a temperature of 45–50 °C57. The presence of this genus in KhS could be due to natural oil seepage from the seabed as PG reservoirs mainly have medium to high temperature and high salinity. Enzymes involved in the degradation of alkane, phenol, toluene and naphthalene under anaerobic conditions were present in MAGs KhS_12 and KhW_31.As mentioned earlier, Flavobacteriales are potent marine indigenous HC degraders that bloom in response to oil pollution58. Flavobacteriales affiliated MAGs (KhW_2, KhW_3, AW_21, and AW_33) were recovered from KhW and AW and mostly contained enzymes that participate in the degradation of aromatic compounds under anaerobic conditions. KhW_2 and KhW_3 also had both alkB/M (alkane monooxygenase) and xylM enzyme, which initiates the alkane and xylene bioremediation in Khark water. Among other recovered MAGs from KhW sample, KhW_18 (UBA724), KhW_24 (clade SAR86), KhW_43 (UBA3478) had alkB/M, and xylM, KhW_24 (clade SAR86) had alkB/M and cddA, and KhW_28 (from Rhodobacteraceae family) had alkB/M and pheA genes in their genome to initiate the degradation process (Fig. 6).Marinobacter (KhW_15) was another MAG reconstructed from KhW sample. This genus is one of the main cultivable genera that play a crucial role in the bioremediation of a wide range of oil derivatives in polluted marine ecosystems54.Marine Group II (MGII) and Poseidonia representatives of Thermoplasmatota that have been reported to be nitrate-reducing Archaea59, were recovered from AW sample (AW_40, AW_45) and contained several enzymes contributing in alkane (alkane monooxygenase, aldehyde dehydrogenase) and naphthalene/phenanthrene/phenol/xylene degradation (decarboxylase) under aerobic conditions. The HC degradation potential of representatives of this phylum has been previously reported60.In the Asalouyeh water sample, MAGs AW_25 (UBA4421) and AW_38 (UBA8337) had cddA, AW_21 (UBA8444) had catA, AW_11 (Poseidonia) and AW_17 (from Rhizobiales order) had both alkB/M and xylM, and AW_4 (UBA8337) had catA and pheA genes and had potential to trigger the breakdown of their corresponding oil derivatives.Other recovered genomes had the potential to metabolize the product of initiating enzymes. For instance, AW_23 contained enzymes involved in the degradation of naphthalene, phenol and cyclododecane and was affiliated to the genus Alteromonas (Fig. 6).Three recovered MAGs of HW affiliated to Pseudomonadales (HW_23), Poseidoniales (HW_24), and Flavobacteriales (HW_30) contained some initiating enzymes to degrade cyclododecane/biphenyl/toluene, alkane/xylene, and alkane/xylene/naphthalene/phenanthrene, respectively. A representative of Heimdallarchaeia that are mainly recovered from sediment samples was reconstructed from the Hormuz water sample (HW_28). It had a completeness of 81% and contained enzymes involved in anaerobic degradation of alkanes. This archaeon could potentially be an input from the neighboring OMZ as this phyla include representatives adopted to microoxic niches61. Containing genes with the potential to initiate the oil derivative degradation in the input water with no oil exposure reiterates the intrinsic ability of marine microbiota for HC degradations and oil bioremediation.While 16S rRNA provides an overview of the community, MAGs provide the possibility to inspect the metabolic capability of the microbiota. We decided to provide both in this manuscript as we believe they are complementary. Having the full picture provided by the combination of these analyses allows for a better understanding of the community structure and their metabolic capabilities. This is even more evident for sediment samples as they are highly diverse, and reconstructing MAGs from sediment metagenomes is still a bottlenecks. In this case, we rely more on the 16S rRNA to provide an overall view of the community composition.This said, we see similar taxonomic distribution in the MAGs and 16S rRNA e.g., the prevalence of Flavobacteriales and Rhodobacterales in KhW and KhS, Synechococcales, and Desulfobacteriales and Flavobacteriales in HW, HS and AW samples, respectively.Additionally, some rare microbiota representatives were recovered among reconstructed MAGs. For example, the Immundisolibacterales showed an abundance of only 0.8% in the KhS sample based on 16S rRNA but the recovered KhS_63 MAG was affiliated to this taxon. Notably, this MAG contained many genes involved in hydrocarbon degradation having the highest potential in hydrocarbon degradation. More

  • in

    Design of synthetic human gut microbiome assembly and butyrate production

    Model-guided procedure guides the exploration of butyrate production landscapesWe aimed to explore the butyrate production landscape as a function of community composition to decipher microbial interactions shaping butyrate production. Exploring the butyrate production functional landscape is a major challenge because the number of sub-communities increases exponentially with the number of species43. To investigate the landscape, we developed a modeling framework to guide the iterative design of informative experiments (Fig. 1a, b). Microbial interactions can impact growth or metabolite production by influencing the availability of ecological niches or facilitating metabolite degradation. To capture these two types of interactions, we implemented a two-stage modeling framework to determine the contributions of microbial interactions to species growth and community assembly or metabolite production. In the first stage, a dynamic ecological model, referred to as the generalized Lotka–Volterra model (gLV), predicts community assembly. The second stage predicts metabolite production as a function of the resulting community composition (Fig. 1b). The gLV model is a set of coupled ordinary differential equations that capture the temporal change in species abundances due to monospecies growth parameters and inter-species growth interactions (see the “Methods” section)16. To estimate parameters for the gLV model, we use Bayesian parameter inference techniques to determine the uncertainty in our parameters based on biological and technical variability in the experimental data44.Fig. 1: Iterative modeling framework to predict microbial community assembly and metabolic function.a Two-stage modeling framework for predicting community assembly and function. The generalized Lotka–Volterra model (gLV) represents community dynamics. A Bayesian Inference approach was used to determine parameter uncertainties due to biological and technical variability. A linear regression model with interactions maps assembled community composition to metabolite concentration. Combining these two models enables prediction of a probability distribution of metabolite concentration from initial species abundances. b Design–Test–Learn cycle for model development. First, we use our model to explore the design space of possible experiments (i.e. different initial conditions of species presence/absence) and design communities that span a desired range of metabolite concentrations. Next, we use high-throughput experiments to measure species abundance and metabolite concentration. Finally, we evaluate the model’s predictive capability and infer an updated set of parameters based on the new experimental measurements. c Phylogenetic tree of the synthetic human gut microbiome composed of 25 highly prevalent and diverse species. Branch color indicates phylum and underlined species denote butyrate producers.Full size imageOur metabolite production model consists of a linear regression model with interaction terms mapping community composition (i.e. abundance of each species) at a specific time point to the concentration of an output metabolite at that time. This model was based on a phenomenological model of metabolite production used in bioprocess engineering expanded to microbial communities (see the “Methods” section). In the regression model, the first-order terms capture the monospecies production per unit biomass and the interaction terms represent the impact of inter-species interactions on metabolite production per unit biomass (i.e. deviations from constant metabolite production per unit biomass19). To estimate parameters for the regression model, we use Lasso regression to identify the most impactful interactions. Altogether, the composite gLV and regression model predicts the probability distribution of the metabolite concentration given an initial condition of species abundances (Fig. 1b, see the “Methods” section).In metabolic and protein engineering, a design–test–learn cycle (DTL) has been used to design biomolecules45 or metabolic pathways46 with properties that satisfy desired performance specifications. We hypothesized that this engineering-inspired approach could be used to explore community design spaces and understand the composition–function mapping for butyrate production. Each cycle consisted of: (1) a design phase wherein we used our model informed by experimental observations to simulate a vast number of potential community compositions to identify sub-communities that satisfied biological objectives (i.e. desired butyrate concentrations), (2) a test phase wherein the selected sub-communities were assembled and species abundance and butyrate concentration were measured, and (3) a learn phase wherein patterns in our experimental data were used to estimate model parameters and to extract information about the key microbial interactions influencing community assembly and butyrate production.Two-stage model enables efficient exploration of low richness community design spaceTo develop a system of microbes representing major metabolic functions in the gut, we selected 25 prevalent bacterial species from all major phyla in the human gut microbiome47 (Fig. 1c, Supplementary Data 1). This community contained five butyrate-producing Firmicutes which have been shown to play important roles in human health and protection from diseases (Fig. 1c, Supplementary Data 1). Due to the lack of a defined medium that universally supports the growth of gut microbes, most in vitro studies use undefined media, making it difficult to interrogate the effects of unknown components on community behaviors48. To maximize our knowledge of the substrates available to the communities, we developed a chemically defined medium to grow the synthetic communities (see the “Methods” section).Based on previous studies using pairwise communities to predict higher richness community behaviors16,18,49, we hypothesized that training our model on single and pairwise community measurements would provide an informative starting point for mapping composition–function relationships determining butyrate production. To do so, we first measured time-resolved growth of single species and observed a wide variety of growth dynamics within each phylum, including disparate growth rates and carrying capacities (Supplementary Fig. 1). We assembled each pairwise community containing at least one butyrate producer (the focal species of our system50) and measured species abundance and the concentrations of organic acid fermentation products (including butyrate, lactate, succinate, and acetate) after 48 h. The pairwise consortia displayed a broad range of butyrate concentrations of 0–50 mM (Fig. 2a).Fig. 2: Exploring the predicted butyrate production of 3–5 member communities with a model trained on 1–2 species communities.a Categorical scatter plot of butyrate production in 1–2 species and 24–25 species communities. Solid datapoints indicate the mean of the biological replicates which are represented by transparent datapoints connected to the mean with transparent lines. The colors indicate which butyrate producer was present in the community with green indicating the presence of multiple butyrate producers. DP− and AC− indicate the 24-species communities lacking Desulfovibrio piger (DP) and Anaerostipes caccae (AC), respectively. b Predicted medians (black line) and 60 percent confidence intervals (gray bars) of butyrate concentration for all 3–5 member communities containing at least one butyrate producer (46,591 community predictions). Colored lines indicate median and 60 percent confidence interval of butyrate production of communities chosen for the experimental design with the color indicating the number of species in the community (156 communities). Subplots separate groups of communities based on the identities of the combination of butyrate producers specified. c Scatter plot of measured butyrate versus predicted butyrate for 3–5 species communities. Colors indicate which butyrate producer was present in the community as in a. Biological replicates (n = 1–5, depending on the community, exact values in source data) are indicated by transparent squares connected to the corresponding mean, which is represented by the large data point. Prediction error bars (x-axis) indicate the 60% confidence interval of the predicted butyrate distribution as in b, with the center being the median prediction. Dashed line indicates the linear regression between the mean measured butyrate and the median predicted butyrate. Indicated statistics are for Pearson correlation (two-sided). Source data are available in the Source Data file.Full size imageSingle-species deletion communities have been used to investigate the contributions of individual species to a community function13,16. Therefore, we characterized the full 25-species community and each single-species deletion sub-community (i.e. 24-member consortia). In stark contrast to the pairwise communities, the 24- and 25-species communities exhibited similar low butyrate production (~2–22 mM Butyrate). The absence of only two species Desulfovibrio piger (DP) (~22 mM Butyrate) and Anaerostipes caccae (AC) (~2 mM Butyrate) resulted in a significant increase or decrease in butyrate concentration compared to the remaining 24-member and 25-member communities (Fig. 2a, Supplementary Fig. 2a). In addition, the concentrations of all measured organic acids spanned a much smaller range in the 24 and 25-member communities than the single and pairwise consortia (Supplementary Fig. 2b). These results suggest that high richness communities may trend towards a similar low butyrate-producing state that is difficult to change by the deletion of most single species and motivates a model-guided design strategy for exploring how community richness shapes butyrate production.To determine whether individual and pairwise communities could predict community composition and butyrate production of low richness communities (i.e. 3–5 species), we estimated the parameters of our model based on experimental measurements. Our initial model was informed only by pairwise communities that contained at least one butyrate producer (Supplementary Data 2, M1) and was thus naïve to all interactions between non-butyrate producers. We assumed that the unobserved growth interactions could be predicted based on trends in measured interactions across phylogenetic relatedness (see the “Methods” section)16. However, the resulting model was unable to predict butyrate production in the 24-and 25-member communities (Supplementary Fig. 3), which we attributed to missing information about non-butyrate producer interactions in our training data. Thus, we used our model to explore a low richness design space of 3–5 species communities based on the assumption that pairwise interactions would be more observable in low than high richness (i.e. >10 species) communities to identify an improved parameter estimate for non-butyrate producer interactions.We used our initial M1 model to predict the probability distributions of butyrate production for all 3–5 species communities containing at least one butyrate producer (46,591 communities). The predicted butyrate production varied substantially based on the combination of butyrate producers present in each community (Fig. 2b). In addition, we observed variations in the shapes of the probability distributions based on how the uncertainty in growth prediction propagated through the regression model. For instance, the butyrate concentration in the AC, Roseburia intestinalis (RI) pairwise community was lower than the AC monoculture, even though RI was low abundance, resulting in a high magnitude negative parameter in the regression model for a production interaction between AC and RI (Supplementary Data 3). Due to the uncertainty in the growth parameters, the model predicted that RI would grow substantially in a subset of the 3–5 member simulations containing both AC and RI. The variability in predicted RI growth combined with the high magnitude negative interaction parameter between AC and RI resulted in distributions where the median butyrate concentration was high (i.e. for simulations where RI did not grow substantially), and the 60 percent confidence interval extended to 0 mM butyrate (i.e. when RI grew substantially) (Fig. 2b). In sum, these results demonstrate that the shape of the predicted probability distributions can provide information about the uncertainty in species growth based on experimental observations.Based on the simulations, we selected 156 communities that spanned a broad range of predicted butyrate concentrations across the butyrate producer groups to evaluate experimentally (Fig. 2b). The model prediction exhibited good agreement with the rank order of butyrate production (Spearman rho = 0.84, p = 9*10−43) (Fig. 2c) and species abundance (Spearman rho = 0.76, p = 3*10−122) (Supplementary Fig. 4a–d), demonstrating that our initial model could predict a wide range of butyrate production in low richness communities.Composition–function landscape predicts contributions of growth and production interactionsEncouraged by our model’s predictive ability, we sought to explore composition–function relationships in higher richness communities (i.e. >10 species) using a model with updated parameters based on measurements of the 3–5 member communities (Supplementary Data 2, M2). Since the human gut microbiome exhibits functional redundancy in butyrate pathways51, we first used model M2 to simulate the assembly of all communities containing all five butyrate producers (5-butyrate producer or 5BP, 1,048,576 total) to map the composition–function landscape for butyrate production (Fig. 3a). In addition, we simulated the assembly of all communities containing the four butyrate producers excluding AC (4-butyrate producer or 4BP, 1,048,576 total) to understand how the composition–function landscape changes in the absence of the most productive butyrate producer (Fig. 3b). The majority of 5BP communities were predicted to have higher butyrate concentration than any of the 4BP communities (Fig. 3a, b), consistent with the substantial decrease in butyrate in the AC deletion community observed previously (Fig. 2a).Fig. 3: Community composition–function landscapes reveal key role of production interactions on A. caccae and negative impact of D. piger on butyrate production.a Scatter plot of predicted total butyrate producer abundance versus predicted butyrate concentration for all possible communities in which all five butyrate producers are present (1,048,576 communities). Histograms indicate the butyrate concentration distribution across the given axis. Communities are colored according to the presence (red) or absence (blue) of D. piger (DP). Blue and red dashed lines indicate the linear regression of communities with (red, y = −2.3x + 25.8, r = −0.34) or without (blue, y = 3.1x + 28.0, r = 0.76) DP. The white star indicates the full 25-member community and black star indicates the community of five butyrate producers alone. Large data points indicate communities chosen for experimental validation. Black triangles indicate leave-one-out communities, black circles indicate designed communities, and gray squares indicate random communities, with open/closed symbols indicate the absence/presence of DP. b Scatter plot of predicted total butyrate producer abundance versus predicted butyrate concentration for all possible communities in which all four butyrate producers excluding AC are present (1,048,576 communities). Histograms indicate the butyrate concentration distribution. Gray dashed line indicates the mean predicted butyrate concentration across all communities. Black dashed line indicates the linear regression of all communities (y = 13.2x−0.1, r = 0.64). The white star indicates the full 24-member community and the black star indicates the four butyrate producers alone. Large data points indicate communities chosen for experimental validation. c, d Scatter plots of experimental measurements of total butyrate producer abundance versus butyrate concentration for communities with (c) and without (d) AC. Data point shapes correspond to the legends in (a) and (b) and represent the mean of biological replicates, which are shown as small datapoints connected to the corresponding mean with lines (n = 2 except for 5 BP, 24 and 25 species communities where n = 5–8). Dashed line in (d) indicates the linear regression (y = 8.9x−0.5). e Comparison of butyrate concentration in communities from a and b with and without DP for both the designed and random experimental sets for the 5BP communities and designed set for the 4BP communities. Data point shapes correspond to the legends in a and b and represent the mean of biological replicates, which are shown as small datapoints connected to the corresponding mean with lines. Box and whisker plots represent the median (center line), quartiles (box), and range (whiskers) of the mean butyrate concentration for each community, excluding outliers (points outside 1.5 times the interquartile range). Indicated p-values are from a Mann–Whitney U test (5BP Designed: n = 28 for DP+ and n = 54 for DP−; 5BP Random: n = 27 for DP+ and n = 55 for DP−; 4BP: n = 42 for DP+ and n = 42 for DP−). f Butyrate concentration per unit biomass as a function of sulfide concentration after 24 h of growth. Butyrate concentration per biomass was normalized to the no sulfide condition. Circles indicate the mean of biological replicates, with individual replicates shown as transparent squares (n = 4). Inset: Butyrate concentration per biomass (mM OD600−1) for AC with and without the addition of 1.6 mM sulfide across time (n = 3). Endpoint sulfide concentrations were higher in the data shown in the inset than in the main figure (Supplementary Fig. 6). Source data are available in the Source Data file.Full size imageThe relationship between butyrate producer abundance and butyrate can provide insight into the contributions of growth and production interactions in the presence and absence of AC (Fig. 3a, b). If butyrate producer abundance correlates with butyrate, then growth interactions drive butyrate production, whereas the contributions of production interactions would reduce the strength of this correlation. The 5BP communities were predicted to have a large contribution of production interactions as evidenced by a weak correlation between butyrate concentration and butyrate producer abundance (Spearman rho = 0.17, p  More

  • in

    A natural constant predicts survival to maximum age

    1.Bailey, D. L., Humm, J. L., Todd-Pokropek, A. & van Aswegen, A. Nuclear Medicine Physics: A Handbook for Teachers and Students. International Atomic Energy Agency (International Atomic Energy Agency, 2014).2.McGraw-Hill. McGraw-Hill encyclopedia of science & technology. (McGraw-Hill, 2007).3.Medawar, P. B. An unsolved problem of biology. in The uniqueness of the individual (ed. Medawar, P. B.) 44–70 (Basic Books, Inc., 1952).4.Leike, A. Demonstration of the exponential decay law using beer froth. Eur. J. Phys. 23, 21–26 (2002).Article 

    Google Scholar 
    5.Pauly, D. On the interrelationships between natural mortality, growth parameters, and mean environmental temperature in 175 fish stocks. ICES J. Mar. Sci. 39, 175–192 (1980).Article 

    Google Scholar 
    6.Vetter, E. F. Estimation of natural mortality in fish stocks: a review. Fish. Bull. 86, 25–43 (1988).
    Google Scholar 
    7.Gosselin, J., Zedrosser, A., Swenson, J. E. & Pelletier, F. The relative importance of direct and indirect effects of hunting mortality on the population dynamics of brown bears. Proc. R. Soc. B Biol. Sci. 282, 1–9 (2015).8.Nowak, D. J., Kuroda, M. & Crane, D. E. Tree mortality rates and tree population projections in Baltimore, Maryland, USA. Urban . Urban Green. 2, 139–147 (2004).Article 

    Google Scholar 
    9.Hoenig, J. M. et al. The logic of comparative life history studies for estimating key parameters, with a focus on natural mortality rate. ICES J. Mar. Sci. 73, 2453–2467 (2016).Article 

    Google Scholar 
    10.Krebs, C. J. Ecology: The Experimental Analysis of Distribution and Abundance. (Pearson Education Limited, 2014).11.Myers, R. A., Bowen, K. G. & Barrowman, N. J. Maximum reproductive rate of fish at low population sizes. Can. J. Fish. Aquat. Sci. 56, 2404–2419 (1999).
    Google Scholar 
    12.Simpfendorfer, C. A., Bonfil, R. & Latour, R. J. Mortality estimation. in. FAO Fish. Tech. Pap. 474, 127 (2005).
    Google Scholar 
    13.Cortés, E. Perspectives on the intrinsic rate of population growth. Methods Ecol. Evol. 7, 1136–1145 (2016).Article 

    Google Scholar 
    14.IUCN. Guidelines for Using the IUCN Red List Categories and Criteria. Version 14. Prepared by the Standards and Petitions Committee. Geographical 14, 1–113 (2019).15.Myers, R. A. & Worm, B. Extinction, survival or recovery of large predatory fishes. Philos. Trans. R. Soc. B Biol. Sci. 360, 13–20 (2005).Article 

    Google Scholar 
    16.Conde, D. A. et al. Data gaps and opportunities for comparative and conservation biology. Proc. Natl. Acad. Sci. U. S. A. 116, 9658–9664 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    17.Gavrilov, L. & Gavrilova, N. The biology of life span: a quantitative approach. (Harwood Academic Publishers, 1991).18.Sekharan, K. Estimates of the stocks of oil sardine and mackerel in the present fishing grounds off the West coast of India. Indian J. Fish. 21, 177–182 (1974).
    Google Scholar 
    19.Alagaraja, K. Simple methods for estimation of parameters for assessing exploited fish stocks. Indian J. Fish. 31, 177–208 (1984).
    Google Scholar 
    20.Cadima, E. L. Fish stock assessment manual. FAO Fish. Tech. Pap. 393, 161 (2003).
    Google Scholar 
    21.Hewitt, D. A. & Hoenig, J. M. Comparison of two approaches for estimating natural mortality based on longevity. Fish. Bull. 103, 433–437 (2005).
    Google Scholar 
    22.Dureuil, M. et al. Unified natural mortality estimation for teleosts and elasmobranchs. Mar. Ecol. Prog. Ser. https://doi.org/10.3354/meps13704 (accepted).23.Litzgus, J. D. Sex differences in longevity in the spotted turtle (Clemmys guttata). Copeia 2, 281–288 (2006).Article 

    Google Scholar 
    24.Calder, W. A. III Body size, mortality, and longevity. J. Theor. Biol. 102, 135–144 (1983).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    25.Botkin, D. B., Janak, J. F. & Wallis, J. R. Some ecological consequences of a computer model of forest growth. J. Ecol. 60, 849–872 (1972).Article 

    Google Scholar 
    26.Holt, S. J. A note on the relation between the mortality rate and the duration of life in an exploited fish population. Int. Comm. Northwest Atl. Fish. Res. Bull. 2, 73–75 (1965).
    Google Scholar 
    27.Hoenig, J. M. Should natural mortality estimators based on maximum age also consider sample size? Trans. Am. Fish. Soc. 146, 136–146 (2017).Article 

    Google Scholar 
    28.Williams, G. C. Pleiotropy, natural selection, and the evolution of senescence. Evolution 11, 398–411 (1957).Article 

    Google Scholar 
    29.Hamilton, W. D. The moulding of senescence by natural selection. J. Theor. Biol. 12, 12–45 (1966).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    30.Kirkwood, T. B. L. Evolution of ageing. Nature 270, 301–304 (1977).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    31.Kirkwood, T. B. L. & Rose, M. R. Evolution of senescence: late survival sacrificed for reproduction. Philos. Trans. R. Soc. Lond., B 332, 15–24 (1991).CAS 
    Article 

    Google Scholar 
    32.Froese, R. & Pauly, D. FishBase. World Wide Web Electronic Publication (2019). Available at: www.fishbase.org. (accessed: 6th February 2018)33.I. C. E. S. Herring (Clupea harengus) in Subarea 4 and divisions 3.a and 7.d, autumn spawners (North Sea, Skagerrak and Kattegat, eastern English Channel). in Report of the ICES Advisory Committee, 2019. ICES Advice 2019, her.27.3a47d 11 (2019).34.Caswell, H. & Shyu, E. Senescence, selection gradients and mortality. in The Evolution of Senescence in the Tree of Life (eds. Shefferson, R. P., Jones, O. R. & Salguero-Gómez, R.) 56–82 (Cambridge University Press, 2017). https://doi.org/10.1017/9781139939867.00435.Promislow, D. E. L. Senescence in natural populations of mammals: a comparative study. Evolution 45, 1869–1887 (1991).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    36.Sibly, R. M., Collett, D., Promislow, D. E. L., Peacock, D. J. & Harvey, P. H. Mortality rates of mammals. J. Zool. 243, 1–12 (1997).Article 

    Google Scholar 
    37.Blumstein, D. T. & Møller, A. P. Is sociality associated with high longevity in North American birds? Biol. Lett. 4, 146–148 (2008).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    38.Nussey, D. H., Froy, H., Lemaitre, J.-F., Gaillard, J.-M. & Austad, S. N. Senescence in natural populations of animals: Widespread evidence and its implications for bio-gerontology. Ageing Res. Rev. 12, 214–225 (2013).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    39.Salguero-Gómez, R. & Jones, O. R.. Life history trade-offs modulate the speed of senescence. in The Evolution of Senescence in the Tree of Life (eds. Shefferson, R. P., Jones, O. R. & Salguero-Gómez, R.) 403–421 (Cambridge University Press, 2017). https://doi.org/10.1017/9781139939867.02040.Hoekstra, L. A., Schwartz, T. S., Sparkman, A. M., Miller, D. A. W. & Bronikowski, A. M. The untapped potential of reptile biodiversity for understanding how and why animals age. Funct. Ecol. 34, 38–54 (2020).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    41.Bonduriansky, R. & Brassil, C. E. Rapid and costly ageing in wild male flies. Nature 420, 377 (2002).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    42.Zajitschek, F., Zajitschek, S. & Bonduriansky, R. Senescence in wild insects: Key questions and challenges. Funct. Ecol. 34, 26–37 (2020).Article 

    Google Scholar 
    43.Roach, D. A. & Smith, E. F. Life-history trade-offs and senescence in plants. Funct. Ecol. 34, 17–25 (2020).Article 

    Google Scholar 
    44.Jones, O. R. et al. Diversity of ageing across the tree of life. Nature 505, 169–173 (2014).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    45.Ruby, J. G., Smith, M. & Buffenstein, R. Naked mole-rat mortality rates defy Gompertzian laws by not increasing with age. Elife 7, 1–18 (2018).Article 

    Google Scholar 
    46.Keller, L. & Genoud, M. Extraordinary lifespans in ants: a test of evolutionary theories of ageing. Nature 389, 958–960 (1997).CAS 
    Article 

    Google Scholar 
    47.Cooke, G. M., Tonkins, B. M. & Mather, J. A. Care and Enrichment for Captive Cephalopods. in The Welfare of Invertebrate Animals (eds. Carere, C. & Mather, J.). 179–208 (Springer, Cham, 2019). https://doi.org/10.1007/978-3-030-13947-6_848.Baudisch, A. et al. The pace and shape of senescence in angiosperms. J. Ecol. 101, 596–606 (2013).Article 

    Google Scholar 
    49.Halley, J. M., Van Houtan, K. S. & Mantua, N. How survival curves affect populations’ vulnerability to climate change. PLoS One 13, 1–18 (2018).Article 
    CAS 

    Google Scholar 
    50.Gompertz, B. On the nature of the function expressive of the law of human mortality, and on a new mode of determining the value of life contingencies. Philos. Trans. R. Soc. Lond. 115, 513–583 (1825).
    Google Scholar 
    51.Makeham, W. M. On the law of mortality and the construction of annuity tables. Assur. Mag. J. Inst. Actuar. 8, 301–310 (1860).Article 

    Google Scholar 
    52.Finch, C. E. & Pike, M. C. Maximum life span predictions from the Gompertz mortality model. J. Gerontol. Biol. Sci. 51A, 183–194 (1996).Article 

    Google Scholar 
    53.Reznick, D. N., Bryant, M. J., Roff, D., Ghalambor, C. K. & Ghalambor, D. E. Effect of extrinsic mortality on the evolution of senescence in guppies. Nature 431, 1095–1099 (2004).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    54.Kirkwood, T. B. L. Deciphering death: a commentary on Gompertz (1825) ‘On the nature of the function expressive of the law of human mortality, and on a new mode of determining the value of life contingencies’. Philos. Trans. R. Soc. B Biol. Sci. 370, 1–8 (2015).Article 

    Google Scholar 
    55.Gavrilov, L. A. & Gavrilova, N. S. New trend in old-age mortality: gompertzialization of mortality trajectory. Gerontology 65, 451–457 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    56.Ohsumi, S. Interspecies relationships among some biological parameters in cetaceans and estimation of the natural mortality coefficient of the Southern Hemisphere minke whale. Rep. Int. Whal. Comm. 29, 397–406 (1979).
    Google Scholar 
    57.Mizroch, S. A. On the relationship between mortality rate and length in baleen whales. Rep. Int. Whal. Comm. 35, 505–510 (1985).
    Google Scholar  More

  • in

    North Atlantic warming over six decades drives decreases in krill abundance with no associated range shift

    1.Astthorsson, O. S. & Palsson, O. K. Predation on euphausiids by cod, Gadus morhua, in winter in Icelandic subarctic waters. Mar. Biol. 96, 327–334 (1987).Article 

    Google Scholar 
    2.MacAulay, M. C., Wishner, K. F. & Daly, K. L. Acoustic scattering from zooplankton and micronekton in relation to a whale feeding site near Georges Bank and Cape Cod. Cont. Shelf Res. 15, 509–537 (1995).Article 

    Google Scholar 
    3.Víkingsson, G. A. Feeding of fin whales (Balaenoptera physalus) off Iceland – diurnal and seasonal variation and possible rates. J. Northwest Atl. Fish. Sci. 22, 77–89 (1997).Article 

    Google Scholar 
    4.Tarling, G. A., Ensor, N. S., Fregin, T., Goodall-Copestake, W. P. & Fretwell, P. An introduction to the biology of Northern krill (Meganyctiphanes norvegica Sars). Adv. Mar. Biol. 57, 1–40 (2010).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    5.McBride, M. M. et al. Krill, climate, and contrasting future scenarios for Arctic and Antarctic fisheries. ICES J. Mar. Sci. 71, 1934–1955 (2014).Article 

    Google Scholar 
    6.Orlova, E.L. et al. Climatic and ecological drivers of euphausiid community structure vary spatially in the Barents Sea: relationships from a long time series (1952–2009). Front. Mar. Sci. 1, 1–13 (2015).Article 

    Google Scholar 
    7.Silva, T. et al. Long-term changes of euphausiids in shelf and oceanic habitats southwest, south and southeast of Iceland. J. Plankton Res. 36, 1262–1278 (2014).Article 

    Google Scholar 
    8.Warner, A. J. & Hays, G. C. Sampling by the Continuous Plankton Recorder Survey. Prog. Oceanogr. 6611, 237–256 (1994).Article 

    Google Scholar 
    9.Williams, R. & Lindley, J. A. Variability in abundance, vertical distribution and ontogenetic migrations of Thysanoessa longicaudata (Crustacea: Euphausiacea) in the north-eastern Atlantic Ocean. Mar. Biol. 69, 321–330 (1982).Article 

    Google Scholar 
    10.Beaugrand, G., Luczak, C. & Edwards, M. Rapid biogeographical plankton shifts in the North Atlantic Ocean. Glob. Chang. Biol. 15, 1790–1803 (2009).Article 

    Google Scholar 
    11.Edwards, M., Beaugrand, G., Helaouët, P., Alheit, J. & Coombs, S. Marine ecosystem response to the Atlantic Multidecadal Oscillation. PLoS ONE 8, e57212 (2013).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    12.Behrenfeld, M. J. et al. Climate-driven trends in contemporary ocean productivity. Nature 444, 752–755 (2006).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    13.Harris, V., Edwards, M. & Olhede, S. C. Multidecadal Atlantic climate variability and its impact on marine pelagic communities. J. Mar. Syst. 133, 55–69 (2014).Article 

    Google Scholar 
    14.Beaugrand, G., Edwards, M., Brander, K., Luczak, C. & Ibanez, F. Causes and projections of abrupt climate-driven ecosystem shifts in the North Atlantic. Ecol. Lett. 11, 1157–1168 (2008).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    15.Edwards, M., Beaugrand, G., Hays, G. C., Koslow, J. A. & Richardson, A. J. Multi-decadal oceanic ecological datasets and their application in marine policy and management. Trends Ecol. Evol. 25, 602–610 (2010).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    16.Harris, V., Olhede, S. C. & Edwards, M. Multidecadal spatial reorganisation of plankton communities in the North East Atlantic. J. Mar. Syst. 142, 16–24 (2015).Article 

    Google Scholar 
    17.Reid, P. C. & Edwards, M. Long-term changes in the pelagos, benthos and fisheries of the North Sea. Senckenbergiana maritima. 32, 107–115 (2001).Article 

    Google Scholar 
    18.Gregory, B., Christophe, L. & Martin, E. Rapid biogeographical plankton shifts in the North Atlantic Ocean. Glob. Chang. Biol. 15, 1790–1803 (2009).Article 

    Google Scholar 
    19.Beaugrand, G., Brander, K. M., Alistair Lindley, J., Souissi, S. & Reid, P. C. Plankton effect on cod recruitment in the North Sea. Nature 426, 661–664 (2003).CAS 
    PubMed 
    Article 

    Google Scholar 
    20.Atkinson, A., Siegel, V., Pakhomov, E. & Rothery, P. Long-term decline in krill stock and increase in salps within the Southern Ocean. Nature 432, 100–103 (2004).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    21.Atkinson, A. et al. Krill (Euphausia superba) distribution contracts southward during rapid regional warming. Nat. Clim. Chang. 9, 142–147 (2019).Article 

    Google Scholar 
    22.Flores, H. et al. Impact of climate change on Antarctic krill. Mar. Ecol. Progr. Ser. 458, 1–19 (2012).Article 

    Google Scholar 
    23.Bograd, S. J., Checkley, D. A. & Wooster, W. S. CalCOFI: a half century of physical, chemical, and biological research in the California Current System. Deep Sea Res. Part II Top. Stud. Oceanogr. 50, 2349–2353 (2003).Article 

    Google Scholar 
    24.Brinton, E. & Townsend, A. Decadal variability in abundances of the dominant euphausiid species in southern sectors of the California Current. Deep Sea Res. Part II Top. Stud. Oceanogr. 50, 2449–2472 (2003).Article 

    Google Scholar 
    25.Polyakov, I. V. et al. Greater role for Atlantic inflows on sea-ice loss in the Eurasian Basin of the Arctic Ocean. Science 356, 285–291 (2017).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    26.Caesar, A.L., Rahmstorf, S., Robinson, A., Feulner, G. & Saba, V. Observed fingerprint of a weakening Atlantic Ocean overturning circulation.Nature 556, 191–196 (2018).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    27.Reid, P. C. et al. A biological consequence of reducing Arctic ice cover: arrival of the Pacific diatom Neodenticula seminae in the North Atlantic for the first time in 800 000 years. Glob. Chang. Biol. 13, 1910–1921 (2007).Article 

    Google Scholar 
    28.Burrows, M. et al. Ocean community warming responses explained by thermal affinities and temperature gradients. Nat. Clim. Chang. 9, 959–963 (2019).Article 

    Google Scholar 
    29.Biri, S. & Klein, B. North Atlantic sub‐polar gyre climate index: a new approach. J. Geophys. Res. Ocean. 124, 4222–4237 (2019).Article 

    Google Scholar 
    30.Batten, S. et al. CPR sampling: the technical background, materials and methods, consistency and comparability. Prog. Oceanogr. 58, 193–215 (2003).Article 

    Google Scholar 
    31.Reid, P. C. et al. The Continuous Plankton Recorder: concepts and history, from plankton indicator to undulating recorders. Prog. Oceanogr. 58, 117–173 (2003).Article 

    Google Scholar 
    32.Richardson, A. J. et al. Using Continuous Plankton Recorder data. Prog. Oceanogr. 68, 27–74 (2006).Article 

    Google Scholar 
    33.Dalpadado, P., Yamaguchi, A., Ellertsen, B. & Johannessen, S. Trophic interactions of macro-zooplankton (krill and amphipods) in the marginal ice zone of the Barents Sea. Deep. Sea Res. Part II Top. Stud. Oceanogr. 55, 2266–2274 (2008).Article 

    Google Scholar 
    34.Letessier, T. B., Cox, M. J. & Brierley, A. S. Drivers of euphausiid species abundance and numerical abundance in the Atlantic Ocean. Mar. Biol. 156, 2539–2553 (2009).Article 

    Google Scholar 
    35.Lowe, M. R., Lawson, G. L. & Fogarty, M. J. Drivers of euphausiid distribution and abundance in the Northeast U.S. Shelf Large Marine Ecosystem. ICES J. Mar. Sci. 75, 1280–1295 (2018).Article 

    Google Scholar 
    36.Lindley, J. A. Population dynamics and production of euphausiids. I. Thysanoessa longicaudata in the North Atlantic Ocean. Mar. Biol. 46, 121–130 (1978).Article 

    Google Scholar 
    37.Lindley, J. A. Population dynamics and production of euphausiids II. Thysanoessa inermis and T. raschii in the North Sea and American Coastal Waters. Mar. Biol. 59, 225–233 (1980).Article 

    Google Scholar 
    38.Lindley, J. A. Population dynamics and production of euphausiids. Mar. Biol. 71, 1–6 (1982).Article 

    Google Scholar 
    39.Rayner, N. A. Global analyses of sea surface temperature, sea ice, and night marine air temperature since the late nineteenth century. J. Geophys. Res. 108, 4407 (2003).Article 

    Google Scholar 
    40.Edwards, M., Johns, D. G. D., Leterme, S. C. S., Svendsen, E. & Richardson, A. J. A. Regional climate change and harmful algal blooms in the northeast Atlantic. Limnol. Oceanogr. 51, 820–829 (2006).Article 

    Google Scholar 
    41.Hélaouët, P., Beaugrand, G. & Reygondeau, G. Reliability of spatial and temporal patterns of C. finmarchicus inferred from the CPR survey. J. Mar. Syst. 153, 18–24 (2016).Article 

    Google Scholar 
    42.Owens, N. J. P. et al. All plankton sampling systems underestimate abundance: response to “Continuous Plankton Recorder underestimates zooplankton abundance” by J.W. Dippner and M. Krause. J. Mar. Syst. 128, 240–242 (2013).Article 

    Google Scholar 
    43.Jonas, T. D., Walne, A., Beaugrand, G., Gregory, L. & Hays, G. C. The volume of water filtered by a Continuous Plankton Recorder sample: the effect of ship speed. J. Plankton Res. 26, 1499–1506 (2004).Article 

    Google Scholar  More

  • in

    Differential gene expression in Drosophila melanogaster and D. nigrosparsa infected with the same Wolbachia strain

    1.Stark, R., Grzelak, M. & Hadfield, J. RNA sequencing: The teenage years. Nat. Rev. Genet. https://doi.org/10.1038/s41576-019-0150-2 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    2.Gutzwiller, F. et al. Dynamics of Wolbachia pipientis gene expression across the Drosophila melanogaster life cycle. G3 Genes Genomes Genet. 5, 2843–2856 (2015).CAS 

    Google Scholar 
    3.Bennuru, S. et al. Stage-specific transcriptome and proteome analyses of the filarial parasite Onchocerca volvulus and its Wolbachia endosymbiont. MBio 7, e02028-e2116 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    4.Baião, G. C., Schneider, D. I., Miller, W. J. & Klasson, L. The effect of Wolbachia on gene expression in Drosophila paulistorum and its implications for symbiont-induced host speciation. BMC Genom. 20, 465 (2019).Article 
    CAS 

    Google Scholar 
    5.Werren, J. H., Baldo, L. & Clark, M. E. Wolbachia: Master manipulators of invertebrate biology. Nat. Rev. Microbiol. 6, 741–751 (2008).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    6.Zug, R. & Hammerstein, P. Still a host of hosts for Wolbachia: Analysis of recent data suggests that 40% of terrestrial arthropod species are infected. PLoS One 7, e38544 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 
    ADS 

    Google Scholar 
    7.Sazama, E. J., Bosch, M. J., Shouldis, C. S., Ouellette, S. P. & Wesner, J. S. Incidence of Wolbachia in aquatic insects. Ecol. Evol. 7, 1165–1169 (2017).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    8.Detcharoen, M., Arthofer, W., Schlick-Steiner, B. C. & Steiner, F. M. Wolbachia megadiversity: 99% of these microorganismic manipulators unknown. FEMS Microbiol. Ecol. 95, fiz151 (2019).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    9.Hosokawa, T., Koga, R., Kikuchi, Y., Meng, X. Y. & Fukatsu, T. Wolbachia as a bacteriocyte-associated nutritional mutualist. Proc. Natl. Acad. Sci. U. S. A. 107, 769–774 (2010).CAS 
    PubMed 
    Article 
    ADS 
    PubMed Central 

    Google Scholar 
    10.Teixeira, L., Ferreira, Á. & Ashburner, M. The bacterial symbiont Wolbachia induces resistance to RNA viral infections in Drosophila melanogaster. PLoS Biol. 6, 2753–2763 (2008).CAS 
    Article 

    Google Scholar 
    11.Hedges, L. M., Brownlie, J. C., O’Neill, S. L. & Johnson, K. N. Wolbachia and virus protection in insects. Science (80-). 322, 702–702 (2008).CAS 
    Article 
    ADS 

    Google Scholar 
    12.Osborne, S. E., Leong, Y. S., O’Neill, S. L. & Johnson, K. N. Variation in antiviral protection mediated by different Wolbachia strains in Drosophila simulans. PLoS Pathog. 5, e1000656 (2009).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    13.Cattel, J., Martinez, J., Jiggins, F., Mouton, L. & Gibert, P. Wolbachia-mediated protection against viruses in the invasive pest Drosophila suzukii. Insect Mol. Biol. 25, 595–603 (2016).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    14.Ranz, J. M., Castillo-Davis, C. I., Meiklejohn, C. D. & Hartl, D. L. Sex-dependent gene expression and evolution of the Drosophila transcriptome. Science (80-). 300, 1742–1745 (2003).CAS 
    Article 
    ADS 

    Google Scholar 
    15.Herbert, R. I. & McGraw, E. A. The nature of the immune response in novel Wolbachia-host associations. Symbiosis 74, 225–236 (2018).Article 

    Google Scholar 
    16.Woodford, L. et al. Vector species-specific association between natural Wolbachia infections and avian malaria in black fly populations. Sci. Rep. 8, 4188 (2018).PubMed 
    PubMed Central 
    Article 
    ADS 
    CAS 

    Google Scholar 
    17.Huigens, M. E., De Almeida, R. P., Boons, P. A. H., Luck, R. F. & Stouthamer, R. Natural interspecific and intraspecific horizontal transfer of parthenogenesis-inducing Wolbachia in Trichogramma wasps. Proc. R. Soc. B Biol. Sci. 271, 509–515 (2004).CAS 
    Article 

    Google Scholar 
    18.Detcharoen, M., Arthofer, W., Jiggins, F. M., Steiner, F. M. & Schlick-Steiner, B. C. Wolbachia affect behavior and possibly reproductive compatibility but not thermoresistance, fecundity, and morphology in a novel transinfected host, Drosophila nigrosparsa. Ecol. Evol. 10, 4457–4470 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    19.Woolfit, M. et al. Genomic evolution of the pathogenic Wolbachia strain, wMelPop. Genome Biol. Evol. 5, 2189–2204 (2013).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    20.Suh, E., Mercer, D. R., Fu, Y. & Dobson, S. L. Pathogenicity of life-shortening Wolbachia in Aedes albopictus after transfer from Drosophila melanogaster. Appl. Environ. Microbiol. 75, 7783–7788 (2009).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    21.McGraw, E. A., Merritt, D. J., Droller, J. N. & O’Neill, S. L. Wolbachia-mediated sperm modification is dependent on the host genotype in Drosophila. Proc. R. Soc. B Biol. Sci. 268, 2565–2570 (2001).CAS 
    Article 

    Google Scholar 
    22.Xie, J., Vilchez, I. & Mateos, M. Spiroplasma bacteria enhance survival of Drosophila hydei attacked by the parasitic wasp Leptopilina heterotoma. PLoS One 5, e12149 (2010).PubMed 
    PubMed Central 
    Article 
    ADS 
    CAS 

    Google Scholar 
    23.Hutchence, K. J., Fischer, B., Paterson, S. & Hurst, G. D. D. How do insects react to novel inherited symbionts? A microarray analysis of Drosophila melanogaster response to the presence of natural and introduced Spiroplasma. Mol. Ecol. 20, 950–958 (2011).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    24.O’Grady, P. M. & DeSalle, R. Phylogeny of the genus Drosophila. Genetics 209, 1–25 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    25.Kellermann, V., Van Heerwaarden, B., Sgrò, C. M. & Hoffmann, A. A. Fundamental evolutionary limits in ecological traits drive Drosophila species distributions. Science (80-). 325, 1244–1246 (2009).CAS 
    Article 
    ADS 

    Google Scholar 
    26.Bächli, G., Viljoen, F., Escher, S. A. & Saura, A. The Drosophilidae (Diptera) of Fennoscandia and Denmark (Brill, 2005).27.Kinzner, M.-C. et al. Life-history traits and physiological limits of the alpine fly Drosophila nigrosparsa (Diptera: Drosophilidae): A comparative study. Ecol. Evol. 8, 2006–2020 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    28.Kinzner, M.-C. et al. Major range loss predicted from lack of heat adaptability in an alpine Drosophila species. Sci. Total Environ. 695, 133753 (2019).CAS 
    PubMed 
    Article 
    ADS 
    PubMed Central 

    Google Scholar 
    29.Kinzner, M.-C. et al. Oviposition substrate of the mountain fly Drosophila nigrosparsa (Diptera: Drosophilidae). PLoS One 11, e0165743 (2016).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    30.Cicconardi, F. et al. Chemosensory adaptations of the mountain fly Drosophila nigrosparsa (Insecta: Diptera) through genomics’ and structural biology’s lenses. Sci. Rep. 7, 43770 (2017).PubMed 
    PubMed Central 
    Article 
    ADS 

    Google Scholar 
    31.Tratter Kinzner, M. et al. Is temperature preference in the laboratory ecologically relevant for the field? The case of Drosophila nigrosparsa. Glob. Ecol. Conserv. 18, e00638 (2019).Article 

    Google Scholar 
    32.Arthofer, W. et al. Genomic resources notes accepted 1 August 2014–30 September 2014. Mol. Ecol. Resour. 15, 228–229 (2015).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    33.Cicconardi, F., Marcatili, P., Arthofer, W., Schlick-Steiner, B. C. & Steiner, F. M. Positive diversifying selection is a pervasive adaptive force throughout the Drosophila radiation. Mol. Phylogenet. Evol. 112, 230–243 (2017).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    34.Verspoor, R. L. & Haddrill, P. R. Genetic diversity, population structure and Wolbachia infection status in a worldwide sample of Drosophila melanogaster and D. simulans populations. PLoS One 6, e26318 (2011).CAS 
    PubMed 
    PubMed Central 
    Article 
    ADS 

    Google Scholar 
    35.Lints, F. A. Size in relation to development-time and egg-density in Drosophila melanogaster. Nature 197, 1128–1130 (1963).Article 
    ADS 

    Google Scholar 
    36.Clemson, A. S., Sgrò, C. M. & Telonis-Scott, M. Thermal plasticity in Drosophila melanogaster populations from eastern Australia: Quantitative traits to transcripts. J. Evol. Biol. 29, 2447–2463 (2016).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    37.Morozova, T. V., Anholt, R. H. & Mackay, T. F. Transcriptional response to alcohol exposure in Drosophila melanogaster. Genome Biol. 7, R95 (2006).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    38.Elya, C., Zhang, V., Ludington, W. B. & Eisen, M. B. Stable host gene expression in the gut of adult Drosophila melanogaster with different bacterial mono-associations. PLoS One 11, e0167357 (2016).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    39.Chrostek, E. et al. Wolbachia variants induce differential protection to viruses in Drosophila melanogaster: A phenotypic and phylogenomic analysis. PLoS Genet. 9, e1003896 (2013).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    40.Zhang, B. et al. Comparative transcriptome analysis of chemosensory genes in two sister leaf beetles provides insights into chemosensory speciation. Insect Biochem. Mol. Biol. 79, 108–118 (2016).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    41.Gazara, R. K. et al. De novo transcriptome sequencing and comparative analysis of midgut tissues of four non-model insects pertaining to Hemiptera, Coleoptera, Diptera and Lepidoptera. Gene 627, 85–93 (2017).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    42.Braig, H. R., Zhou, W., Dobson, S. L. & O’Neill, S. L. Cloning and characterization of a gene encoding the major surface protein of the bacterial endosymbiont Wolbachia pipientis. J. Bacteriol. 180, 2373–2378 (1998).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    43.Bolger, A. M., Lohse, M. & Usadel, B. Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120 (2014).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    44.Patro, R., Duggal, G., Love, M. I., Irizarry, R. A. & Kingsford, C. Salmon provides fast and bias-aware quantification of transcript expression. Nat. Methods 14, 417–419 (2017).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    45.R Core Team. R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, 2020), https://www.R-project.org.46.Soneson, C., Love, M. I. & Robinson, M. D. Differential analyses for RNA-seq: Transcript-level estimates improve gene-level inferences. F1000Research 4, 1521 (2016).PubMed Central 
    Article 

    Google Scholar 
    47.Thurmond, J. et al. FlyBase 2.0: The next generation. Nucleic Acids Res. 47, D759–D765 (2019).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    48.Hardcastle, T. J. & Kelly, K. A. BaySeq: Empirical Bayesian methods for identifying differential expression in sequence count data. BMC Bioinform. 11, 422 (2010).Article 

    Google Scholar 
    49.Huang, D. W., Sherman, B. T. & Lempicki, R. A. Bioinformatics enrichment tools: Paths toward the comprehensive functional analysis of large gene lists. Nucleic Acids Res. 37, 1–13 (2009).Article 
    CAS 

    Google Scholar 
    50.Gaujoux, R. & Seoighe, C. A flexible R package for nonnegative matrix factorization. BMC Bioinform. 11, 367 (2010).Article 
    CAS 

    Google Scholar 
    51.Oksanen, J. et al. vegan: Community Ecology Package. R package version 2.5-6. (2019) https://cran.r-project.org/web/packages/vegan/.52.Love, M. I., Huber, W. & Anders, S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550 (2014).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    53.Wickham, H. ggplot2. Wiley Interdiscip. Rev. Comput. Stat. 3, 180–185 (2011).54.Wittkopp, P. J. Variable gene expression in eukaryotes: A network perspective. J. Exp. Biol. 210, 1567–1575 (2007).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    55.Lin, Y., Chen, Z.-X., Oliver, B. & Harbison, S. T. Microenvironmental gene expression plasticity among individual Drosophila melanogaster. G3 Genes Genomes Genet. 6, 4197–4210 (2016).CAS 

    Google Scholar 
    56.Kristensen, T. N., Sørensen, P., Pedersen, K. S., Kruhøffer, M. & Loeschcke, V. Inbreeding by environmental interactions affect gene expression in Drosophila melanogaster. Genetics 173, 1329–1336 (2006).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    57.Dunning, L. T., Dennis, A. B., Sinclair, B. J., Newcomb, R. D. & Buckley, T. R. Divergent transcriptional responses to low temperature among populations of alpine and lowland species of New Zealand stick insects (Micrarchus). Mol. Ecol. 23, 2712–2726 (2014).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    58.Lambert, A. J. & Brand, M. D. Reactive oxygen species production by mitochondria. In Mitochondrial DNA. Methods in Molecular Biology (ed. Stuart, J. A.) vol. 554 165–181 (Humana Press, 2009).
    Google Scholar 
    59.Kurz, M. et al. Structural and functional characterization of the oxidoreductase α-DsbA1 from Wolbachia pipientis. Antioxidants Redox Signal. 11, 1485–1500 (2009).CAS 
    Article 

    Google Scholar 
    60.Zug, R. & Hammerstein, P. Wolbachia and the insect immune system: What reactive oxygen species can tell us about the mechanisms of Wolbachia-host interactions. Front. Microbiol. 6, 1201 (2015).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    61.Ratzka, C., Gross, R. & Feldhaar, H. Endosymbiont tolerance and control within insect hosts. Insects 3, 553–572 (2012).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    62.Pan, X. et al. Wolbachia induces reactive oxygen species (ROS)-dependent activation of the Toll pathway to control dengue virus in the mosquito Aedes aegypti. Proc. Natl. Acad. Sci. U. S. A. 109, E23-31 (2012).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    63.Brennan, L. J., Haukedal, J. A., Earle, J. C., Keddie, B. & Harris, H. L. Disruption of redox homeostasis leads to oxidative DNA damage in spermatocytes of Wolbachia-infected Drosophila simulans. Insect Mol. Biol. 21, 510–520 (2012).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    64.Blagrove, M. S. C., Arias-Goeta, C., Failloux, A.-B. & Sinkins, S. P. Wolbachia strain wMel induces cytoplasmic incompatibility and blocks dengue transmission in Aedes albopictus. Proc. Natl. Acad. Sci. 109, 255–260 (2012).CAS 
    PubMed 
    Article 
    ADS 
    PubMed Central 

    Google Scholar 
    65.Andrews, E. S., Crain, P. R., Fu, Y., Howe, D. K. & Dobson, S. L. Reactive oxygen species production and Brugia pahangi survivorship in Aedes polynesiensis with artificial Wolbachia infection types. PLoS Pathog. 8, e1003075 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    66.Oliveira, M. F. et al. Haem detoxification by an insect. Nature 400, 517–518 (1999).CAS 
    PubMed 
    Article 
    ADS 
    PubMed Central 

    Google Scholar 
    67.Paiva-Silva, G. O. et al. A heme-degradation pathway in a blood-sucking insect. Proc. Natl. Acad. Sci. U. S. A. 103, 8030–8035 (2006).CAS 
    PubMed 
    PubMed Central 
    Article 
    ADS 

    Google Scholar 
    68.Levi, S. & Rovida, E. The role of iron in mitochondrial function. Biochim. Biophys. Acta Gen. Subj. 1790, 629–636 (2009).CAS 
    Article 

    Google Scholar 
    69.Kremer, N. et al. Wolbachia interferes with ferritin expression and iron metabolism in insects. PLoS Pathog. 5, e1000630 (2009).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    70.Kremer, N. et al. Influence of Wolbachia on host gene expression in an obligatory symbiosis. BMC Microbiol. 12, S7 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 
    ADS 

    Google Scholar 
    71.Peng, Y., Nielsen, J. E., Cunningham, J. P. & McGraw, E. A. Wolbachia infection alters olfactory-cued locomotion in Drosophila spp. Appl. Environ. Microbiol. 74, 3943–3948 (2008).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    72.Peng, Y. & Wang, Y. Infection of Wolbachia may improve the olfactory response of Drosophila. Chin. Sci. Bull. 54, 1369–1375 (2009).
    Google Scholar 
    73.Fattouh, N., Cazevieille, C. & Landmann, F. Wolbachia endosymbionts subvert the endoplasmic reticulum to acquire host membranes without triggering ER stress. PLoS Negl. Trop. Dis. 13, e0007218 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    74.Chagas-Moutinho, V. A., Silva, R., de Souza, W. & Motta, M. C. Identification and ultrastructural characterization of the Wolbachia symbiont in Litomosoides chagasfilhoi. Parasit. Vectors 8, 74 (2015).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    75.Serbus, L. R., Casper-Lindley, C., Landmann, F. & Sullivan, W. The genetics and cell biology of Wolbachia-host interactions. Annu. Rev. Genet. 42, 683–707 (2008).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    76.Ping, Y. et al. Linking Aβ42-induced hyperexcitability to neurodegeneration, learning and motor deficits, and a shorter lifespan in an Alzheimer’s model. PLoS Genet. 11, e1005025 (2015).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    77.Ping, Y. et al. Shal/Kv4 channels are required for maintaining excitability during repetitive firing and normal locomotion in Drosophila. PLoS One 6, e16043 (2011).CAS 
    PubMed 
    PubMed Central 
    Article 
    ADS 

    Google Scholar 
    78.Ping, Y. & Tsunoda, S. Inactivity-induced increase in nAChRs upregulates Shal K+ channels to stabilize synaptic potentials. Nat. Neurosci. 15, 90–97 (2012).CAS 
    Article 

    Google Scholar 
    79.Kim, W. J., Jan, L. Y. & Jan, Y. N. A PDF/NPF neuropeptide signaling circuitry of male Drosophila melanogaster controls rival-induced prolonged mating. Neuron 80, 1190–1205 (2013).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    80.King, A. N. et al. A peptidergic circuit links the circadian clock to locomotor activity. Curr. Biol. 27, 1915-1927.e5 (2017).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    81.Kim, Y. J., Žitňan, D., Galizia, C. G., Cho, K. H. & Adams, M. E. A command chemical triggers an innate behavior by sequential activation of multiple peptidergic ensembles. Curr. Biol. 16, 1395–1407 (2006).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    82.Rapaport, F. et al. Comprehensive evaluation of differential gene expression analysis methods for RNA-seq data. Genome Biol. 14, 3158 (2013).Article 
    CAS 

    Google Scholar 
    83.Kvam, V. M., Liu, P. & Si, Y. A comparison of statistical methods for detecting differentially expressed genes from RNA-seq data. Am. J. Bot. 99, 248–256 (2012).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    84.Guo, Y., Li, C. I., Ye, F. & Shyr, Y. Evaluation of read count based RNAseq analysis methods. BMC Genom. 14, S2 (2013).Article 

    Google Scholar  More

  • in

    Energetic context determines the effects of multiple upwelling-associated stressors on sea urchin performance

    1.Krumhansl, K. A. et al. Global patterns of kelp forest change over the past half-century. Proc. Natl. Acad. Sci. USA 113, 13785–13790 (2016).CAS 
    PubMed 
    Article 

    Google Scholar 
    2.Filbee-Dexter, K. & Scheibling, R. E. Sea urchin barrens as alternative stable states of collapsed kelp ecosystems. Mar. Ecol. Prog. Ser. 495, 1–25 (2014).ADS 
    Article 

    Google Scholar 
    3.Steneck, R. S. et al. Kelp forest ecosystems: biodiversity, stability, resilience and future. Environ. Conserv. 29, 436–459 (2002).Article 

    Google Scholar 
    4.Steneck, R. S. Regular sea urchins as drivers of shallow benthic marine community structure. Dev. Aquacult. Fish. Sci. 43, 255–279 (2020).
    Google Scholar 
    5.Rogers-Bennett, L. & Catton, C. A. Marine heat wave and multiple stressors tip bull kelp forest to sea urchin barrens. Sci. Rep. 9, 1–9 (2019).CAS 
    Article 

    Google Scholar 
    6.Pearse, J. S. Ecological role of purple sea urchins. Science 31, 940–941 (2006).ADS 
    Article 
    CAS 

    Google Scholar 
    7.Harrold, C. & Reed, D. C. Food availability, sea urchin grazing, and kelp forest community structure. Ecology 66, 1160–1169 (1985).Article 

    Google Scholar 
    8.Kriegisch, N., Reeves, S. E., Flukes, E. B., Johnson, C. R. & Ling, S. D. Drift-kelp suppresses foraging movement of overgrazing sea urchins. Oecologia 190, 665–677 (2019).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    9.Pearse, J. S. & Hines, A. H. Long-term population dynamics of sea urchins in a central California kelp forest: rare recruitment and rapid decline. Mar. Ecol. Prog. Ser. 39, 275–283 (1987).ADS 
    Article 

    Google Scholar 
    10.Watanabe, J. M. & Harrold, C. Destructive grazing by sea urchins Strongylocentrotus spp. in a central California kelp forest: potential roles of recruitment, depth, and predation. Mar. Ecol. Prog. Ser. 71, 125–141 (1991).ADS 
    Article 

    Google Scholar 
    11.Reid, J. et al. The economic value of the recreational red abalone fishery in northern California. Calif. Fish Game 102, 119–130 (2016).
    Google Scholar 
    12.Menge, B. A. & Menge, D. N. Dynamics of coastal meta-ecosystems: the intermittent upwelling hypothesis and a test in rocky intertidal regions. Ecol. Monogr. 83, 283–310 (2013).Article 

    Google Scholar 
    13.Breitburg, D. L., Loher, T., Pacey, C. A. & Gerstein, A. Varying effects of low dissolved oxygen on trophic interactions in an estuarine food web. Ecol. Monogr. 67, 489–507 (1997).Article 

    Google Scholar 
    14.Hauri, C. et al. (2009) Ocean acidification in the California current system. Oceanography 22, 60–71 (2009).Article 

    Google Scholar 
    15.Connell, S. D. & Russell, B. D. The direct effects of increasing CO2 and temperature on non-calcifying organisms: increasing the potential for phase shifts in kelp forests. Proc. R. Soc. B 277, 1409–1415 (2010).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    16.Sellers, A. J. et al. Seasonal upwelling reduces herbivore control of tropical rocky intertidal algal communities. Ecology e03335 https://doi.org/10.1002/ecy.3335(2021).17.Moulin, L., Grosjean, P., Leblud, J., Batigny, A. & Dubois, P. Impact of elevated pCO2 on acid-base regulation of the sea urchin Echinometra mathaei and its relation to resistance to ocean acidification: a study in mesocosms. J. Exp. Mar. Biol. Ecol. 457, 97–104 (2014).CAS 
    Article 

    Google Scholar 
    18.Siikavuopio, S. I., Dale, T., Mortensen, A. & Foss, A. Effects of hypoxia on feed intake and gonad growth in the green sea urchin, Strongylocentrotus droebachiensis. Aquaculture 266, 112–116 (2007).Article 

    Google Scholar 
    19.Low, H. N. N. The Effects of Upwelling-driven Hypoxia on Sea Urchins in California Current Kelp Forests. PhD dissertation, Stanford University, Stanford, CA (2018).20.Low, N. H. & Micheli, F. Lethal and functional thresholds of hypoxia in two key benthic grazers. Mar. Ecol. Prog. Ser. 594, 165–173 (2018).ADS 
    CAS 
    Article 

    Google Scholar 
    21.Low, N. H. & Micheli, F. Short-and long-term impacts of variable hypoxia exposures on kelp forest sea urchins. Sci. Rep. 10, 1–9 (2020).CAS 
    Article 

    Google Scholar 
    22.Huyer, A. Coastal upwelling in the California current system. Prog. Oceanogr. 12, 259–284 (1983).ADS 
    Article 

    Google Scholar 
    23.Frieder, C. A., Nam, S. H., Martz, T. R. & Levin, L. A. High temporal and spatial variability of dissolved oxygen and pH in a nearshore California kelp forest. Biogeosciences 9, 3917–3930 (2012).24.Feely, R. A. et al. The combined effects of acidification and hypoxia on pH and aragonite saturation in the coastal waters of the California current ecosystem and the northern Gulf of Mexico. Cont. Shelf Res. 152, 50–60 (2018).ADS 
    Article 

    Google Scholar 
    25.Chan, F. et al. Persistent spatial structuring of coastal ocean acidification in the California Current System. Sci. Rep. 7, 1–7 (2017).Article 
    CAS 

    Google Scholar 
    26.Orr, J. C. et al. Anthropogenic ocean acidification over the twenty-first century and its impact on calcifying organisms. Nature 437, 681–686 (2005).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    27.Feely, R. A. et al. Chemical and biological impacts of ocean acidification along the west coast of North America. Estuar. Coast. Shelf Sci. 183, 260–270 (2016).ADS 
    CAS 
    Article 

    Google Scholar 
    28.Iles, A. C. et al. Climate-driven trends and ecological implications of event-scale upwelling in the California Current System. Glob. Change Biol. 18, 783–796 (2012).ADS 
    Article 

    Google Scholar 
    29.CeNCOOS. Real-Time Sensor Feeds of Oceanographic and Atmospheric Models’ Online Tool to Extract Temperature, pH, and Dissolved Oxygen. https://data.cencoos.org (2020).30.Bakun, A. Global climate change and intensification of coastal ocean upwelling. Science 247, 198–201 (1990).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    31.McGregor, H. V., Dima, M., Fischer, H. W. & Mulitza, S. Rapid 20th-century increase in coastal upwelling off northwest Africa. Science 315, 637–639 (2007).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    32.Narayan, N., Paul, A., Mulitza, S. & Schulz, M. Trends in coastal upwelling intensity during the late 20th century. Ocean Sci. 6, 815–823 (2010).ADS 
    Article 

    Google Scholar 
    33.Barton, E. D. D., Field, D. B. B. & Roy, C. Canary current upwelling: more or less?. Prog. Oceanogr. 116, 167–178 (2013).ADS 
    Article 

    Google Scholar 
    34.Mote, P. W. & Mantua, N. J. Coastal upwelling in a warmer future. Geophys. Res. Lett. 29, 2138 (2002).ADS 
    Article 

    Google Scholar 
    35.Bakun, A. et al. Anticipated effects of climate change on coastal upwelling ecosystems. Curr. Clim. Change Rep. 1, 85–93 (2015).Article 

    Google Scholar 
    36.Wang, D., Gouhier, T. C., Menge, B. A. & Ganguly, A. R. Intensification and spatial homogenization of coastal upwelling under climate change. Nature 518, 390–394 (2015).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    37.Snyder, M. A., Sloan, L. C., Diffenbaugh, N. S. & Bell, J. L. Future climate change and upwelling in the California Current. Geophys. Res. Lett. 30, 1823 (2003).38.García‐Reyes, M. & Largier, J. Observations of increased wind‐driven coastal upwelling off central California. J. Geophys. Res. Oceans 115, 1–8 (2010).39.Varela, R., Álvarez, I., Santos, F., DeCastro, M. & Gómez-Gesteira, M. Has upwelling strengthened along worldwide coasts over 1982–2010?. Sci. Rep. 5, 1–15 (2015).40.Varela, R., Lima, F. P., Seabra, R., Meneghesso, C. & Gómez-Gesteira, M. Coastal warming and wind-driven upwelling: a global analysis. Sci. Total Environ. 639, 1501–1511 (2018).41.Abrahams, A., Schlegel, R. W. & Smit, A. J. Variation and change of upwelling dynamics detected in the world’s eastern boundary upwelling systems. Front. Mar. Sci. 8, 626411 (2021).Article 

    Google Scholar 
    42.IPCC Climate change 2014: Synthesis report. In Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change Vol. 151 (eds Core Writing Team et al.) (IPCC, Geneva, 2014).
    Google Scholar 
    43.Rykaczewski, R. R. & Dunne, J. P. Enhanced nutrient supply to the California Current Ecosystem with global warming and increased stratification in an earth system model. Geophys. Res. Lett. 37, 1-5 (2010).44.Somero, G. N. et al. What changes in the carbonate system, oxygen, and temperature portend for the northeastern Pacific Ocean: a physiological perspective. Bioscience 66, 14–26 (2016).Article 

    Google Scholar 
    45.Frölicher, T. L., Fischer, E. M. & Gruber, N. Marine heatwaves under global warming. Nature 560, 360–364 (2018).ADS 
    PubMed 
    Article 
    CAS 

    Google Scholar 
    46.Filbee-Dexter, K. et al. Marine heatwaves and the collapse of marginal North Atlantic kelp forests. Sci. Rep. 10, 13388 (2020).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    47.Sokolova, I. M., Frederich, M., Bagwe, R., Lannig, G. & Sukhotin, A. A. Energy homeostasis as an integrative tool for assessing limits of environmental stress tolerance in aquatic invertebrates. Mar. Environ. Res. 79, 1–15 (2012).CAS 
    PubMed 
    Article 

    Google Scholar 
    48.Sokolova, I. M. Energy-limited tolerance to stress as a conceptual framework to integrate the effects of multiple stressors. Integr. Comp. Biol. 53, 597–608 (2013).PubMed 
    Article 

    Google Scholar 
    49.Fitzgerald-Dehoog, L., Browning, J. & Allen, B. J. Food and heat stress in the California mussel: evidence for an energetic trade-off between survival and growth. Biol. Bull. 223, 205–216 (2012).PubMed 
    Article 

    Google Scholar 
    50.Ramajo, L. et al. Food supply confers calcifiers resistance to ocean acidification. Sci. Rep. 6, 19374 (2016).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    51.Brown, N. E., Bernhardt, J. R., Anderson, K. M. & Harley, C. D. Increased food supply mitigates ocean acidification effects on calcification but exacerbates effects on growth. Sci. Rep. 8, 1–9 (2018).
    Google Scholar 
    52.Wahle, R. A. & Peckham, S. H. Density-related reproductive trade-offs in the green sea urchin, Strongylocentrotus droebachiensis. Mar. Biol. 134, 127–137 (1999).Article 

    Google Scholar 
    53.Rogers-Bennett, L., Allen, B. L. & Rothaus, D. P. Status and habitat associations of the threatened northern abalone: importance of kelp and coralline algae. Aquat. Conserv. Mar. Freshw. Ecosyst. 21, 573–581 (2011).Article 

    Google Scholar 
    54.Brown, M. B., Edwards, M. S. & Kim, K. Y. Effects of climate change on the physiology of giant kelp, Macrocystis pyrifera, and grazing by purple urchin, Strongylocentrotus purpuratus. Algae 29, 203–215 (2014).CAS 
    Article 

    Google Scholar 
    55.Klinger, T. S. & Lawrence, J. M. Distance perception of food and the effect of food quantity on feeding behavior of Lytechinus variegatus (Lamarck) (Echinodermata: Echinoidea). Mar. Freshw. Behav. Physiol. 11, 327–344 (1985).Article 

    Google Scholar 
    56.Trowbridge, C. D. Establishment of the green alga Codium fragile ssp. tomentosoides on New Zealand rocky shores: current distribution and invertebrate grazers. J. Ecol. 83, 949–965 (1995).Article 

    Google Scholar 
    57.Meidel, S. K. & Scheibling, R. E. Effects of food type and ration on reproductive maturation and growth of the sea urchin Strongylocentrotus droebachiensis. Mar. Biol. 134, 155–166 (1999).Article 

    Google Scholar 
    58.Harianto, J., Nguyen, H. D., Holmes, S. P. & Byrne, M. The effect of warming on mortality, metabolic rate, heat-shock protein response and gonad growth in thermally acclimated sea urchins (Heliocidaris erythrogramma). Mar. Biol. 165, 1–12 (2018).CAS 
    Article 

    Google Scholar 
    59.Brown, J. H., Gillooly, J. F., Allen, A. P., Savage, V. M. & West, G. B. Toward a metabolic theory of ecology. Ecology 85, 1771–1789 (2004).Article 

    Google Scholar 
    60.Hobday, A. J. et al. A hierarchical approach to defining marine heatwaves. Prog. Oceanogr. 141, 227–238 (2016).ADS 
    Article 

    Google Scholar 
    61.Spicer, J. I., Widdicombe, S., Needham, H. R. & Berge, J. A. Impact of CO2-acidified seawater on the extracellular acid-base balance of the northern sea urchin Strongylocentrotus dröebachiensis. J. Exp. Mar. Biol. Ecol. 407, 19–25 (2011).CAS 
    Article 

    Google Scholar 
    62.Catarino, A. I., Bauwens, M. & Dubois, P. Acid–base balance and metabolic response of the sea urchin Paracentrotus lividus to different seawater pH and temperatures. Environ. Sci. Pollut. Res. 19, 2344–2353 (2012).CAS 
    Article 

    Google Scholar 
    63.Rogers-Bennett, L., Bennett, W. A., Fastenau, H. C. & Dewees, C. M. Spatial variation in red sea urchin reproduction and morphology: implications for harvest refugia. Ecol. Appl. 5, 1171–1180 (1995).Article 

    Google Scholar 
    64.Quinn, J. F., Wing, S. R. & Botsford, L. W. Harvest refugia in marine invertebrate fisheries: models and applications to the red sea urchin, Strongylocentrotus franciscanus. Am. Zool. 33, 537–550 (1993).Article 

    Google Scholar 
    65.Eurich, J. G., Selden, R. L. & Warner, R. R. California spiny lobster preference for urchins from kelp forests: implications for urchin barren persistence. Mar. Ecol. Prog. Ser. 498, 217–225 (2014).ADS 
    Article 

    Google Scholar 
    66.Steneck, R. S., Leland, A., McNaught, D. C. & Vavrinec, J. Ecosystem flips, locks, and feedbacks: the lasting effects of fisheries on Maine’s kelp forest ecosystem. Bull. Mar. Sci. 89, 31–55 (2013).Article 

    Google Scholar 
    67.Gerard, V. A. Growth and utilization of internal nitrogen reserves by the giant kelp Macrocystis pyrifera in a low-nitrogen environment. Mar. Biol. 66(1), 27–35 (1982).CAS 
    Article 

    Google Scholar 
    68.Simonson, E. J., Scheibling, R. E. & Metaxas, A. Kelp in hot water: I. Warming seawater temperature induces weakening and loss of kelp tissue. Mar. Ecol. Prog. Ser. 537, 89–104 (2015).ADS 
    CAS 
    Article 

    Google Scholar 
    69.Thomsen, M. S. et al. Local extinction of bull kelp (Durvillaea spp.) due to a marine heatwave. Front. Mar. Sci. 6, 84 (2019).Article 

    Google Scholar 
    70.O’Donnell, M. J., Hammond, L. M. & Hofmann, G. E. Predicted impact of ocean acidification on a marine invertebrate: elevated CO2 alters response to thermal stress in sea urchin larvae. Mar. Biol. 156, 439–446 (2009).Article 
    CAS 

    Google Scholar 
    71.Dupont, S., Dorey, N., Stumpp, M., Melzner, F. & Thorndyke, M. Long-term and trans-life-cycle effects of exposure to ocean acidification in the green sea urchin Strongylocentrotus droebachiensis. Mar. Biol. 160, 1835–1843 (2013).CAS 
    Article 

    Google Scholar 
    72.Marcel, E. V. et al. Variable responses to large-scale climate change in European Parus populations. Proc. R. Soc. B 270, 367–372 (2003).Article 

    Google Scholar 
    73.Parker, L. M., Ross, P. M. & O’Connor, W. A. Populations of the Sydney rock oyster, Saccostrea glomerata, vary in response to ocean acidification. Mar. Biol. 158, 689–697 (2011).Article 

    Google Scholar 
    74.Kroeker, K. J., Kordas, R. L. & Harley, C. D. Embracing interactions in ocean acidification research: confronting multiple stressor scenarios and context dependence. Biol. Lett. 13, 20160802 (2017).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    75.Conor, J. J. Gonad growth in the sea urchin, Strongylocentrotus purpuratus (Stimpson) (Echinodermata: Echinoidea) and the assumptions of gonad index methods. J. Exp. Mar. Biol. Ecol. 10, 89–103 (1972).Article 

    Google Scholar 
    76.Bandstra, L., Hales, B. & Takahashi, T. High-frequency measurements of total CO2: method development and first oceanographic observations. Mar. Chem. 100, 24–38 (2006).CAS 
    Article 

    Google Scholar 
    77.Hales, B., Chipman, D. & Takahashi, T. High-frequency measurement of partial pressure and total concentration of carbon dioxide in seawater using microporous hydrophobic membrane contactors. Limnol. Oceanogr. Methods 2, 356–364 (2004).Article 

    Google Scholar 
    78.Lavigne, H., Epitalon, J. M. & Gattuso, J. P. Seacarb: Seawater Carbonate Chemistry with R. R package version 3.0 http://CRAN.R-project.org/package=seacarb (2011).79.Gattuso, J. P., Epitalon, J. M., Lavigne, H. & Orr, J. Seacarb: seawater carbonate chemistry. R package version 3.2.10. http://CRAN.R-project.org/package=seacarb (2018).80.Murie, K. A. & Bourdeau, P. E. Fragmented kelp forest canopies retain their ability to alter local seawater chemistry. Sci. Rep. 10, 1–13 (2020).Article 
    CAS 

    Google Scholar 
    81.R Core Team. R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna, Austria. http://www.R-project.org/ (2013). More