More stories

  • in

    The population sizes and global extinction risk of reef-building coral species at biogeographic scales

    1.
    Wilkinson, C. Status of Coral Reefs of the World: 2008 (Global Coral Reef Monitoring Network and Reef and Rainforest Research Centre, 2008).
    2.
    Jackson, J. B. C., Donovan, M. K., Cramer, K. L. & Lam, V. V. Status and Trends of Caribbean Coral Reefs: 1970–2012 (Global Coral Reef Monitoring Network, 2014).

    3.
    Eakin, C. M. et al. Caribbean corals in crisis: record thermal stress, bleaching, and mortality in 2005. PLoS ONE 5, e13969 (2010).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    4.
    Baker, A. C., Glynn, P. W. & Riegl, B. Climate change and coral reef bleaching: an ecological assessment of long-term impacts, recovery trends and future outlook. Estuar. Coast. Shelf Sci. 80, 435–471 (2008).
    Article  Google Scholar 

    5.
    Hughes, T. P. et al. Global warming transforms coral reef assemblages. Nature 556, 492–496 (2018).
    CAS  Article  PubMed  Google Scholar 

    6.
    Hughes, T. P. et al. Spatial and temporal patterns of mass bleaching of corals in the Anthropocene. Science 359, 80–83 (2018).
    CAS  PubMed  Article  Google Scholar 

    7.
    De’ath, G., Fabricius, K. E., Sweatman, H. & Puotinen, M. The 27-year decline of coral cover on the Great Barrier Reef and its causes. Proc. Natl Acad. Sci. USA 109, 17995–17999 (2012).
    PubMed  Article  Google Scholar 

    8.
    Gardner, T. A. Long-term region-wide declines in Caribbean corals. Science 301, 958–960 (2003).
    CAS  PubMed  Article  Google Scholar 

    9.
    Carpenter, K. E. et al. One-third of reef-building corals face elevated extinction risk from climate change and local impacts. Science 321, 560–563 (2008).
    CAS  PubMed  Article  Google Scholar 

    10.
    ter Steege, H. et al. Estimating the global conservation status of more than 15,000 Amazonian tree species. Sci. Adv. 1, e1500936 (2015).
    PubMed  PubMed Central  Article  Google Scholar 

    11.
    Fauset, S. et al. Hyperdominance in Amazonian forest carbon cycling. Nat. Commun. 6, 6857 (2015).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    12.
    Crowther, T. W. et al. Mapping tree density at a global scale. Nature 525, 201–205 (2015).
    CAS  PubMed  Article  Google Scholar 

    13.
    Connell, J., Hughes, T. & Wallace, C. A 30-year study of coral abundance, recruitment, and disturbance at several scales in space and time. Ecol. Monogr. 67, 461–488 (1997).
    Article  Google Scholar 

    14.
    Hughes, T. P. & Jackson, J. B. C. Population dynamics and life histories of foliaceous corals. Ecol. Monogr. 55, 141–166 (1985).
    Article  Google Scholar 

    15.
    ter Steege, H. et al. Hyperdominance in the Amazonian tree flora. Science 342, 1243092 (2013).
    PubMed  Article  CAS  Google Scholar 

    16.
    Gaston, K. J. & Blackburn, T. M. How many birds are there? Biodivers. Conserv. 6, 615–625 (1997).
    Article  Google Scholar 

    17.
    Kerry, J. T. & Bellwood, D. R. Do tabular corals constitute keystone structures for fishes on coral reefs? Coral Reefs 34, 41–50 (2015).
    Article  Google Scholar 

    18.
    Connolly, S. R., Hughes, T. P., Bellwood, D. R. & Karlson, R. H. Community structure of corals and reef fishes at multiple scales. Science 309, 1363–1365 (2005).
    CAS  PubMed  Article  Google Scholar 

    19.
    Connolly, S. R., Hughes, T. P. & Bellwood, D. R. A unified model explains commonness and rarity on coral reefs. Ecol. Lett. 20, 477–486 (2017).
    PubMed  Article  Google Scholar 

    20.
    Hubbell, S. P. Estimating the global number of tropical tree species, and Fisher’s paradox. Proc. Natl Acad. Sci. USA 112, 7343–7344 (2015).
    CAS  PubMed  Article  Google Scholar 

    21.
    Hughes, T. P., Bellwood, D. R. & Connolly, S. R. Biodiversity hotspots, centres of endemicity, and the conservation of coral reefs. Ecol. Lett. 5, 775–784 (2002).
    Article  Google Scholar 

    22.
    Hughes, T. P., Bellwood, D. R., Connolly, S. R. & Cornell, H. V. Double jeopardy and global extinction risk in corals and reef fishes. Curr. Biol. 24, 2946–2951 (2014).
    CAS  PubMed  Article  Google Scholar 

    23.
    Kinlan, B. P. & Gaines, S. D. Propagule dispersal in marine and terrestrial environments: a community perspective. Ecology 84, 2007–2020 (2003).
    Article  Google Scholar 

    24.
    Hull, P. M., Darroch, S. A. F. & Erwin, D. H. Rarity in mass extinctions and the future of ecosystems. Nature 528, 345–351 (2015).
    CAS  PubMed  Article  Google Scholar 

    25.
    Cardoso, P., Borges, P. A. V., Triantis, K. A., Ferrández, M. A. & Martín, J. L. Adapting the IUCN Red List criteria for invertebrates. Biol. Conserv. 144, 2432–2440 (2011).
    Article  Google Scholar 

    26.
    Cardoso, P., Borges, P. A. V., Triantis, K. A., Ferrández, M. A. & Martín, J. L. The underrepresentation and misrepresentation of invertebrates in the IUCN Red List. Biol. Conserv. 149, 147–148 (2012).
    Article  Google Scholar 

    27.
    Estes, J. A., Duggins, D. O. & Rathbun, G. B. The ecology of extinctions in kelp forest communities. Conserv. Biol. 3, 252–264 (1989).
    Article  Google Scholar 

    28.
    Oliver, J. & Babcock, R. Aspects of the fertilization ecology of broadcast spawning corals: sperm dilution effects and in situ measurements of fertilization. Biol. Bull. 183, 409–417 (1992).
    CAS  PubMed  Article  Google Scholar 

    29.
    Knowlton, N., Lang, J. C. & Keller, B. D. Case study of natural population collapse: post-hurricane predation on Jamaican staghorn corals. Smithson. Contrib. Mar. Sci. 31, 1–25 (1990).
    Google Scholar 

    30.
    Gaston, K. J. & Fuller, R. A. Commonness, population depletion and conservation biology. Trends Ecol. Evol. 23, 14–19 (2008).
    PubMed  Article  Google Scholar 

    31.
    Säterberg, T., Sellman, S. & Ebenman, B. High frequency of functional extinctions in ecological networks. Nature 499, 468–470 (2013).
    PubMed  Article  CAS  Google Scholar 

    32.
    Pratchett, M. S. Dietary overlap among coral-feeding butterflyfishes (Chaetodontidae) at Lizard Island, northern Great Barrier Reef. Mar. Biol. 148, 373–382 (2005).
    Article  Google Scholar 

    33.
    Huang, D., Licuanan, W. Y., Baird, A. H. & Fukami, H. Cleaning up the ‘Bigmessidae’: molecular phylogeny of scleractinian corals from Faviidae, Merulinidae, Pectiniidae and Trachyphylliidae. BMC Evol. Biol. 11, 37 (2011).
    PubMed  PubMed Central  Article  Google Scholar 

    34.
    Knowlton, N. & Jackson, J. B. C. New taxonomy and niche partitioning on coral reefs: jack of all trades or master of some? Trends Ecol. Evol. 9, 7–9 (1994).
    CAS  PubMed  Article  Google Scholar 

    35.
    Gilpin, M. E. & Soulé, M. E. in Conservation Biology: The Science of Scarcity and Diversity (ed, Soulé, M. E.) 19–34 (Sinauer Associates, 1986).

    36.
    Bak, R. P. M. & Meesters, E. H. Population structure as a response of coral communities to global change. Am. Zool. 39, 56–65 (1999).
    Article  Google Scholar 

    37.
    McClanahan, T. R., Ateweberhan, M. & Omukoto, J. Long-term changes in coral colony size distributions on Kenyan reefs under different management regimes and across the 1998 bleaching event. Mar. Biol. 153, 755–768 (2008).
    Article  Google Scholar 

    38.
    Riegl, B. M., Bruckner, A. W., Rowlands, G. P., Purkis, S. J. & Renaud, P. Red Sea coral reef trajectories over 2 decades suggest increasing community homogenization and decline in coral size. PLoS ONE 7, e38396 (2012).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    39.
    Hughes, T. P. et al. Coral reefs in the Anthropocene. Nature 546, 82–90 (2017).
    CAS  Article  Google Scholar 

    40.
    Global Distribution of Coral Reefs (UNEP-WCMC, WorldFish Centre, WRI & TNC, 2018); https://data.unep-wcmc.org/datasets/

    41.
    Bruno, J. F. & Valdivia, A. Coral reef degradation is not correlated with local human population density. Sci. Rep. 6, 29778 (2016).

    42.
    Bruno, J. Data from: Coral reef degradation is not correlated with local human population density. Dryad Digital Repository https://doi.org/10.5061/dryad.48r68 (2016).

    43.
    Karlson, R. H., Cornell, H. V. & Hughes, T. P. Coral communities are regionally enriched along an oceanic biodiversity gradient. Nature 429, 867–870 (2004).
    CAS  PubMed  Article  Google Scholar 

    44.
    Cornell, H. V., Karlson, R. H. & Hughes, T. P. Scale-dependent variation in coral community similarity across sites, islands, and island groups. Ecology 88, 1707–1715 (2007).
    PubMed  Article  Google Scholar 

    45.
    Cornell, H. V., Karlson, R. H. & Hughes, T. P. Local-regional species richness relationships are linear at very small to large scales in west-central Pacific corals. Coral Reefs 27, 145–151 (2008).
    Article  Google Scholar 

    46.
    Connolly, S. R., Dornelas, M., Bellwood, D. R. & Hughes, T. P. Testing species abundance models: a new bootstrap approach applied to Indo-Pacific coral reefs. Ecology 90, 3138–3149 (2009).
    PubMed  Article  Google Scholar 

    47.
    Reef Habitat Maps (NOAA-NCCOS, accessed 10 November 2017); https://products.coastalscience.noaa.gov/collections/benthic/default.aspx

    48.
    Purkis, S. J. et al. High-resolution habitat and bathymetry maps for 65,000 sq. km of Earth’s remotest coral reefs. Coral Reefs 38, 467–488 (2019).
    Article  Google Scholar 

    49.
    Roelfsema, C., Phinn, S., Jupiter, S., Comley, J. & Albert, S. Mapping coral reefs at reef to reef-system scales, 10s–1000s km2, using object-based image analysis. Int. J. Remote Sens. 34, 6367–6388 (2013).
    Article  Google Scholar 

    50.
    Bürkner, P.-C. brms: an R package for Bayesian multilevel models using Stan. J. Stat. Softw. 80, 1–28 (2017).
    Article  Google Scholar 

    51.
    Warton, D. I. & Hui, F. K. C. The arcsine is asinine: the analysis of proportions in ecology. Ecology 92, 3–10 (2011).
    PubMed  Article  Google Scholar 

    52.
    Marsh, L. M., Bradbury, R. H. & Reichelt, R. E. Determination of the physical parameters of coral distributions using line transect data. Coral Reefs 2, 175–180 (1984).
    Google Scholar 

    53.
    Hughes, T. P. Population dynamics based on individual size rather than age: a general model with a reef coral example. Am. Nat. 123, 778–795 (1984).
    Article  Google Scholar 

    54.
    Hall, V. R. & Hughes, T. P. Reproductive strategies of modular organisms: comparative studies of reef-building corals. Ecology 77, 950–963 (1996).
    Article  Google Scholar 

    55.
    Hughes, T. P., Connolly, S. R. & Keith, S. A. Geographic ranges of reef corals (Cnidaria: Anthozoa: Scleractinia) in the Indo-Pacific. Ecology 94, 1659 (2013).
    Article  Google Scholar 

    56.
    Bar-On, Y. M., Phillips, R. & Milo, R. The biomass distribution on Earth. Proc. Natl Acad. Sci. USA 115, 6506–6511 (2018).
    CAS  PubMed  Article  Google Scholar 

    57.
    van den Hoogen, J. et al. Soil nematode abundance and functional group composition at a global scale. Nature 572, 194–198 (2019).
    PubMed  Article  CAS  Google Scholar 

    58.
    Hubbell, S. P. et al. How many tree species are there in the Amazon and how many of them will go extinct? Proc. Natl Acad. Sci. USA 105, 11498–11504 (2008).
    CAS  PubMed  Article  Google Scholar 

    59.
    Atkinson, A., Siegel, V., Pakhomov, E. A., Jessopp, M. J. & Loeb, V. A re-appraisal of the total biomass and annual production of Antarctic krill. Deep-Sea Res. I 56, 727–740 (2009).
    Article  Google Scholar 

    60.
    Current World Population (Worldometer, accessed 13 May 2020); https://www.worldometers.info/world-population/

    61.
    California Condor Recovery Program: 2017 Annual Population Status (US Fish and Wildlife Service, 2017).

    62.
    Goodrich, J. M. et al. Panthera tigris. The IUCN Red List of Threatened Species 2015 Report number e.T15955A50659951 (IUCN, 2015). More

  • in

    Gharial nesting in a reservoir is limited by reduced river flow and by increased bank vegetation

    1.
    Strayer, D. L. & Dudgeon, D. Freshwater biodiversity conservation: Recent progress and future challenges. Freshw. Sci. 29, 344–358 (2010).
    Google Scholar 
    2.
    Dudgeon, D. et al. Freshwater biodiversity: Importance, threats, status and conservation challenges. Biol. Rev. 81, 163–182 (2006).
    PubMed  Article  Google Scholar 

    3.
    Reid, A. J. et al. Emerging threats and persistent conservation challenges for freshwater biodiversity. Biol. Rev. 94, 849–873 (2019).
    PubMed  Article  Google Scholar 

    4.
    He, F. et al. Freshwater megafauna diversity: Patterns, status and threats. Divers. Distrib. 24, 1395–1404 (2018).
    Article  Google Scholar 

    5.
    He, F. et al. Disappearing giants: A review of threats to freshwater megafauna. WIREs Water 4, e1208 (2017).
    Article  Google Scholar 

    6.
    Nilsson, C. & Berggren, K. Alterations of riparian ecosystems caused by river regulation: Dam operations have caused global-scale ecological changes in riparian ecosystems. How to protect river environments and human needs of rivers remains one of the most important questions of our time. BioScience 50, 783–792 (2000).

    7.
    Nilsson, C., Reidy, C. A., Dynesius, M. & Revenga, C. Fragmentation and flow regulation of the world’s large river systems. Science 308, 405–408 (2005).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    8.
    Nilsson, C. & Svedmark, M. Basic principles and ecological consequences of changing water regimes: Riparian plant communities. Environ. Manag. 30, 468–480 (2002).
    Article  Google Scholar 

    9.
    Lytle, D. A. & Poff, N. L. Adaptation to natural flow regimes. Trends Ecol. Evol. 19, 94–100 (2004).
    PubMed  Article  Google Scholar 

    10.
    Junk, W. J. & Wantzen, K. M. The flood pulse concept: New aspects, approaches and applications-an update. in Proceedings of the Second International Symposium on the Management of Large Rivers for Fisheries (eds. Welcomme, R. L. & Petr, T.) 117–149 (Bangkok: Food and Agriculture Organization and Mekong River Commission, FAO Regional Office for Asia and the Pacific, 2004).

    11.
    Wiens, J. A. Riverine landscapes: Taking landscape ecology into the water. Freshw. Biol. 47, 501–515 (2002).
    Article  Google Scholar 

    12.
    Benda, L. et al. The network dynamics hypothesis: How channel networks structure riverine habitats. Bioscience 54, 413–427 (2004).
    Article  Google Scholar 

    13.
    Poff, N. L. Beyond the natural flow regime? Broadening the hydro-ecological foundation to meet environmental flows challenges in a non-stationary world. Freshw. Biol. 63, 1011–1021 (2018).
    Article  Google Scholar 

    14.
    Castro, J. M. & Thorne, C. R. The stream evolution triangle: Integrating geology, hydrology, and biology. River Res. Appl. 35, 315–326 (2019).
    Article  Google Scholar 

    15.
    Palmer, M. & Ruhi, A. Linkages between flow regime, biota, and ecosystem processes: Implications for river restoration. Science 365, eaaw2087 (2019).

    16.
    Van Looy, K. et al. The three Rs of river ecosystem resilience: Resources, recruitment, and refugia. River Res. Appl. 35, 107–120 (2019).
    Article  Google Scholar 

    17.
    Braulik, G. T., Arshad, M., Noureen, U. & Northridge, S. P. Habitat fragmentation and species extirpation in freshwater ecosystems; causes of range decline of the Indus River Dolphin (Platanista gangetica minor). PLoS ONE 9, e101657 (2014).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    18.
    Lang, J., Chowfin, S. & Ross, J. P. Gavialis gangeticus. in The IUCN Red List of Threatened Species 2019: e.T8966A3148543 (2019). https://doi.org/10.2305/IUCN.UK.2019-1.RLTS.T8966A3148543.en.

    19.
    He, F. et al. The global decline of freshwater megafauna. Glob. Change Biol. 25, 3883–3892 (2019).
    ADS  Article  Google Scholar 

    20.
    Khanal, G. et al. Irrigation demands aggravate fishing threats to river dolphins in Nepal. Biol. Conserv. 204, 386–393 (2016).
    Article  Google Scholar 

    21.
    Paudel, S., Timilsina, Y. P., Lewis, J., Ingersoll, T. & Jnawali, S. R. Population status and habitat occupancy of endangered river dolphins in the Karnali River system of Nepal during low water season. Mar. Mammal Sci. 31, 707–719 (2015).
    Article  Google Scholar 

    22.
    Whitaker, R. & Basu, D. The Gharial (Gavialis gangeticus): A review. J. Bombay Nat. Hist. Soc. 79, 531–548 (1983).
    Google Scholar 

    23.
    Vesipa, R., Camporeale, C. & Ridolfi, L. Effect of river flow fluctuations on riparian vegetation dynamics: Processes and models. Adv. Water Resour. 110, 29–50 (2017).
    ADS  CAS  Article  Google Scholar 

    24.
    Merritt, D. M. & Cooper, D. J. Riparian vegetation and channel change in response to river regulation: a comparative study of regulated and unregulated streams in the Green River Basin, USA. Regul. Rivers Res. Mgmt. 16, 543–564 (2000).
    Article  Google Scholar 

    25.
    Latterell, J. J., Bechtold, J. S., O’keefe, T. C., Pelt, R. V. & Naiman, R. J. Dynamic patch mosaics and channel movement in an unconfined river valley of the Olympic Mountains. Freshw. Biol. 51, 523–544 (2006).

    26.
    Braatne, J. H., Rood, S. B., Goater, L. A. & Blair, C. L. Analyzing the impacts of dams on riparian ecosystems: A review of research strategies and their relevance to the Snake River through Hells Canyon. Environ. Manag. 41, 267–281 (2008).
    ADS  Article  Google Scholar 

    27.
    Merritt, D. M., Scott, M. L., LeRoy, P. N., Auble, G. T. & Lytle, D. A. Theory, methods and tools for determining environmental flows for riparian vegetation: Riparian vegetation-flow response guilds. Freshw. Biol. 55, 206–225 (2010).
    Article  Google Scholar 

    28.
    Poff, N. L. & Zimmerman, J. K. Ecological responses to altered flow regimes: A literature review to inform the science and management of environmental flows. Freshw. Biol. 55, 194–205 (2010).
    Article  Google Scholar 

    29.
    Miller, K. A., Webb, J. A., de Little, S. C. & Stewardson, M. J. Environmental flows can reduce the encroachment of terrestrial vegetation into river channels: A systematic literature review. Environ. Manag. 52, 1202–1212 (2013).
    ADS  Article  Google Scholar 

    30.
    Tonkin, J. D., Merritt, D. M., Olden, J. D., Reynolds, L. V. & Lytle, D. A. Flow regime alteration degrades ecological networks in riparian ecosystems. Nat. Ecol. Evol. 2, 86–93 (2018).
    PubMed  Article  Google Scholar 

    31.
    Liro, M. Dam reservoir backwater as a field-scale laboratory of human-induced changes in river biogeomorphology: A review focused on gravel-bed rivers. Sci. Total Environ. 651, 2899–2912 (2019).
    ADS  CAS  PubMed  Article  Google Scholar 

    32.
    Volke, M. A., Johnson, W. C., Dixon, M. D. & Scott, M. L. Emerging reservoir delta-backwaters: Biophysical dynamics and riparian biodiversity. Ecol. Monogr. 89, e01363 (2019).
    Article  Google Scholar 

    33.
    Choudhury, S. Seasonal habitat use and resource partitioning between two sympatric crocodilian populations (Gavialis gangeticus & Crocodylus palustris) in Katerniaghat Wildlife Sanctuary, India. Master’s thesis submitted to Saurashtra University, Rajkot, Gujarat, India (2011)

    34.
    MacClune, K. et al. Urgent case for recovery: What we can learn from the August 2014 Karnali River floods in Nepal. in Technical Report. Zurich Insurance Group Ltd, Zurich, ISET-International, Boulder 1–44 (2015).

    35.
    Lang, J. W. & Kumar, P. Behavioral ecology of gharial on the chambal river, India. in Crocodiles. Proceedings of the 22nd Working Meeting of the IUCN-SSC Specialist Group. 42–52 (IUCN, Gland, 2013)

    36.
    Lang, J. W. & Kumar, P. Chambal gharial ecology project-2016 update. in Crocodiles. Proceedings of the 24th Working Meeting of the IUCN-SSC Specialist Group. 136–148 (IUCN, Gland, 2016)

    37.
    Gladfelter, S. R. Training rivers, Training people: Interrogating the making of disasters and the politics of response in Nepal’s lower Karnali River basin. Master’s thesis, University of Colorado (2017). https://floodresilience.net/resources/item/training-rivers-training-people-interrogating-the-making-of-disasters-and-the-politics-of-response-in-nepals-lower-karnali-river-basin.

    38.
    Kolbe, J. J. & Janzen, F. J. Impact of nest-site selection on nest success and nest temperature in natural and disturbed habitats. Ecology 83, 269–281 (2002).
    Article  Google Scholar 

    39.
    Brown, G. P. & Shine, R. Maternal nest-site choice and offspring fitness in a tropical snake (Tropidonophis mairii, Colubridae). Ecology 85, 1627–1634 (2004).
    Article  Google Scholar 

    40.
    López-Luna, M. A., Hidalgo-Mihart, M. G., Aguirre-León, G., González-Ramón, M. D. C. & Rangel-Mendoza, J. A. Effect of nesting environment on incubation temperature and hatching success of Morelet’s crocodile (Crocodylus moreletii) in an urban lake of Southeastern Mexico. J. Therm. Boil. 49, 66–73 (2015).
    Article  Google Scholar 

    41.
    Calverley, P. M. & Downs, C. T. The past and present nesting ecology of Nile crocodiles in Ndumo Game Reserve, South Africa: Reason for concern?. J. Herpetol. 51, 19–26 (2017).
    Article  Google Scholar 

    42.
    Somaweera, R., Brien, M. L., Platt, S. G., Manolis, C. & Webber, B. L. Direct and indirect interactions with vegetation shape crocodylian ecology at multiple scales. Freshw. Biol. 64, 257–268 (2019).
    Google Scholar 

    43.
    Lang, J. W. & Andrews, H. V. Temperature-dependent sex determination in crocodilians. J. Exp. Zool. 270, 28–44 (1994).
    Article  Google Scholar 

    44.
    Andrews, H. V. & Whitaker, N. Captive breeding and reproductive biology of the Indian Gharial Gavialis gangeticus (Gmelin). in Crocodiles. Proceedings of the 17th Working Meeting of the IUCN-SSC Crocodile Specialist Group. 401–411 (IUCN, Gland, 2004).

    45.
    Rhen, T. & Lang, J. W. Phenotypic effects of incubation temperature in reptiles. In Temperature-dependent sex determination in vertebrates (eds. Valenzuela, N. & Lance, V. A.) 90–98 (Smithsonian Books, Washington, 2004).

    46.
    Singh, V. P. Status of the gharial in Uttar Pradesh and its rehabilitation. J. Bombay Nat. Hist. Soc. 75(3), 668–683 (1979).
    ADS  Google Scholar 

    47.
    Basu, D. The gharial of Katerniaghat. Sanctuary 11, 36–43 (1991).
    Google Scholar 

    48.
    Srivastava, A. K. The biology of Indian gharial, Gavialis gangeticus, with special reference to its behaviour. PhD thesis submitted at University of Lucknow, Uttar Pradesh, India (1981).

    49.
    Singh, V. P. Evaluation of gharial rehabilitation U.P. forestry project. Report prepared for biodiversity research, aided by World bank. 1–49 (2003).

    50.
    Andrews, H. V. Status of the Indian gharial, conservation action and assessment of key locations in North India. Unpublished report to Cleveland Metro Park. 1–8 (2006).

    51.
    Whitaker, R. The gharial: Going extinct again. Iguana 14, 24–33 (2007).
    Google Scholar 

    52.
    Chaudhari, S. Gharial reproduction and mortality. Iguana 15, 150–153 (2008).
    Google Scholar 

    53.
    Converse L. Katerniaghat Gharial Project 2008–2009. Report of Preliminary Findings. A Report to GCA and James Cook University, Australia. 1–8 (2009).

    54.
    Das, A., Basu, D., Converse, L. & Choudhury, S. C. Herpetofauna of Katerniaghat Wildlife Sanctuary, Uttar Pradesh, India. JoTT. 4, 2553–2568 (2012).
    Google Scholar 

    55.
    Choudhary, S., Choudhury, B. C. & Gopi, G. V. Differential response to disturbance factors for the population of sympatric crocodilians (Gavialis gangeticus and Crocodylus palustris) in Katarniaghat Wildlife Sanctuary, India. Aquat. Conserv. 27, 946–952 (2017).
    Article  Google Scholar 

    56.
    Kuussaari, M. et al. Extinction debt: A challenge for biodiversity conservation. Trends Ecol. Evol. 24, 564–571 (2009).
    PubMed  Article  Google Scholar 

    57.
    Figueiredo, L., Krauss, J., Steffan‐Dewenter, I. & Sarmento Cabral, J. Understanding extinction debts: Spatio-temporal scales, mechanisms and a roadmap for future research. Ecography 42, 1973–1990 (2019).

    58.
    Bashyal, A. et al. Gharials (Gavialis gangeticus) in Bardiya National Park of Nepal: Population, habitat, and threats. Aquat. Conserv. (in press).

    59.
    Grill, G. et al. Mapping the world’s free-flowing rivers. Nature 569, 215–221 (2019).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    60.
    Jensen, J. R. Remote Sensing of the Environment: An Earth Resource Perspective (Pearson Prentice Hall, Upper Saddle River, 2007).
    Google Scholar 

    61.
    Cohen, J. A. Coefficient of agreement for nominal scales. Educ. Psychol. Meas. 20, 37–46 (1960).
    Article  Google Scholar 

    62.
    Killick, R., Haynes, K. & Eckley, I. A. Changepoint: An R package for changepoint analysis. R package version 2.2.2 (2016). https://CRAN.R-project.org/package=changepoint

    63.
    Carpenter, S. R. & Kinne, O. Regime Shifts in Lake Ecosystems: Pattern and Variation, Vol. 15 (International Ecology Institute, Oldendorf/Luhe, 2003).

    64.
    Whited, D. C. et al. Climate, hydrologic disturbance, and succession: drivers of floodplain pattern. Ecology 88, 940–953 (2007).
    PubMed  Article  Google Scholar 

    65.
    Heffernan, J. B. Wetlands as an alternative stable state in desert streams. Ecology 89, 1261–1271 (2008).
    PubMed  Article  Google Scholar 

    66.
    R Core Team. R: A Language and Environment for Statistical Computing. (R Foundation for Statistical Computing, Vienna, 2013). http://www.R-project.org/. More

  • in

    Ecology-guided prediction of cross-feeding interactions in the human gut microbiome

    Overview of the GutCP algorithm
    Our approach uses the idea that we can leverage cross-feeding interactions—which comprise knowing the metabolites that each microbial species is capable of consuming and producing—to mechanistically connect the levels of microbes and metabolites in the human gut. Several different mechanistic models in past studies have shown that this is indeed possible18,20,29,36,37. While GutCP is generalizable and can be used with any of these models, in this paper, we use a previously published consumer-resource model20. We use this model because of its context and performance: it is built specifically for the human gut and is best able to explain the experimentally measured species composition of the gut microbiome with its resulting metabolic environment, or fecal metabolome (compared with other state-of-the-art methods, such as ref. 29). To predict the metabolome from the microbiome, it relies on a manually curated set of known cross-feeding interactions9. It then uses these known interactions to follow the stepwise flow of metabolites through the gut. At each step (ecologically, at each trophic level), the metabolites available to the gut are utilized by microbial species that are capable of consuming them, and a fraction of these metabolites are secreted as metabolic byproducts. These byproducts are then available for consumption by another set of species in the next trophic level. After several such steps, the metabolites that are left unconsumed constitute the fecal metabolome.
    We hypothesized that adding new, yet-undiscovered cross-feeding interactions would improve our ability to predict the levels of metabolites with our mechanistic and causal model. Specifically, we predict that the set of undiscovered interactions resulting in the most accurate and optimal improvement in predictions would be the most likely candidates for true cross-feeding interactions. Inferring such an optimal set of new cross-feeding interactions or reactions is the main logic driving GutCP. In what follows, we sometimes refer to cross-feeding reactions (i.e., metabolite consumption or production by microbes) as “links” in an overall cross-feeding network of the gut microbiome, whose nodes are microbes and metabolites (Fig. 1a; metabolites in blue, microbes in orange); the links themselves are directed edges connecting the nodes. Links can be of two types: consumption or nutrient uptake reactions (from nutrients to microbes) and production or nutrient secretion reactions (from microbes to their metabolic byproducts).
    Fig. 1: Overview of the GutCP algorithm.

    a Schematic of the original set of known cross-feeding interactions (top) and bar plot of the prediction error for each metabolite and microbe (bottom). The cross-feeding interactions are represented as a network, whose nodes are either metabolites (cyan circles) or microbial species (orange ellipses), and directed links represent the abilities of different species to consume (red arrows) and produce (blue arrows) individual metabolites. b GutCP adds a new consumption link (red) and production link (blue) as added links reduce the prediction errors for metabolites and microbes.

    Full size image

    The salient aspects of our method are outlined in Fig. 1. We start with the known set of consumption and production links that were originally used by the model; these links are known from direct experiments and represent a ground-truth dataset or original cross-feeding network9. These are shown in Fig. 1a through the pink and blue arrows connecting nutrients 1 through 6 with microbes (a) through (c). For each sample, using only the species abundance from the microbiome, we use the model to quantitatively estimate the microbiome’s species and metabolomic composition. Briefly, we assume that a defined set of polysaccharides, common to human diets, are available as the nutrient intake to the gut (nutrients 1 and 4 in Fig. 1a). We calculate the microbiome and metabolome profiles separately for each individual, which contain a different set of microbial species in their guts. At the first trophic level, all microbial species that are capable of using the polysaccharides (indicated by the pink arrows in Fig. 1a) consume each of them in proportion to their abundances (microbes a, b, and c in Fig. 1a). They subsequently secrete a fixed fraction of the consumed nutrients as metabolic byproducts; every species at this trophic level secretes all the metabolic byproducts it is known to secrete (blue arrows in Fig. 1a) in equal proportion (nutrients 2–6 in Fig. 1a). At the next trophic level, all species detected in the individual’s gut which can consume the newly secreted byproducts consume them as nutrients, secreting a new set of byproducts, and this continues for four trophic levels (not shown in Fig. 1a for simplicity). At the end of this process, all metabolites which remain unconsumed by the community comprise the metabolome of the individual and the microbial species which consume nutrients and grow comprise the microbiome of the individual (for a complete description, see “Methods” and previous work20).
    For each metabolite and microbial species, there can be two kinds of prediction errors, or biases: individual (the sample-specific difference between predicted and measured levels) and systematic (average difference across all samples). We focused on the “systematic bias” for each metabolite and microbial species: the average deviation of the predicted levels from the measured levels across all samples in our dataset (Fig. 1a, bottom). The systematic bias for each metabolite and microbe tells us whether our model generally tends to predict their level to be greater than observed (overpredicted), less than observed (underpredicted), or neither (well-predicted). We assume that metabolites and microbes with a large systematic bias are most likely to harbor missing consumption or production links that are relevant across many samples. We prioritize adding links to them in proportion to their systematic biases.
    After measuring the systematic bias for each metabolite and microbe, GutCP proceeds in discrete steps (Fig. 1a, b). At each step, we attempt to add a new link to the current cross-feeding network. This new link is chosen randomly from the entire set of combinatorially possible links (see “Methods”; for S species, M metabolites, and two kinds of links (consumption and production), there are a total of 2SM combinatorially possible links). We accept this link—keeping it in the current network—if it leads to an overall improvement in the agreement between the predicted and measured levels of microbes and metabolites. We repeat the process of adding new links—accepting or rejecting them—until the improvements in the levels of metabolites and microbes became insignificant. Overall, GutCP can add several links to improve the agreement between the predicted and measured levels of microbes and metabolites (in Fig. 1a, b, bottom, adding the extra red and blue link at the top results in improved predictions for metabolite (1), metabolite (3), and microbe (b). Figure 2a shows how the cross-feeding network improves over a typical GutCP run via the red trajectory, starting from the original network (Fig. 2a, top left) to the final network state (Fig. 2a, bottom right). Trajectories from 100 other runs are shown in gray. GutCP repeatably reduces both the error of the metabolome predictions (y axis; measured as ({text{log}}_{10}(frac{,text{pred}-text{meas}}{text{measurement},}))) and improves the correlation between the predicted and measured metabolomes (x axis).
    Fig. 2: Improvement in predictions using GutCP.

    a Improvement in log error (({text{log}}_{10}(frac{,text{pred}-text{meas}}{text{measurement},}))) and the correlation between the prediction and measured fecal metabolome during 100 typical runs of the GutCP algorithm. The gray point at the top left indicates the performance of the original cross-feeding network of Ref. 9, and the black points at the bottom right, that of improved networks predicted using GutCP. A trajectory example, highlighting how performance improves over a GutCP run, is shown in red, and others are shown in gray. b Rarefaction curve showing the number of unique cross-feeding interactions discovered by GutCP over 100 runs of the algorithm. c Prevalence of links, i.e., the number of GutCP runs in which they repeatedly appeared (red dots; total 100 runs) and for comparison, a corresponding binomial distribution with the same mean (black dotted line). P values for different prevalences are estimated using the one-sided binomial test.

    Full size image

    Cross-validating the newly predicted interactions
    To test if the cross-feeding interactions predicted by GutCP are generalizable to unknown datasets, we performed fourfold cross-validation. We used a sample -omics dataset of the gut microbiome and metabolome sampled from 41 human individuals, comprising 221 metabolites and 72 microbial species (data from ref. 38). We split our -omics dataset into two subsets: training (three-fourths of the individuals) and test (one-fourth of the individuals) subsets. We then ran GutCP on the training subset to discover new interactions and added them to the ground-truth interactions taken from ref. 9. Doing so resulted in a network of cross-feeding interactions learned only from the training subset of the data. Finally, we evaluated the improvement in accuracy of metabolome predictions resulting from the trained network on the unseen, test subset of the data. We repeated this process three times, each time splitting the full dataset into a training subset (with a randomly chosen three-fourths of the individuals) and test subset (with the remaining one-fourth of the individuals); finally, we calculated the average improvement in prediction accuracy over all four splits.
    We found that both the training and test set performances after using the links predicted by GutCP were significantly better than the baseline given by the original cross-feeding network (Table 1). Specifically, both measures of model performance, namely the logarithmic error and the average correlation, improved by 64% and 20%, respectively, after adding GutCP’s discovered interactions. In addition, the test set performance was comparable to the training set performance (6% difference; Table 1). This suggests that the cross-feeding interactions inferred by GutCP are not likely to be a result of over-fitting.
    Table 1 Cross-validating the newly predicted interactions.
    Full size table

    Building a consensus-based atlas of predicted cross-feeding interactions
    Having confirmed that GutCP is unlikely to over-fit data, we pooled the entire sample dataset of 41 individuals and ran 100 independent instances of our prediction algorithm on it; we verified that incorporating more instances did not qualitatively affect our results (Fig. 2b shows a rarefaction curve, which highlights the number of new links discovered by GutCP as we perform more runs the algorithm). Each run of the algorithm resulted in an average of 140 newly predicted cross-feeding interactions. Then, based on consensus from many runs, we assigned a confidence level to each predicted interaction, namely what fraction of GutCP runs it was discovered in. By calculating a null distribution (Fig. 2c, black), which predicts the fraction of GutCP runs where a random link would be discovered by chance, we assigned a P value to each link and set a threshold at P = 10−3 (Fig. 3c, red; see “Methods” for details). Doing so finally resulted in a complete consensus-based atlas of 293 predicted cross-feeding interactions, which we have provided as a resource for experimental verification in Supplementary Table 1. Figure 3a shows a condensed version of these interactions obtained from the simulation with the best performance (the trajectory example in Fig. 2a with the lowest log error and highest correlation coefficient) in the form of a matrix; specifically, newly added interactions are in dark colors, and old interactions in faded colors. Supplementary Fig. 3 shows a complete version of this matrix. Note that some of the predicted interactions in Fig. 3a are unrealistic, e.g., the production of certain sugars like D-Fructose and D-Sorbitol. Such interactions are unlikely to be predicted in repeated simulations, and thus will not be part of the final consensus set. This illustrates the power of pooling results from several simulations to arrive at a set of highly probable predictions.
    Fig. 3: New cross-feeding interactions predicted by GutCP.

    a Concise matrix representation of the improved cross-feeding network of the gut microbiome predicted by GutCP (the trajectory example in Fig. 2a with the best performance). The rows are metabolites, and columns, microbial species. Faded cells represent the original, known set of cross-feeding interactions, both production (light blue), consumption (light red), and bidirectional links (gray). The new cross-feeding interactions predicted by GutCP are shown in dark colors: production links in dark blue, consumption links in dark red, and bidirectional links in black. b Network of 293 new links predicted by GutCP (with a P value  More

  • in

    Bacteria enhance the production of extracellular polymeric substances by the green dinoflagellate Lepidodinium chlorophorum

    1.
    Siano, R. et al. Citizen participation in monitoring phytoplankton seawater discolorations. Mar. Policy 117, 1–11. https://doi.org/10.1016/j.marpol.2018.01.022 (2018).
    Article  Google Scholar 
    2.
    Elbrächter, M. & Schnepf, E. Gymnodinium chlorophorum, a new, green, bloom-forming dinoflagellate (Gymnodiniales, Dinophyceae) with a vestigial prasinophyte endosymbiont. Phycologia 35, 381–393 (1996).
    Article  Google Scholar 

    3.
    Hansen, G., Botes, L. & De Salas, M. Ultrastructure and large subunit rDNA sequences of Lepidodinium viride reveal a close relationship to Lepidodinium chlorophorum comb. Nov. (=Gymnodinium chlorophorum). Phycol. Res. 55, 25–41. https://doi.org/10.1111/j.1440-1835.2006.00442.x (2007).
    CAS  Article  Google Scholar 

    4.
    Gavalás-Olea, A. et al. 19,19′-diacyloxy signature: An atypical level of structural evolution in carotenoid pigments. Org. Lett. 18, 4642–4645. https://doi.org/10.1021/acs.orglett.6b02272 (2016).
    CAS  Article  PubMed  Google Scholar 

    5.
    Jackson, C., Knoll, A. H., Chan, C. X. & Verbruggen, H. Plastid phylogenomics with broad taxon sampling further elucidates the distinct evolutionary origins and timing of secondary green plastids. Sci. Rep. 8, 1523. https://doi.org/10.1038/s41598-017-18805-w (2018).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    6.
    Kamikawa, R. et al. Plastid genome-based phylogeny pinpointed the origin of the green-colored plastid in the dinoflagellate Lepidodinium chlorophorum. Genome Biol. Evol. 7, 1133–1140. https://doi.org/10.1093/gbe/evv060 (2015).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    7.
    Chapelle, A., Lazure, P. & Ménesguen, A. Modelling eutrophication events in a coastal ecosystem. Sensitivity analysis. Estuar. Coast. Shelf Sci. 39, 529–548. https://doi.org/10.1016/S0272-7714(06)80008-9 (1994).
    ADS  CAS  Article  Google Scholar 

    8.
    Sournia, A. et al. The repetitive and expanding occurrence of a green, bloom-forming dinoflagellate (Dinophyceae) on the coast of France. Cryptogam. Algol. 13, 1–13 (1992).
    Google Scholar 

    9.
    Claquin, P., Probert, I., Lefebvre, S. & Veron, B. Effects of temperature on photosynthetic parameters and TEP production in eight species of marine microalgae. Aquat. Microb. Ecol. 51, 1–11. https://doi.org/10.3354/ame01187 (2008).
    Article  Google Scholar 

    10.
    Alldredge, A. L., Passow, U. & Logan, B. E. The abundance and significance of a class of large, transparent organic particles in the ocean. Deep-Sea Res. 40, 1131–1140. https://doi.org/10.1016/0967-0637(93)90129-Q (1993).
    CAS  Article  Google Scholar 

    11.
    Passow, U. Transparent exopolymer particles (TEP) in aquatic environments. Prog. Oceanogr. 55, 287–333. https://doi.org/10.1016/S0079-6611(02)00138-6 (2002).
    ADS  Article  Google Scholar 

    12.
    Verdugo, P. et al. The oceanic gel phase: A bridge in the DOM-POM continuum. Mar. Chem. 92, 67–85. https://doi.org/10.1016/j.marchem.2004.06.017 (2004).
    CAS  Article  Google Scholar 

    13.
    Azam, F. & Malfatti, F. Microbial structuring of marine ecosystems. Nature 5, 782–791. https://doi.org/10.1038/nrmicro1747 (2007).
    CAS  Article  Google Scholar 

    14.
    Bittar, T. B., Passow, U., Hamaraty, L., Bidle, K. D. & Harvey, E. L. An updated method for the calibration of transparent exopolymer particle measurements. Limnol. Oceanogr. Methods. 16, 621–628. https://doi.org/10.1002/lom3.10268 (2018).
    Article  Google Scholar 

    15.
    Mari, X., Passow, U., Migon, C., Burd, A. B. & Legendre, L. Transparent exopolymer particles: Effects on carbon cycling in the ocean. Prog. Oceanogr. 151, 13–37. https://doi.org/10.1016/j.pocean.2016.11.002 (2017).
    ADS  Article  Google Scholar 

    16.
    Passow, U. et al. The origin of transparent exopolymer particles (TEP) and their role in the sedimentation of particulate matter. Cont. Shelf. Res. 21, 327–346. https://doi.org/10.1016/S0278-4343(00)00101-1 (2001).
    ADS  Article  Google Scholar 

    17.
    Jenkinson, I. R. Oceanographic implications of non-newtonian properties found in phytoplankton cultures. Nature 323, 435–437. https://doi.org/10.1038/323435a0 (1986).
    ADS  Article  Google Scholar 

    18.
    Alldredge, A. L. & Gotschalk, C. C. Direct observations of the mass flocculation of diatom blooms: Characteristics, settling velocities and formation of diatom aggregates. Deep-Sea Res. 36, 159–171. https://doi.org/10.1016/0198-0149(89)90131-3 (1989).
    ADS  CAS  Article  Google Scholar 

    19.
    Schapira, M., McQuaid, C. D. & Froneman, P. W. Free-living and particle-associated prokaryote metabolism in giant kelp forests: Implications for carbon flux in a sub-Antarctic coastal area. Estuar. Coast. Shelf. Sci. 106, 69–79. https://doi.org/10.1016/j.ecss.2012.04.031 (2012).
    ADS  CAS  Article  Google Scholar 

    20.
    Schapira, M., McQuaid, C. D. & Froneman, P. W. Metabolism of free-living particle-associated prokaryotes: Consequences for carbon flux around a Southern Ocean archipelago. J. Mar. Syst. 90, 58–66. https://doi.org/10.1016/j.jmarsys.2011.08.009 (2012).
    Article  Google Scholar 

    21.
    Bhaskar, P.V. & Bhosle, N.B. Microbial extracellular polymeric substances in marine biogeochemical processes. Curr. Sci. 88, 45–53. http://drs.nio.org/drs/handle/2264/89 (2005).

    22.
    Passow, U. & Alldredge, A. L. A dye-binding assay for the spectrophotometric measurement of transparent exopolymer particles (TEP). Limnol. Oceanogr. 40, 1326–1335. https://doi.org/10.4319/lo.1995.40.7.1326 (1995).
    ADS  CAS  Article  Google Scholar 

    23.
    Gärdes, A., Iversen, M. H., Grossart, H. P., Passow, U. & Ullrich, M. S. Diatom-associated bacteria are required for aggregation of Thalassiosira weissflogii. ISME J. 5, 436–445. https://doi.org/10.1038/ismej.2010.145 (2011).
    CAS  Article  PubMed  Google Scholar 

    24.
    Nosaka, Y., Yamashita, Y. & Suzuki, K. Dynamics and origin of transparent exopolymer particles in the Oyashio region of the Western Subarctic Pacific during the spring diatom bloom. Front. Mar. Sci. 4, 1–16. https://doi.org/10.3389/fmars.2017.00079 (2017).
    Article  Google Scholar 

    25.
    Burns, W. G., Marchetti, A. & Ziervogel, K. Enhanced formation of transparent exopolymer particles (TEP) under turbulence during phytoplankton growth. J. Plankton Res. 41, 349–361. https://doi.org/10.1093/plankt/fbz018 (2019).
    CAS  Article  Google Scholar 

    26.
    Riebesell, U., Reigstad, M., Wassmann, P., Noji, T. & Passow, U. On the trophic fate of Phaeocystis pouchetii (hariot): Significance of Phaeocystis-derived mucus for vertical flux. Neth. J. Sea Res. 33, 193–203. https://doi.org/10.1016/0077-7579(95)90006-3 (1995).
    Article  Google Scholar 

    27.
    Alderkamp, A. C., Buma, A. G. J. & van Rijssel, M. The carbohydrates of Phaeocystis and their degradation in the microbial food web. Biogeochemistry 83, 1–3. https://doi.org/10.1007/s10533-007-9078-2 (2007).
    CAS  Article  Google Scholar 

    28.
    Grossart, H. P., Simon, M. & Logan, B. E. Formation of macroscopic organic aggregates (lake snow) in a large lake: The significance of transparent exopolymer particles, phytoplankton, and zooplankton. Limnol. Oceanogr. 42, 1651–1659. https://doi.org/10.4319/lo.1997.42.8.1651 (1997).
    ADS  CAS  Article  Google Scholar 

    29.
    Iuculano, F., Mazuecos, I. P., Reche, I. & Agusti, S. Prochlorococcus as a possible source for transparent exopolymer particles (TEP). Front. Microbiol. 8, 1–11. https://doi.org/10.3389/fmicb.2017.00709 (2017).
    Article  Google Scholar 

    30.
    Thornton, D. C. O. Dissolved organic matter (DOM) release by phytoplankton in the contemporary and future ocean. Eur. J. Phycol. 49, 20–46. https://doi.org/10.1080/09670262.2013.875596 (2014).
    CAS  Article  Google Scholar 

    31.
    Zhang, Z. et al. The fate of marine bacterial exopolysaccharide in natural marine microbial communities. PLoS One 10, 1–16. https://doi.org/10.1371/journal.pone.0142690 (2015).
    CAS  Article  Google Scholar 

    32.
    Xiao, R. & Zheng, Y. Overview of microalgal extracellular polymeric substances (EPS) and their applications. Biotechnol. Adv. 34, 1225–1244. https://doi.org/10.1016/j.biotechadv.2016.08.004 (2016).
    CAS  Article  PubMed  Google Scholar 

    33.
    Thavasi, R. & Banat, I. M. Biosurfactant and bioemulsifiers from marine sources. In Biosurfactants: Research Trends and Applications, ***Chap 5 (eds Mulligan, C. N. et al.) 125–146 (CRC Press, Boca Raton, 2014).
    Google Scholar 

    34.
    Decho, A. W. & Gutierrez, T. Microbial extracellular polymeric substances (EPSs) in ocean systems. Front. Microbiol. 8, 1–28. https://doi.org/10.3389/fmicb.2017.00922 (2017).
    Article  Google Scholar 

    35.
    Parker, C. The effect of environmental stressors on biofilm formation of Chlorella vulgaris. Master thesis Appalachian State University (2013).

    36.
    Zhou, J., Mopper, K. & Passow, U. The role of surface-active carbohydrates in the formation of transparent exopolymer particles by bubble adsorption of seawater. Limnol. Oceanogr. 43, 1860–1871. https://doi.org/10.4319/lo.1998.43.8.1860 (1998).
    ADS  CAS  Article  Google Scholar 

    37.
    Fukao, T., Kimoto, K. & Kotani, Y. Production of transparent exopolymer particles by four diatom species. Fish Sci. 76, 755–760. https://doi.org/10.1007/s12562-010-0265-z (2010).
    CAS  Article  Google Scholar 

    38.
    Seebah, S., Fairfield, C., Ullrich, M. S. & Passow, U. Aggregation and sedimentation of Thalassiosira weissflogii (diatom) in a warmer and more acidified Future Ocean. PLoS One 9, 1–9. https://doi.org/10.1371/journal.pone.0112379 (2014).
    CAS  Article  Google Scholar 

    39.
    Staats, N., Stal, L. J. & Mur, L. R. Exopolysaccharide production by the epipelic diatom Cylindrotheca fusiformis: Effects of nutrient conditions. J. Exp. Mar. Biol. Ecol. 249, 13–27. https://doi.org/10.1016/S0022-0981(00)00166-0 (2000).
    CAS  Article  PubMed  Google Scholar 

    40.
    Underwood, G. J. C., Boulcott, M., Raines, C. A. & Waldron, K. Environmental effects on exopolymer production by marine benthic diatoms: Dynamics, changes in composition, and pathways of production. J. Phycol. 40, 293–304. https://doi.org/10.1111/j.1529-8817.2004.03076.x (2004).
    CAS  Article  Google Scholar 

    41.
    Engel, A. et al. Impact of CO2 enrichment on organic matter dynamics during nutrient induced coastal phytoplankton blooms. J. Plankton Res. 36, 641–657. https://doi.org/10.1093/plankt/fbt125 (2014).
    CAS  Article  Google Scholar 

    42.
    Thornton, D. C. O. & Chen, J. Exopolymer production as a function of cell permeability and death in a diatom (Thalassiosira weissflogii) and a cyanobacterium (Synechococcus elongatus). J. Phycol. 53, 245–260. https://doi.org/10.1111/jpy.12470 (2017).
    CAS  Article  PubMed  Google Scholar 

    43.
    Sugimoto, K., Fukuda, H., Abdul Baki, M. & Koike, I. Bacterial contribution to formation of transparent exopolymer particles (TEP) and seasonal trends in coastal waters of Sagami Bay, Japan. Aquat. Microb. Ecol. 46, 31–41. https://doi.org/10.3354/ame046031 (2007).
    Article  Google Scholar 

    44.
    Gordillo, F. J. L., Jiménez, C., Chavarria, J. & Niell, F. X. Photosynthetic acclimation to photon irradiance and its relation to chlorophyll fluorescence and carbon assimilation in the halotolerant green alga Dunaliella viridis. Photosynth. Res. 68, 225–235. https://doi.org/10.1023/a:1012969324756 (2001).
    CAS  Article  PubMed  Google Scholar 

    45.
    Ekelund, N. G. A. & Aronsson, K. A. Changes in chlorophyll a fluorescence in Euglena gracilis and Chlamydomonas reinhardii after exposure to wood-ash. Environ. Exp. Bot. 59, 92–98. https://doi.org/10.1016/j.envexpbot.2005.10.004 (2007).
    CAS  Article  Google Scholar 

    46.
    Cole, J. J. Interactions between bacteria and algae in aquatic ecosystems. Ann. Rev. Ecol. Syst. 13, 291–314. https://doi.org/10.1146/annurev.es.13.110182.001451 (1982).
    Article  Google Scholar 

    47.
    Joint, I. et al. Competition for inorganic nutrients between phytoplankton and bacterioplankton in nutrient manipulated mesocosms. Aquat. Microb. Ecol. 29, 145–159. https://doi.org/10.3354/ame029145 (2002).
    Article  Google Scholar 

    48.
    Amin, S. A., Parker, M. S. & Armbrust, E. V. Interactions between diatoms and bacteria. Microbiol. Mol. Biol. Rev. 76, 667–684. https://doi.org/10.1128/MMBR.00007-12 (2012).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    49.
    Ramanan, R., Kim, B. H., Cho, D. H., Oh, H. M. & Kim, H. S. Algae-bacteria interactions: Evolution, ecology and emerging applications. Biotechnol. Adv. 34, 14–29. https://doi.org/10.1016/j.biotechadv.2015.12.003 (2016).
    CAS  Article  PubMed  Google Scholar 

    50.
    Ray, S. & Bagchi, S. N. Nutrients and pH regulate algicide accumulation in cultures of the cyanobacterium Oscillatoria laetevirens. New Phytol. 149, 455–460. https://doi.org/10.1046/j.1469-8137.2001.00061.x (2001).
    CAS  Article  Google Scholar 

    51.
    Oremland, R. S. & Capone, D. G. Use of “specific” inhibitors in biogeochemistry and microbial ecology. Adv. Microb. Ecol. 10, 285–383. https://doi.org/10.1007/978-1-4684-5409-3_8 (1988).
    CAS  Article  Google Scholar 

    52.
    Middelburg, J. J. & Nieuwenhuize, J. Nitrogen uptake by heterotrophic bacteria and phytoplankton in the nitrate-rich Thames estuary. Mar. Ecol. Prog. Ser. 203, 13–21. https://doi.org/10.3354/meps203013 (2000).
    ADS  CAS  Article  Google Scholar 

    53.
    Mulholland, M. R., Rocha, A. M. & Boncillo, G. E. Incorporation of leucine and thymidine by estuarine phytoplankton: Implications for bacteria productivity estimates. Estuar. Coasts 34, 310–325. https://doi.org/10.1007/s12237-010-9366-2 (2010).
    CAS  Article  Google Scholar 

    54.
    Prieto, A. et al. Assessing the role of phytoplankton–bacterioplankton coupling in the response of microbial plankton to nutrient additions. J. Plankton Res. 38, 55–63. https://doi.org/10.1093/plankt/fbv101 (2016).
    CAS  Article  Google Scholar 

    55.
    Dakhama, A., de la Noüe, J. & Lavoie, M. C. Isolation and identification of antialgal substances produced by Pseudomonas aeruginosa. J. Appl. Phycol. 5, 297–306. https://doi.org/10.1007/BF02186232 (1993).
    CAS  Article  Google Scholar 

    56.
    Bowman, L. P. Bioactive compound synthetic capacity and ecological significance of marine bacterial genus Pseudoalteromonas. Mar. Drugs 5, 220–241. https://doi.org/10.3390/md504220 (2007).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    57.
    Meseck, S. L., Smith, B. C., Wikfors, G. H., Alix, J. H. & Kapareiko, D. Nutrient interactions between phytoplankton and bacterioplankton under different carbon dioxide regimes. J. Appl. Phycol. 19, 229–237. https://doi.org/10.1007/s10811-006-9128-5 (2007).
    CAS  Article  Google Scholar 

    58.
    Guerrini, F., Mazzotti, A., Boni, L. & Pistocchi, R. Bacterial-algal interactions in polysaccharide production. Aquat. Microb. Ecol. 15, 247–253. https://doi.org/10.3354/ame015247 (1998).
    Article  Google Scholar 

    59.
    Lu, X. et al. A marine algicidal Thalassiosira and its active substance against the harmful algal bloom species Karenia mikimotoi. Appl. Microbiol. Biotechnol. 100, 5131–5139. https://doi.org/10.1007/s00253-016-7352-8 (2016).
    CAS  Article  PubMed  Google Scholar 

    60.
    Li, Y. et al. Chitinase producing bacteria with direct algicidal activity on marine diatoms. Sci. Rep. 6, 1–13. https://doi.org/10.1038/srep21984 (2016).
    CAS  Article  Google Scholar 

    61.
    Li, Y. et al. The first evidence of deinoxanthin from Deinococcus sp Y35 with strong algicidal effect on the toxic dinoflagellate Alexandrium tamarense. J. Hazard. Mater. 290, 87–95. https://doi.org/10.1016/j.jhazmat.2015.02.070 (2015).
    ADS  CAS  Article  PubMed  Google Scholar 

    62.
    Lovejoy, C., Bowman, J. P. & Hallegraeff, G. M. Algicidal effects of a novel marine Pseudomonas isolate (class Proteobacteria, gamma subdivision) on harmful algal bloom species of the genera Chattonella, Gymnodinium and Heterosigma. Appl. Environ. Microbiol. 64, 2806–2813 (1998).
    CAS  Article  Google Scholar 

    63.
    Honsell, G. & Talarico, L. Gymnodinium chlorophorum (Dinophyceae) in the Adriatic Sea: Electron microscopical observations. Bot. Mar. 47, 152–166. https://doi.org/10.1515/BOT.2004.016 (2004).
    Article  Google Scholar 

    64.
    Iriarte, J. L., Quiñones, R. A. & González, R. R. Relationship between biomass and enzymatic activity of a bloom-forming dinoflagellate (Dinophyceae) in southern Chile (41°S): A field approach. J. Plankton. Res. 27, 159–161. https://doi.org/10.1093/plankt/fbh167 (2005).
    CAS  Article  Google Scholar 

    65.
    Gárate-Lizárraga, I., Muñetón-Gómez, M. S., Pérez-Cruz, B. & Díaz-Ortíz, J. A. Bloom of Gonyaulax spinifera (Dinophyceae: Gonyaulacales) in Ensenada de la Paz Lagoon, Gulf of California. CICIMAR Oceán. 29, 1–18 (2014).
    Google Scholar 

    66.
    McCarthy, P.M. Census of Australian Marine Dinoflagellates. Australian Biological Resources Study, Canberra. http://www.anbg.gov.au/abrs/Dinoflagellates/index_Dino.html. Accessed 11 July 2013 (2013).

    67.
    Azam, F. & Smith, D. C. Bacterial influence on the variability in the ocean’s biogeochemical state: A mechanistic view. In Particle Analysis in Oceanography. NATO ASI Series (Series G: Ecological Sciences), ***27 (ed. Demers, S.) (Springer, Berlin, 1991). https://doi.org/10.1007/978-3-642-75121-9_9.
    Google Scholar 

    68.
    Smith, D. C., Steward, G. F., Long, R. A. & Azam, F. Bacterial mediation of carbon fluxes during a diatom bloom in a mesocosm. Deep-Sea Res. 442, 75–97. https://doi.org/10.1016/0967-0645(95)00005-B (1995).
    ADS  Article  Google Scholar 

    69.
    Schuster, S. & Herndl, G. J. Formation and significance of transparent exopolymeric particles in the northern Adriatic Sea. Mar. Ecol. Prog. Ser. 124, 227–236. https://doi.org/10.3354/meps124227 (1995).
    ADS  Article  Google Scholar 

    70.
    Engel, A. & Passow, U. Carbon and nitrogen content of transparent exopolymer particles (TEP) in relation to their Alcian Blue adsorption. Mar. Ecol. Prog. Ser. 219, 1–10. https://doi.org/10.3354/meps219001 (2001).
    ADS  CAS  Article  Google Scholar 

    71.
    Hasui, M., Matsuda, M., Okutani, K. & Shigeta, S. In vitro antiviral activities of sulfated polysaccharides from a marine microalga (Cochlodinium polykrikoides) against human immunodeficiency virus and other enveloped viruses. Int. J. Biol. Macromol. 17, 293–297. https://doi.org/10.1016/0141-8130(95)98157-T (1995).
    CAS  Article  PubMed  Google Scholar 

    72.
    Yim, J. H., Kim, S. J., Ahn, S. H. & Lee, H. K. Characterization of a novel bioflocculant, p-KG03, from a marine dinoflagellate, Gyrodinium impudicum KG03. Bioresour. Technol. 98, 361–367. https://doi.org/10.1016/j.biortech.2005.12.021 (2007).
    CAS  Article  PubMed  Google Scholar 

    73.
    Mandal, S. K., Singh, R. P. & Patel, V. Isolation and characterization of exopolysaccharide secreted by a toxic dinoflagellate, Amphidinium carterae Hulburt 1957 and its probable role in harmful algal blooms (HABs). Microb. Ecol. 62, 518–527. https://doi.org/10.1007/s00248-011-9852-5 (2011).
    CAS  Article  PubMed  Google Scholar 

    74.
    Kesaulya, I., Leterme, S. C., Mitchell, J. G. & Seuront, L. The impact of turbulence and phytoplankton dynamics on foam formation, seawater viscosity and chlorophyll concentration in the eastern English Channel. Oceanologia 50, 167–182 (2008).
    Google Scholar 

    75.
    Seuront, L. & Vincent, D. Increased seawater viscosity, Phaeocystis globosa spring bloom and Temora longicornis feeding and swimming behaviours. Mar. Ecol. Prog. Ser. 363, 131–145. https://doi.org/10.3354/meps07373 (2008).
    ADS  CAS  Article  Google Scholar 

    76.
    Seuront, L., Vincent, D. & Mitchell, J. G. Biologically induced modification of seawater viscosity in the Eastern English Channel during a Phaeocystis globosa spring bloom. J. Mar. Syst. 61, 118–133. https://doi.org/10.1016/j.jmarsys.2005.04.010 (2006).
    Article  Google Scholar 

    77.
    Seuront, L. et al. The influence of Phaeocystis globosa on microscale spatial patterns of chlorophylla and bulk-phase seawater viscosity. Biogeochemistry 83, 173–188. https://doi.org/10.1007/s10533-007-9097-z (2007).
    CAS  Article  Google Scholar 

    78.
    Seuront, L. et al. Role of microbial and phytoplankton communities in the control of seawater viscosity off East Antarctica (30–80° E). Deep-Sea Res. 57, 877–886. https://doi.org/10.1016/j.dsr2.2008.09.018 (2010).
    ADS  CAS  Article  Google Scholar 

    79.
    Stoderegger, K. E. & Herndl, G. J. Production of exopolymer particles by marine bacterioplankton under contrasting turbulence conditions. Mar. Ecol. Prog. Ser. 189, 9–16. https://doi.org/10.3354/meps189009 (1999).
    ADS  CAS  Article  Google Scholar 

    80.
    Alunno-Bruscia, M. et al. A single bio-energetics growth and reproduction model for the oyster Crassostrea gigas in six Atlantic ecosystems. J. Sea Res. 66, 340–348. https://doi.org/10.1016/j.seares.2011.07.008 (2011).
    ADS  Article  Google Scholar 

    81.
    Thomas, Y. et al. Global change and climate-driven invasion of the Pacific oyster (Crassostrea gigas) along European coasts: A bioenergetics modelling approach. J. Biogeogr. 43, 568–579. https://doi.org/10.1111/jbi.12665 (2016).
    Article  Google Scholar 

    82.
    Guillard, R. & Hargraves, P. Stichochrysis immobilis is a diatom, not a chrysophyte. Phycologia 32, 234–236. https://doi.org/10.2216/i0031-8884-32-3-234.1 (1993).
    Article  Google Scholar 

    83.
    Scholin, C. A., Herzog, M., Sogin, M. & Anderson, D. M. Identification of group- and strain-specific genetic markers for globally distributed Alexandrium (Dinophyceae). II. Sequence analysis of a fragment of the LSU rRNA gene. J. Phycol. 30, 999–1011. https://doi.org/10.1111/j.0022-3646.1994.00999.x (1994).
    CAS  Article  Google Scholar 

    84.
    Nunn, G. B., Theisen, B. F., Christensen, B. & Arctander, P. Simplicity-correlated size growth of the nuclear 28S ribosomal RNA D3 expansion segment in the crustacean order Isopoda. J. Mol. Evol. 42, 211–223. https://doi.org/10.1007/BF02198847 (1996).
    ADS  CAS  Article  PubMed  Google Scholar 

    85.
    Marie, D., Partensky, F., Jacquet, S. & Vaulot, D. Enumeration and cell cycle analysis of natural populations of marine picoplankton by flow cytometry using the nucleic acid stain SYBR Green I. Appl. Environ. Microbiol. 63, 186–193. https://doi.org/10.1128/aem.63.1.186-193.1997 (1997).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    86.
    Wood, A. M., Everroad, R. C. & Wingard, L. M. Measuring growth rates in microalgal cultures. In Algal Culturing Techniques (ed. Anderson, R. A.) 269–285 (Elsevier, Amsterdam, 2005).
    Google Scholar 

    87.
    Kromkamp, J. C. & Forster, R. M. The use of variable fluorescence measurements in aquatic ecosystems: Differences between multiple and single turnover measuring protocols and suggested terminology. Eur. J. Phycol. 38, 103–112. https://doi.org/10.1080/0967026031000094094 (2003).
    Article  Google Scholar 

    88.
    Aminot, A. & Kérouel, R. Dosage Automatique des Nutriments dans les Eaux Marines: Méthodes en flux Continu (in French) (Ed. Ifremer, Plouzané, 2007).
    Google Scholar 

    89.
    Smith, P. K. et al. Measurement of protein using bicinchoninic acid. Anal. Biochem. 150, 76–85. https://doi.org/10.1016/0003-2697(85)90442-7 (1985).
    CAS  Article  PubMed  Google Scholar 

    90.
    Kamerling, J. P., Gerwig, G. J., Vliegenthart, J. F. G. & Clamp, J. R. Characterization by gas-liquid chromatography mass spectrometry of pertrimethylsilyl methyl glycosides obtained in the methanolysis of glycoproteins and glycolipids. Biochem. J. 151, 491–495. https://doi.org/10.1042/bj1510491 (1975).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    91.
    Laemmli, U. K. Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature 22, 680–685. https://doi.org/10.1038/227680a0 (1970).
    ADS  Article  Google Scholar 

    92.
    Rigouin, C., Delbarre Ladrat, C., Sinquin, C., Colliec-Jouault, S. & Dion, M. Assessment of biochemical methods to detect enzymatic depolymerization of polysaccharides. Carbohydr. Polym. 76, 279–284. https://doi.org/10.1016/j.carbpol.2008.10.022 (2009).
    CAS  Article  Google Scholar 

    93.
    Dubray, G. & Bezard, G. A highly sensitive periodic acid-silver stain for 1,2-diol groups of glycoproteins and polysaccharides in polyacrylamide gels. Anal. Biochem. 119, 325–329. https://doi.org/10.1016/0003-2697(82)90593-0 (1982).
    CAS  Article  PubMed  Google Scholar 

    94.
    Aminot, A. & Kérouel, R. Hydrologie des Écosystèmes Marins: Paramètres et Analyses (in French) (Ed Ifremer, Plouzané, 2004).
    Google Scholar 

    95.
    R Core Team R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna. https://www.R-project.org (2018).

    96.
    Lê, S., Josse, J. & Husson, F. FactoMineR: An R package for multivariate analysis. J. Stat. Softw. 25, 1–18 (2008).
    Article  Google Scholar  More

  • in

    A DNA barcode-based survey of wild urban bees in the Loire Valley, France

    1.
    Hallmann, C. A. et al. More than 75 percent decline over 27 years in total flying insect biomass in protected areas. PLoS ONE 12, e0185809 (2017).
    PubMed  PubMed Central  Article  CAS  Google Scholar 
    2.
    Macgregor, C. J., Williams, J. H., Bell, J. R. & Thomas, C. D. Moth biomass increases and decreases over 50 years in Britain. Nat. Ecol. Evol. 3, 1645–1649 (2019).
    PubMed  Article  Google Scholar 

    3.
    Sánchez-Bayo, F. & Wyckhuys, K. A. G. Worldwide decline of the entomofauna: A review of its drivers. Biol. Conserv. 232, 8–27 (2019).
    Article  Google Scholar 

    4.
    Thomas, C. D., Jones, T. H. & Hartley, S. E. “Insectageddon”: A call for more robust data and rigorous analyses. Glob. Chang. Biol. 25, 1891–1892 (2019).
    ADS  PubMed  Article  Google Scholar 

    5.
    van Klink, R. et al. Meta-analysis reveals declines in terrestrial but increases in freshwater insect abundances. Science (80-) 368, 417–420 (2020).
    ADS  Article  CAS  Google Scholar 

    6.
    Potts, S. G. et al. Global pollinator declines: Trends, impacts and drivers. Trends Ecol. Evol. 25, 345–353 (2010).
    PubMed  Article  Google Scholar 

    7.
    Pérez-Méndez, N. et al. The economic cost of losing native pollinator species for orchard production. J. Appl. Ecol. 57, 599–608 (2020).
    Article  Google Scholar 

    8.
    Porto, R. G. et al. Pollination ecosystem services: A comprehensive review of economic values, research funding and policy actions. Food Secur. 12, 1425–1442 (2020).
    Article  Google Scholar 

    9.
    Winfree, R., Aguilar, R., Vázquez, D. P., LeBuhn, G. & Aizen, M. A. A meta-analysis of bees’ responses to anthropogenic disturbance. Ecology 90, 2068–2076 (2009).
    PubMed  PubMed Central  Article  Google Scholar 

    10.
    Godfray, H. C. J. et al. A restatement of the natural science evidence base concerning neonicotinoid insecticides and insect pollinators. Proc. R. Soc. B Biol. Sci. 281, 20140558 (2014).
    Article  Google Scholar 

    11.
    Fortel, L. et al. Decreasing abundance, increasing diversity and changing structure of the Wild Bee Community (Hymenoptera: Anthophila) along an urbanization gradient. PLoS ONE 9, e104679 (2014).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    12.
    Geslin, B. et al. The proportion of impervious surfaces at the landscape scale structures wild bee assemblages in a densely populated region. Ecol. Evol. 6, 6599–6615 (2016).
    PubMed  PubMed Central  Article  Google Scholar 

    13.
    Geslin, B., Le Féon, V., Kuhlmann, M., Vaissière, B. E. & Dajoz, I. The bee fauna of large parks in downtown Paris, France. Ann. la Société Entomol. Fr. 51, 487–493 (2015).
    Article  Google Scholar 

    14.
    Baldock, K. C. R. et al. A systems approach reveals urban pollinator hotspots and conservation opportunities. Nat. Ecol. Evol. 3, 363–373 (2019).
    PubMed  PubMed Central  Article  Google Scholar 

    15.
    Lerman, S. B., Contosta, A. R., Milam, J. & Bang, C. To mow or to mow less: Lawn mowing frequency affects bee abundance and diversity in suburban yards. Biol. Conserv. 221, 160–174 (2018).
    Article  Google Scholar 

    16.
    Kerr, J. T. et al. Climate change impacts on bumblebees converge across continents. Science 349, 177–180 (2015).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    17.
    Soroye, P., Newbold, T. & Kerr, J. Climate change contributes to widespread declines among bumble bees across continents. Science 367, 685–688 (2020).
    ADS  CAS  PubMed  Article  Google Scholar 

    18.
    McFrederick, Q. S. & LeBuhn, G. Are urban parks refuges for bumble bees Bombus spp. (Hymenoptera: Apidae)?. Biol. Conserv. 129, 372–382 (2006).
    Article  Google Scholar 

    19.
    Hall, D. M. et al. The city as a refuge for insect pollinators. Conserv. Biol. 31, 24–29 (2017).
    PubMed  Article  Google Scholar 

    20.
    Ropars, L., Dajoz, I. & Geslin, B. La ville un désert pour les abeilles sauvages? J. Bot. Soc. Bot. Fr. 79, 29–35 (2017).
    Google Scholar 

    21.
    Falk, S. et al. Evaluating the ability of citizen scientists to identify bumblebee (Bombus) species. PLoS ONE 14, e0218614 (2019).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    22.
    Bloom, E. H. & Crowder, D. W. Promoting data collection in pollinator citizen science projects. Citiz. Sci. Theory Pract. 5(1), 3 https://doi.org/10.5334/cstp.217 (2020).

    23.
    Levé, M., Baudry, E. & Bessa-Gomes, C. Domestic gardens as favorable pollinator habitats in impervious landscapes. Sci. Total Environ. 647, 420–430 (2019).
    ADS  PubMed  Article  CAS  Google Scholar 

    24.
    Mason, L. & Arathi, H. S. Assessing the efficacy of citizen scientists monitoring native bees in urban areas. Glob. Ecol. Conserv. 17, e00561 (2019).
    Article  Google Scholar 

    25.
    Sheffield, C. S. et al. Contribution of DNA barcoding to the study of the bees (Hymenoptera: Apoidea) of Canada: Progress to date. Can. Entomol. 149, 736–754 (2017).
    Article  Google Scholar 

    26.
    Sheffield, C. S., Hebert, P. D. N., Kevan, P. G. & Packer, L. DNA barcoding a regional bee (Hymenoptera: Apoidea) fauna and its potential for ecological studies. Mol. Ecol. Resour. 9, 196–207 (2009).
    CAS  PubMed  Article  Google Scholar 

    27.
    Schmidt, S., Schmid-Egger, C., Morinière, J., Haszprunar, G. & Hebert, P. D. N. DNA barcoding largely supports 250 years of classical taxonomy: Identifications for Central European bees (Hymenoptera, Apoidea partim ). Mol. Ecol. Resour. 15, 985–1000 (2015).
    CAS  PubMed  Article  Google Scholar 

    28.
    Packer, L. & Ruz, L. DNA barcoding the bees (Hymenoptera: Apoidea) of Chile: Species discovery in a reasonably well known bee fauna with the description of a new species of Lonchopria (Colletidae). Genome 60, 414–430 (2017).
    CAS  PubMed  Article  Google Scholar 

    29.
    Tang, M. et al. High-throughput monitoring of wild bee diversity and abundance via mitogenomics. Methods Ecol. Evol. 6, 1034–1043 (2015).
    PubMed  PubMed Central  Article  Google Scholar 

    30.
    Sonet, G. et al. Using next-generation sequencing to improve DNA barcoding: Lessons from a small-scale study of wild bee species (Hymenoptera, Halictidae). Apidologie 49, 671–685 (2018).
    CAS  Article  Google Scholar 

    31.
    Creedy, T. J. et al. A validated workflow for rapid taxonomic assignment and monitoring of a national fauna of bees (Apiformes) using high throughput DNA barcoding. Mol. Ecol. Resour. 20, 40–53 (2020).
    CAS  PubMed  Article  Google Scholar 

    32.
    Gueuning, M. et al. Evaluating next-generation sequencing (NGS) methods for routine monitoring of wild bees: Metabarcoding, mitogenomics or NGS barcoding. Mol. Ecol. Resour. 19, 847–862 (2019).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    33.
    Lanner, J., Curto, M., Pachinger, B., Neumüller, U. & Meimberg, H. Illumina midi-barcodes: Quality proof and applications. Mitochondrial DNA Part A 30, 490–499 (2019).
    CAS  Article  Google Scholar 

    34.
    González-Vaquero, R. A., Roig-Alsina, A. & Packer, L. DNA barcoding as a useful tool in the systematic study of wild bees of the tribe Augochlorini (Hymenoptera: Halictidae). Genome 59, 889–898 (2016).
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    35.
    Gibbs, J. DNA barcoding a nightmare taxon: Assessing barcode index numbers and barcode gaps for sweat bees. Genome 61, 21–31 (2018).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    36.
    Dorey, J. P., Schwarz, M. P. & Stevens, M. I. Review of the bee genus Homalictus Cockerell (Hymenoptera: Halictidae) from Fiji with description of nine new species. Zootaxa 4674, 1–46 (2019).
    Article  Google Scholar 

    37.
    Williams, P. H. et al. Unveiling cryptic species of the bumblebee subgenus Bombus s. str. worldwide with COI barcodes (Hymenoptera: Apidae). Syst. Biodivers. 10, 21–56 (2012).
    Article  Google Scholar 

    38.
    Magnacca, K. N. & Brown, M. J. F. DNA barcoding a regional fauna: Irish solitary bees. Mol. Ecol. Resour. 12, 990–998 (2012).
    CAS  PubMed  Article  Google Scholar 

    39.
    de Waard, J. R. et al. A reference library for Canadian invertebrates with 1.5 million barcodes, voucher specimens, and DNA samples. Sci. Data 6, 308 (2019).
    Article  CAS  Google Scholar 

    40.
    Hua, F. et al. Opportunities for biodiversity gains under the world’s largest reforestation programme. Nat. Commun. 7, 12717 (2016).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    41.
    Gueuning, M., Frey, J. E. & Praz, C. Ultraconserved yet informative for species delimitation: UCEs resolve long-standing systematic enigma in Central European bees. Mol. Ecol. Mec. https://doi.org/10.1111/mec.15629 (2020).
    Article  Google Scholar 

    42.
    Phillips, J. D., French, S. H., Hanner, R. H. & Gillis, D. J. HACSim: An R package to estimate intraspecific sample sizes for genetic diversity assessment using haplotype accumulation curves. PeerJ Comput. Sci. 6, e243 (2020).
    Article  Google Scholar 

    43.
    Phillips, J. D., Gwiazdowski, R. A., Ashlock, D. & Hanner, R. An exploration of sufficient sampling effort to describe intraspecific DNA barcode haplotype diversity: Examples from the ray-finned fishes (Chordata: Actinopterygii). DNA Barcodes 3(1), 66–73 (2015).

    44.
    Phillips, J. D., Gillis, D. J. & Hanner, R. H. Incomplete estimates of genetic diversity within species: Implications for DNA barcoding. Ecol. Evol. 9, 2996–3010 (2019).
    PubMed  PubMed Central  Article  Google Scholar 

    45.
    Muséum national d’Histoire naturelle (ed). 2003-2020. Inventaire National du Patrimoine Naturel. https://inpn.mnhn.fr.

    46.
    Zayed, A., Constantin, ŞA. & Packer, L. Successful biological invasion despite a severe genetic load. PLoS ONE 2, e868 (2007).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    47.
    Lecocq, T. et al. The alien’s identity: Consequences of taxonomic status for the international bumblebee trade regulations. Biol. Conserv. 195, 169–176 (2016).
    Article  Google Scholar 

    48.
    Danforth, B. N. Phylogeny of the bee genus Lasioglossum (Hymenoptera: Halictidae) based on mitochondrial COI sequence data. Syst. Entomol. 24, 377–393 (1999).
    Article  Google Scholar 

    49.
    Hebert, P. D. N. et al. A Sequel to Sanger: Amplicon sequencing that scales. BMC Genom. 19, 219 (2018).
    Article  CAS  Google Scholar 

    50.
    Ratnasingham, S. & Hebert, P. D. N. A DNA-based registry for all animal species: The Barcode Index Number (BIN) system. PLoS ONE 8, e66213 (2013).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    51.
    Carolan, J. C. et al. Colour patterns do not diagnose species: Quantitative evaluation of a DNA barcoded cryptic bumblebee complex. PLoS ONE 7, e29251 (2012).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    52.
    Praz, C., Müller, A. & Genoud, D. Hidden diversity in European bees: Andrena amieti sp. n., a new Alpine bee species related to Andrena bicolor (Fabricius, 1775) (Hymenoptera, Apoidea, Andrenidae). Alp. Entomol. 3, 11–38 (2019).
    Article  Google Scholar 

    53.
    Pauly, A. Abeilles de Belgique et des régions limitrophes (Insecta: Hymenoptera: Apoidea) Famille Halictidae. (Institut royal des sciences naturelles de Belgique, 2019).

    54.
    Gonçalves, R. B. & Oliveira, P. S. Preliminary results of bowl trapping bees (Hymenoptera, Apoidea) in a southern Brazil forest fragment. J. Insect Biodivers. 1, 1–9 (2013).
    Article  Google Scholar 

    55.
    Buri, P., Humbert, J.-Y. & Arlettaz, R. Promoting pollinating insects in intensive agricultural matrices: Field-scale experimental manipulation of hay-meadow mowing regimes and its effects on bees. PLoS One 9(1), e85635 (2014).

    56.
    Rhoades, P. et al. Sampling technique affects detection of habitat factors influencing wild bee communities. J. Insect Conserv. 21, 703–714 (2017).
    Article  Google Scholar 

    57.
    Lettow, M. C. et al. Bee community responses to a gradient of oak savanna restoration practices. Restor. Ecol. 26, 882–890 (2018).
    Article  Google Scholar 

    58.
    Onuferko, T. M., Skandalis, D. A., Cordero, R. L. & Richards, M. H. Rapid initial recovery and long-term persistence of a bee community in a former landfill. Insect Conserv. Divers. 11, 88–99 (2018).
    Article  Google Scholar 

    59.
    Geroff, R. K., Gibbs, J. & McCravy, K. W. Assessing bee (Hymenoptera: Apoidea) diversity of an Illinois restored tallgrass prairie: Methodology and conservation considerations. J. Insect Conserv. 18, 951–964 (2014).
    Article  Google Scholar 

    60.
    Griffin, S. R., Bruninga-Socolar, B., Kerr, M. A., Gibbs, J. & Winfree, R. Wild bee community change over a 26-year chronosequence of restored tallgrass prairie. Restor. Ecol. 25, 650–660 (2017).
    Article  Google Scholar 

    61.
    Ropars, L., Dajoz, I. & Geslin, B. La diversité des abeilles parisiennes. Osmia 7, 14–19 (2018).
    Article  Google Scholar 

    62.
    Portman, Z. M., Bruninga-Socolar, B. & Cariveau, D. P. The state of bee monitoring in the United States: A call to refocus away from bowl traps and towards more effective methods. Ann. Entomol. Soc. Am. 113, 337–342 (2020).
    Article  Google Scholar 

    63.
    Magnacca, K. N. & Brown, M. J. Mitochondrial heteroplasmy and DNA barcoding in Hawaiian Hylaeus (Nesoprosopis) bees (Hymenoptera: Colletidae). BMC Evol. Biol. 10, 174 (2010).

    64.
    Ballare, K. M. et al. Utilizing field collected insects for next generation sequencing: Effects of sampling, storage, and DNA extraction methods. Ecol. Evol. 9, 13690–13705 (2019).
    PubMed  PubMed Central  Article  Google Scholar 

    65.
    Hill, G. E. Mitonuclear coevolution as the genesis of speciation and the mitochondrial DNA barcode gap. Ecol. Evol. 6, 5831–5842 (2016).
    PubMed  PubMed Central  Article  Google Scholar 

    66.
    Ascher, J. S. & Pickering, J.  Life bee species guide and world checklist (Hymenoptera Apoidea Anthophila). http://www.discoverlife.org/mp/20q?guide=Apoidea_species (2020).

    67.
    LaBerge, W. E. A revision of the bees of the genus Andrena of the western hemisphere. Part XI. Minor subgenera and subgeneric key. Trans. Am. Entomol. Soc. 111, 441–567 (1985).
    Google Scholar 

    68.
    Warncke, K. Die Untergattungen der westpalaarktischen Bienengattung Andrena F. Memorias e Estud Muséu Zool. da Univ. Coimbra 307, 1–110 (1968).
    Google Scholar 

    69.
    Amiet, F., Herrmann, M., Müller, A. & Neumeyer, R. Apidae 6: Andrena, Melitturga, Panurginus, Panurgus. Fauna Helv. 26, 1–317 (2010).

    70.
    Michener, C. The bees of the world. (Johns Hopkins University Press, Baltimore, 2000).
    Google Scholar 

    71.
    Michener, C. D. The Social Behavior of the Bees: A Comparative Study (Harvard University Press, Cambridge, 1974).
    Google Scholar 

    72.
    Pauly, A., Noël, G., Sonet, G., Notton, D. G. & Boevé, J.-L. Integrative taxonomy resuscitates two species in the Lasioglossum villosulum complex (Kirby, 1802) (Hymenoptera: Apoidea: Halictidae). Eur. J. Taxon. 541 (2019).

    73.
    Eberle, J., Ahrens, D., Mayer, C., Niehuis, O. & Misof, B. A plea for standardized nuclear markers in metazoan DNA taxonomy. Trends Ecol. Evol. 35, 336–345 (2020).
    PubMed  Article  Google Scholar 

    74.
    Roulston, T. H., Smith, S. A. & Brewster, A. L. A comparison of pan trap and intensive net sampling techniques for documenting a bee (Hymenoptera: Apiformes) Fauna. J. Kansas Entomol. Soc. 80, 179–181 (2007).
    Article  Google Scholar 

    75.
    Westphal, C. et al. Measuring bee diversity in different European habitats and biogeographical regions. Ecol. Monogr. 78, 653–671 (2008).
    Article  Google Scholar 

    76.
    Amiet, F., Herrmann, M., Müller, A. & Neumeyer, R. Apidae 5: Ammobates, Ammobatoides, Anthophora, Biastes, Ceratina, Dasypoda, Epeoloides, Epeolus, Eucera, Macropis, Melecta, Melitta, Nomada, Pasites, Tetralonia, Thyreus, Xylocopa. Fauna Helv. 20, 1–356 (2007).

    77.
    Amiet, F., Herrmann, M., Müller, A. & Neumeyer, R. Apidae 2: Colletes, Dufourea, Hylaeus, Nomia, Nomioides, Rhophitoides, Rophites, Sphecodes, Systropha. Fauna Helv. 4, 1–239 (1999).

    78.
    Amiet, F., Herrmann, M., Müller, A. & Neumeyer, R. Apidae 3: Halictus, Lasioglossum. Fauna Helv. 6, 1–208 (2001).

    79.
    Amiet, F., Herrmann, M., Müller, A. & Neumeyer, R. Apidae 4: Anthidium, Chelostoma, Coelioxys, Dioxys, Heriades, Lithurgus, Megachile, Osmia, Stelis. Fauna Helv. 9, 1–273 (2004).

    80.
    Folmer, O., Black, M., Hoeh, W., Lutz, R. & Vrijenhoek, R. DNA primers for amplification of mitochondrial cytochrome c oxidase subunit I from diverse metazoan invertebrates. Mol. Mar. Biol. Biotechnol. 3, 294–299 (1994).
    CAS  PubMed  Google Scholar 

    81.
    Ratnasingham, S. & Hebert, P. D. N. BOLD: The barcode of life data system. Mol. Ecol. Notes 7, 355–364 (2007).

    82.
    Katoh, K. MAFFT: A novel method for rapid multiple sequence alignment based on fast Fourier transform. Nucleic Acids Res. 30, 3059–3066 (2002).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    83.
    Kimura, M. A simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide sequences. J. Mol. Evol. 16, 111–120 (1980).
    ADS  CAS  PubMed  Article  Google Scholar 

    84.
    Price, M. N., Dehal, P. S. & Arkin, A. P. FastTree 2—Approximately maximum-likelihood trees for large alignments. PLoS ONE 5, e9490 (2010).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    85.
    Letunic, I. & Bork, P. Interactive Tree Of Life (iTOL) v4: Recent updates and new developments. Nucleic Acids Res. 47, W256–W259 (2019).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    86.
    Wickham, H. ggplot2 (Springer, New York, 2009). https://doi.org/10.1007/978-0-387-98141-3.
    Google Scholar 

    87.
    Rozas, J. et al. DnaSP 6: DNA sequence polymorphism analysis of large data sets. Mol. Biol. Evol. 34, 3299–3302 (2017).
    CAS  PubMed  Article  Google Scholar 

    88.
    Nei, M. Molecular evolutionary genetics (Columbia University Press, New York, 1987). More

  • in

    Native and invasive ants affect floral visits of pollinating honey bees in pumpkin flowers (Cucurbita maxima)

    1.
    Vitousek, P. M., D’Antonio, C. M., Loope, L. L., Rejmánek, M. & Westbrooks, R. Introduced species: a significant component of human-caused global change. N. Z. J. Ecol. 21, 1–16 (1997).
    Google Scholar 
    2.
    Courchamp, F. et al. Invasion biology: specific problems and possible solutions. Trends Ecol. Evol. 32, 13–22 (2017).
    PubMed  Article  PubMed Central  Google Scholar 

    3.
    Sanders, N. J., Gotelli, N. J., Heller, N. E. & Gordon, D. M. Community disassembly by an invasive species. PNAS 100, 2474–2477 (2003).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    4.
    Christian, C. E. Consequences of a biological invasion reveal the importance of mutualism for plant communities. Nature 413, 635–639 (2001).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    5.
    Suarez, A. V., McGlynn, T. P. & Tsutsui, N. D. Biogeographic and taxonomic patterns of introduced ants. In Ant ecology (eds Lach, L. et al.) 233–244 (Oxford University Press, Oxford, 2010).
    Google Scholar 

    6.
    Moloney, S. D. & Vanderwoude, C. Potential ecological impacts of red imported fire ants in eastern Australia. J. Agric. Urban Entomol. 20, 131–142 (2003).
    Google Scholar 

    7.
    Rajesh, T. P., Ballullaya, U. P., Unni, A. P., Parvathy, S. & Sinu, P. A. Interactive effects of urbanization and year on invasive and native ant diversity of sacred groves of South India. Urban Ecosyst. 23, 1335–1348 (2020).
    Article  Google Scholar 

    8.
    Hoffmann, B. D., Luque, G. M., Bellard, C., Holmes, N. D. & Donlan, C. J. Improving invasive ant eradication as a conservation tool: a review. Biol. Conserv. 198, 37–49 (2016).
    Article  Google Scholar 

    9.
    Traveset, A. & Richardson, D. M. Biological invasions as disruptors of plant reproductive mutualisms. Trends Ecol. Evol. 21, 208–216 (2006).
    PubMed  Article  PubMed Central  Google Scholar 

    10.
    Carney, S. E., Byerley, M. B. & Holway, D. A. Invasive Argentine ants (Linepithema humile) do not replace native ants as seed dispersers of Dendromecon rigida (Papaveraceae) in California, USA. Oecologia 135, 576–582 (2003).
    ADS  PubMed  Article  PubMed Central  Google Scholar 

    11.
    Styrsky, J. D. & Eubanks, M. D. Ecological consequences of interactions between ants and honeydew-producing insects. Proc. R. Soc. B Biol. Sci. 274, 151–164 (2007).
    Article  Google Scholar 

    12.
    Lach, L. Invasive ants: unwanted partners in ant-plant interactions?. Ann. Mo. Bot. Garden 90, 91–108 (2003).
    Article  Google Scholar 

    13.
    Willmer, P. G. et al. Floral volatiles controlling ant behaviour. Funct. Ecol. 23, 888–900 (2009).
    Article  Google Scholar 

    14.
    Vilamil, N., Boege, K. & Stone, G. N. Testing the distraction hypothesis: do extrafloral nectaries reduce ant-pollinator conflict?. J. Ecol. 107, 1377–1391 (2019).
    Article  Google Scholar 

    15.
    Raine, N., Willmer, P. & Stone, G. Spatial structuring and floral avoidance behavior prevent ant-pollinator conflict in a Mexican antacacia. Ecology 83, 3086–3096 (2002).
    Google Scholar 

    16.
    Weber, M. G., Porturas, L. D. & Keeler, K. H. World list of plants with extrafloral nectaries. www.extrafloralnectaries.org (2015)

    17.
    Dutton, E. M. & Frederickson, M. E. Why ant pollination is rare: new evidence and implications of the antibiotic hypothesis. Arthropod-Plant Interact. 6, 561–569 (2012).
    Article  Google Scholar 

    18.
    Gonzálvez, F. G., Santamaría, L., Corlett, R. T. & Rodríguez-Gironés, M. A. Flowers attract weaver ants that deter less effective pollinators. J. Ecol. 101, 78–85 (2013).
    Article  Google Scholar 

    19.
    Cembrowski, A. R., Tan, M. G., Thomson, J. D. & Frederickson, M. E. Ants and ant scent reduce bumblebee pollination of artificial flowers. Am. Nat. 183, 133–139 (2014).
    PubMed  Article  PubMed Central  Google Scholar 

    20.
    Sinu, P. A. et al. Invasive ant (Anoplolepis gracilipes) disrupts pollination in pumpkin. Biol. Invasions 19, 2599–2607 (2017).
    Article  Google Scholar 

    21.
    Hanna, C. et al. Floral visitation by the Argentine ant reduces bee visitation and plant seed set. Ecology 96, 222–230 (2015).
    PubMed  Article  PubMed Central  Google Scholar 

    22.
    LeVan, K. E., Hung, K.-L.-J., McCann, K. R., Ludka, J. T. & Holway, D. A. Floral visitation by the Argentine ant reduces pollinator visitation and seed set in the coast barrel cactus, Ferocactus viridescens. Oecologia 174, 163–171 (2014).
    ADS  PubMed  Article  PubMed Central  Google Scholar 

    23.
    Fuster, F., Kaiser-Bunbury, C. N. & Traveset, A. Pollination effectiveness of specialist and opportunistic nectar feeders influenced by invasive alien ants in the Seychelles. Am. J. Bot. 107, 957–969 (2020).
    PubMed  Article  PubMed Central  Google Scholar 

    24.
    Bissessur, P., Baider, C. & Florens, F. B. V. Infestation by pollination-disrupting alien ants varies temporally and spatially and is worsened by alien plant invasion. Biol. Invasions 22, 2573–2585 (2020).
    Article  Google Scholar 

    25.
    Del-Claro, K., Rodriguez-Morales, D., Calixto, E. S., Martins, A. S. & Torezan-Silingardi, H. M. Ant pollination of Paepalanthus lundii (Eriocaulaceae) in Brazilian savanna. Ann. Bot. 123, 1159–1165 (2019).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    26.
    Kuriakose, G., Sinu, P. A. & Shivanna, K. R. Ant pollination of Syzygium occidentale, an endemic tree species of tropical rain forests of the Western Ghats, India. Arthropod-Plant Interact. 12, 647–655 (2018).
    Article  Google Scholar 

    27.
    Galen, C. & Cuba, J. Down the tube: pollinators, predators, and the evolution of flower shape in the alpine skypilot Polemonium viscosum. Evolution 55, 1963–1971 (2001).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    28.
    Tsuji, K., Hasyim Harlion, A. & Nakamura, K. Asian weaver ants, Oecophylla smaragdina, and their repelling of pollinators. Ecol. Res. 19, 669–673 (2004).
    Article  Google Scholar 

    29.
    Ness, J. H. A mutualism’s indirect costs: the most aggressive plant bodyguards also deter pollinators. Oikos 113, 506–514 (2006).
    Article  Google Scholar 

    30.
    Lach, L. Argentine ants displace floral arthropods in a biodiversity hotspot. Divers. Distrib. 14, 281–290 (2008).
    Article  Google Scholar 

    31.
    Hansen, D. M. & Müller, C. B. Invasive ants disrupt gecko pollination and seed dispersal of the endangered plant Roussea simplex in Mauritius. Biotropica 41, 202–208 (2009).
    Article  Google Scholar 

    32.
    Lach, L. Interference and exploitation competition of three nectar-thieving invasive ant species. Insectes Soc. 52, 257–262 (2005).
    Article  Google Scholar 

    33.
    Blancafort, X. & Gómez, C. Consequences of the Argentine ant, Linepithema humile (Mayr), invasion on pollination of Euphorbia characias (L.) (Euphorbiaceae). Acta Oecol. 28, 49–55 (2005).
    ADS  Article  Google Scholar 

    34.
    Holway, D. A. Competitive mechanisms underlying the displacement of native ants by the invasive Argentine ant. Ecology 80, 238–251 (1999).
    Article  Google Scholar 

    35.
    Holway, D. A., Lach, L., Suarez, A. V., Tsutsui, N. D. & Case, T. J. The causes and consequences of ant invasions. Annu. Rev. Ecol. Syst. 33, 181–233 (2002).
    Article  Google Scholar 

    36.
    Silverman, J. & Buczkowski, G. 13 Behaviours mediating ant invasions. in Biological Invasions and Animal Behaviour 221 (2016).

    37.
    Sinu, P. A. et al. Effect of flower sex ratio on fruit set in pumpkin (Cucurbita maxima). Sci. Hortic. 246, 1005–1008 (2019).
    Article  Google Scholar 

    38.
    Sinu, P. A., Pooja, A. R. & Aneha, K. Overhead sprinkler irrigation affects pollinators and pollination in pumpkin (Cucurbita maxima). Sci. Hortic. 258, 108803 (2019).
    Article  Google Scholar 

    39.
    Bharti, H., Guénard, B., Bharti, M. & Economo, E. P. An updated checklist of the ants of India with their specific distributions in Indian states (Hymenoptera, Formicidae). ZooKeys 551, 1–83 (2016).
    Article  Google Scholar 

    40.
    R Core Team. R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, Vienna, 2018).
    Google Scholar 

    41.
    Anusree, T. et al. Flower sex expression in cucurbit crops of Kerala: implications for pollination and fruitset. Curr. Sci. 109, 2299–2302 (2015).
    Article  Google Scholar 

    42.
    das Vidal, M. G., de Jong, D., Wien, H. C. & Morse, R. A. Produção de néctar e pólen em abóbora (Cucurbita pepo L). Braz. J. Bot. 29, 267–273 (2006).
    Article  Google Scholar 

    43.
    Junker, R., Chung, A. Y. & Blüthgen, N. Interaction between flowers, ants and pollinators: additional evidence for floral repellence against ants. Ecol. Res. 22, 665–670 (2007).
    Article  Google Scholar 

    44.
    Ibarra-Isassi, J. & Sendoya, S. F. Ants as floral visitors of Blutaparon portulacoides (A. St-Hil.) Mears (Amaranthaceae): an ant pollination system in the Atlantic Rainforest. Arthropod-Plant Interact. 10, 221–227 (2016).
    Article  Google Scholar 

    45.
    Beattie, A. J., Turnbull, C., Knox, R. B. & Williams, E. G. Ant inhibition of pollen function: a possible reason why ant pollination is rare. Am. J. Bot. 71, 421–426 (1984).
    Article  Google Scholar 

    46.
    Hickman, J. C. Pollination by ants: a low-energy system. Science 184, 1290–1292 (1974).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    47.
    Galen, C. The effects of nectar thieving ants on seedset in floral scent morphs of Polemonium viscosum. Oikos 41, 245–249 (1983).
    Article  Google Scholar 

    48.
    Witte, V., Attygalle, A. B. & Meinwald, J. Complex chemical communication in the crazy ant Paratrechina longicornis Latreille (Hymenoptera: Formicidae). Chemoecology 17, 57–62 (2007).
    Article  Google Scholar 

    49.
    Wetterer, J. Worldwide spread of the ghost ant, Tapinoma melanocephalum (Hymenoptera: Formicidae). Myrmecol. News 12, 23–33 (2009).
    Google Scholar 

    50.
    Todd, B. D. et al. Habitat alteration increases invasive fire ant abundance to the detriment of amphibians and reptiles. Biol. Invasions 10, 539–546 (2008).
    Article  Google Scholar  More

  • in

    Mayetiola destructor (Diptera: Cecidmyiidae) host preference and survival on small grains with respect to leaf reflectance and phytohormone concentrations

    1.
    Wiseman, B. R. Plant-resistance to insects in integrated pest-management. Plant Dis. 78, 927–932. https://doi.org/10.1094/pd-78-0927 (1994).
    Article  Google Scholar 
    2.
    Painter, R. H. Insect resistance in crop plants. Soil Sci. 72 (1951).

    3.
    Orr, D. B. & Boethel, D. J. Influence of plant antibiosis through four trophic levels. Oecologia 70, 242–249. https://doi.org/10.1007/BF00379247 (1986).
    ADS  CAS  Article  PubMed  Google Scholar 

    4.
    Smith, C. M. & Clement, S. L. Molecular Bases of Plant Resistance to Arthropods. Annu. Rev. Entomol. 57, 309–328. https://doi.org/10.1146/annurev-ento-120710-100642 (2011).
    CAS  Article  PubMed  Google Scholar 

    5.
    Radcliffe, R. H. in Radcliffe’s IPM world textbook Vol. https://ipmworld.umn.edu/ratcliffe-hessian-fly (eds Radcliffe E.B. & Hutchison W.D.) (University of Minnesota, 1997).

    6.
    Kosma, D. K., Nemacheck, J. A., Jenks, M. A. & Williams, C. E. Changes in properties of wheat leaf cuticle during interactions with Hessian fly. Plant J 63, 31–43 (2010).
    CAS  PubMed  Google Scholar 

    7.
    Smiley, R. W., Gourlie, J. A., Whittaker, R. G., Easley, S. A. & Kidwell, K. K. Economic impact of Hessian fly (Diptera: Cecidomyiidae) on spring wheat in Oregon and additive yield losses with Fusarium crown rot and lesion nematode. J Econ Entomol 97, 397–408 (2004).
    Article  Google Scholar 

    8.
    Harris, M. O., Sandanayaka, M. & Griffin, A. Oviposition preferences of the Hessian fly and their consequences for the survival and reproductive potential of offspring. Ecol. Entomol. 26, 473–486. https://doi.org/10.1046/j.1365-2311.2001.00344.x (2001).
    Article  Google Scholar 

    9.
    Ganehiarachchi, G. A. S. M., Anderson, K. M., Harmon, J. & Harris, M. O. Why oviposit there? Fitness consequences of a gall midge choosing the plant’s youngest leaf. Environ Entomol 42, 123–130 (2013).
    CAS  Article  Google Scholar 

    10.
    Kanno, H. & Harris, M. O. Physical features of grass leaves influence the placement of eggs within the plant by the Hessian fly. Entomol. Exp. Appl. 96, 69–80. https://doi.org/10.1046/j.1570-7458.2000.00680.x (2000).
    Article  Google Scholar 

    11.
    Harris, M. O. & Rose, S. Chemical, color, and tactile cues influencing oviposition behavior of the Hessian fly (Diptera, Cecidomyiidae). Environ. Entomol. 19, 303–308. https://doi.org/10.1093/ee/19.2.303 (1990).
    Article  Google Scholar 

    12.
    Kanno, H. & Harris, M. O. Leaf physical and chemical features influence selection of plant genotypes by hessian fly. J. Chem. Ecol. 26, 2335–2354 (2000).
    CAS  Article  Google Scholar 

    13.
    Cervantes, D. E., Eigenbrode, S. D., Ding, H. J. & Bosque-Perez, N. A. Oviposition responses by Hessian fly, Mayetiola destructor, to wheats varying in surface waxes. J. Chem. Ecol. 28, 193–210 (2002).
    CAS  Article  Google Scholar 

    14.
    Morris, B. D., Foster, S. P. & Harris, M. O. Identification of 1-octacosanal and 6-methoxy-2-benzoxazolinone from wheat as ovipositional stimulants for Hessian fly, Mayetiola destructor. J. Chem. Ecol. 26, 859–873 (2000).
    CAS  Article  Google Scholar 

    15.
    Harris, M. O., Rose, S. & Malsch, P. The role of vision in the host plant-finding behavior of the Hessian fly. Physiol. Entomol. 18, 31–42. https://doi.org/10.1111/j.1365-3032.1993.tb00446.x (1993).
    Article  Google Scholar 

    16.
    Rohfritsch, O. A fungus associated gall midge, Lasioptera arundinis (Schiner), on Phragmites australis (Cav) Trin. Bull. Soc. Bot. France Lett. Bot. 139, 45–59. https://doi.org/10.1080/01811797.1992.10824942 (1992).
    Article  Google Scholar 

    17.
    Schmid, R. B., Knutson, A., Giles, K. L. & McCornack, B. P. Hessian fly (Diptera: Cecidomyiidae) biology and management in wheat. J. Integr. Pest Manag. 9, 12. https://doi.org/10.1093/jipm/pmy008 (2018).
    Article  Google Scholar 

    18.
    Gagné, R. J. & Hatchett, J. H. Instars of the Hessian Fly (Diptera: Cecidomyiidae). Ann. Entomol. Soc. Am. 82, 73–79. https://doi.org/10.1093/aesa/82.1.73 (1989).
    Article  Google Scholar 

    19.
    Lidell, M. C. & Schuster, M. F. Distribution of the Hessian fly and its control in Texas. Southwestern Entomologist 15, 133–145 (1990).
    Google Scholar 

    20.
    Morgan, G., Sansone, C. & Knutson, A. Hessian fly in Texas wheat. E-350 (Texas A&M, 2005).

    21.
    Flanders, K. L., Reisig, D. D., Buntin, G. D., Herbert, J. D. A. & Johnson, D. W. Biology and management of Hessian fly in the Southeast. ANR1069 (Alabama Cooperative Extension System, 2013).

    22.
    Wellso, S. G. Aestivation and Phenology of the Hessian Fly (Diptera: Cecidomyiidae) in Indiana. Environ. Entomol. 20, 795–801. https://doi.org/10.1093/ee/20.3.795 (1991).
    Article  Google Scholar 

    23.
    Boyd, M. L. & Bailey, W. C. Hessian fly management on wheat. G7180 (Missouri Extension, University of Missouri-Columbia, 2000).

    24.
    Ando, K. et al. Genome-wide associations for multiple pest resistances in a Northwestern United States elite spring wheat panel. PLoS One 13, e0191305/0191301-e0191305/0191325. https://doi.org/10.1371/journal.pone.0191305 (2018).

    25.
    Anderson, K. M. & Harris, M. O. Susceptibility of North Dakota Hessian Fly (Diptera: Cecidomyiidae) to 31 H Genes Mediating Wheat Resistance. J. Econ. Entomol. 112, 2398–2406 (2019).
    Article  Google Scholar 

    26.
    Sardesai, N., Nemacheck, J. A., Subramanyam, S. & Williams, C. E. Identification and mapping of H32, a new wheat gene conferring resistance to Hessian fly. Theor. Appl. Genet. 111, 1167–1173 (2005).
    CAS  Article  Google Scholar 

    27.
    Zhu, L., Liu, X. & Chen, M.-S. Differential accumulation of phytohormones in wheat seedlings attacked by avirulent and virulent Hessian fly (Diptera: Cecidomyiidae) larvae. J. Econ. Entomol. 103, 178–185 (2010).
    CAS  Article  Google Scholar 

    28.
    Mithöfer, A. & Boland, W. Recognition of Herbivory-Associated Molecular Patterns. Plant Physiol. 146, 825. https://doi.org/10.1104/pp.107.113118 (2008).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    29.
    Stuart, J. J., Chen, M.-S., Shukle, R. & Harris, M. O. Gall midges (Hessian flies) as plant pathogens. Annu. Rev. Phytopathol. 50, 339–357 (2012).
    CAS  Article  Google Scholar 

    30.
    Liu, X. et al. Gene expression of different wheat genotypes during attack by virulent and avirulent Hessian fly (Mayetiola destructor) larvae. J. Chem. Ecol. 33, 2171–2194 (2007).
    CAS  Article  Google Scholar 

    31.
    Subramanyam, S. et al. Expression of two wheat defense-response genes, Hfr-1 and Wci-1, under biotic and abiotic stresses. Plant Sci. 170, 90–103. https://doi.org/10.1016/j.plantsci.2005.08.006 (2006).
    CAS  Article  Google Scholar 

    32.
    Wu, J. et al. Differential responses of wheat inhibitor-like genes to Hessian fly, Mayetiola destructor, attacks during compatible and incompatible interactions. J. Chem. Ecol. 34, 1005–1012 (2008).
    CAS  Article  Google Scholar 

    33.
    Giovanini, M. P. et al. A novel wheat gene encoding a putative chitin-binding lectin is associated with resistance against Hessian fly. Mol. Plant Pathol. 8, 69–82 (2007).
    CAS  Article  Google Scholar 

    34.
    Liu, X. et al. Reactive oxygen species are involved in plant defense against a gall midge. Plant Physiol. 152, 985. https://doi.org/10.1104/pp.109.150656 (2010).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    35.
    Bari, R. & Jones, J. D. G. Role of plant hormones in plant defence responses. Plant Mol. Biol. 69, 473–488. https://doi.org/10.1007/s11103-008-9435-0 (2009).
    CAS  Article  PubMed  Google Scholar 

    36.
    Denancé, N., Sánchez-Vallet, A., Goffner, D. & Molina, A. Disease resistance or growth: the role of plant hormones in balancing immune responses and fitness costs. Frontiers in Plant Science 4. https://doi.org/10.3389/fpls.2013.00155 (2013)

    37.
    Dinh, S. T., Baldwin, I. T. & Galis, I. The HERBIVORE ELICITOR-REGULATED1 gene enhances abscisic acid levels and defenses against herbivores in Nicotiana attenuate plants. Plant Physiol. 162, 2106–2124 (2013).
    CAS  Article  Google Scholar 

    38.
    War, A. R., Paulraj, M. G., War, M. Y. & Ignacimuthu, S. Role of salicylic acid in induction of plant defense system in chickpea (Cicer arietinum L.). Plant signaling & behavior 6, 1787–1792. https://doi.org/10.4161/psb.6.11.17685 (2011).

    39.
    Nguyen, D., Rieu, I., Mariani, C. & van Dam, N. M. How plants handle multiple stresses: hormonal interactions underlying responses to abiotic stress and insect herbivory. Plant Mol. Biol. 91, 727–740. https://doi.org/10.1007/s11103-016-0481-8 (2016).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    40.
    Lee, A. et al. Inverse correlation between jasmonic acid and salicylic acid during early wound response in rice. Biochem. Biophys. Res. Commun. 318, 734–738. https://doi.org/10.1016/j.bbrc.2004.04.095 (2004).
    CAS  Article  PubMed  Google Scholar 

    41.
    Kunkel, B. N. & Brooks, D. M. Cross talk between signaling pathways in pathogen defense. Curr. Opin. Plant Biol. 5, 325–331. https://doi.org/10.1016/S1369-5266(02)00275-3 (2002).
    CAS  Article  PubMed  Google Scholar 

    42.
    Farmer, E. E., Alméras, E. & Krishnamurthy, V. Jasmonates and related oxylipins in plant responses to pathogenesis and herbivory. Curr. Opin. Plant Biol. 6, 372–378. https://doi.org/10.1016/S1369-5266(03)00045-1 (2003).
    CAS  Article  PubMed  Google Scholar 

    43.
    Loake, G. & Grant, M. Salicylic acid in plant defence—the players and protagonists. Curr. Opin. Plant Biol. 10, 466–472. https://doi.org/10.1016/j.pbi.2007.08.008 (2007).
    CAS  Article  PubMed  Google Scholar 

    44.
    Felton, G. W., Bi, J. L., Summers, C. B., Mueller, A. J. & Duffey, S. S. Potential role of lipoxygenases in defense against insect herbivory. J. Chem. Ecol. 20, 651–666. https://doi.org/10.1007/BF02059605 (1994).
    CAS  Article  PubMed  Google Scholar 

    45.
    Audenaert, K., De Meyer, G. B. & Höfte, M. M. Abscisic Acid Determines Basal Susceptibility of Tomato to Botrytis cinerea and Suppresses Salicylic Acid-Dependent Signaling Mechanisms. Plant Physiol. 128, 491. https://doi.org/10.1104/pp.010605 (2002).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    46.
    Mohr, P. G. & Cahill, D. M. Suppression by ABA of salicylic acid and lignin accumulation and the expression of multiple genes, in Arabidopsis infected with Pseudomonas syringae pv. tomato. Functional & Integrative Genomics 7, 181–191, https://doi.org/10.1007/s10142-006-0041-4 (2007).

    47.
    Harris, M. O., Dando, J. L., Griffin, W. & Madie, C. Susceptibility of cereal and non-cereal grasses to attack by Hessian fly (Mayetiola destructor (Say)). N. Zeal. J. Crop Hortic. Scie.ce 24, 229–238. https://doi.org/10.1080/01140671.1996.9513957 (1996).
    Article  Google Scholar 

    48.
    Gitelson, A. A. & Merzlyak, M. N. Signature analysis of leaf reflectance spectra: algorithm development for remote sensing of chlorophyll. J. Plant Physiol. 148, 494–500. https://doi.org/10.1016/S0176-1617(96)80284-7 (1996).
    CAS  Article  Google Scholar 

    49.
    Foster, S. P. & Harris, M. O. Foliar chemicals of wheat and related grasses influencing oviposition by Hessian fly, Mayetiola destructor (Say) (Diptera: Cecidomyiidae). J Chem Ecol 18, 1965–1980 (1992).
    CAS  Article  Google Scholar 

    50.
    Gagne, R. J., Hatchett, J. H., Lhaloui, S. & El Bouhssini, M. Hessian fly and barley stem gall midge, two different species of mayetiola (Diptera: Cecidomyiidae) in Morocco. Ann. Entomol. Soc. Am. 84, 436–443. https://doi.org/10.1093/aesa/84.4.436 (1991).
    Article  Google Scholar 

    51.
    Cherif, A., Kinoshita, N., Taylor, D. & Mediouni Ben Jemâa, J. Molecular characterization and phylogenetic comparisons of three Mayetiola species (Diptera: Cecidomyiidae) infesting cereals in Tunisia. Applied Entomology and Zoology 52, 543–551, https://doi.org/10.1007/s13355-017-0507-y (2017).

    52.
    Gould, F. Simulation models for predicting durability of insect-resistant germ plasm: hessian fly (Diptera: Cecidomyiidae)-resistant Winter Wheat. Environ. Entomol. 15, 11–23. https://doi.org/10.1093/ee/15.1.11 (1986).
    Article  Google Scholar 

    53.
    Chen, M.-S., Liu, X., Wang, H. & El-Bouhssini, M. Hessian fly (Diptera: Cecidomyiidae) interactions with barley, rice, and wheat seedlings. J Econ Entomol 102, 1663–1672 (2009).
    Article  Google Scholar 

    54.
    Ratcliffe, R. H., Safranski, G. G., Patterson, F. L., Ohm, H. W. & Taylor, P. L. Biotype status of Hessian fly (Diptera, Cecidomyiidae) populations from the eastern United-States and their response to 14 Hessian fly resistance genes. J. Econ. Entomol. 87, 1113–1121. https://doi.org/10.1093/jee/87.4.1113 (1994).
    Article  Google Scholar 

    55.
    Tooker, J. F. & Frank, S. D. Genotypically diverse cultivar mixtures for insect pest management and increased crop yields. J. Appl. Ecol. 49, 974–985. https://doi.org/10.1111/j.1365-2664.2012.02173.x (2012).
    Article  Google Scholar 

    56.
    Erb, M., Meldau, S. & Howe, G. A. Role of phytohormones in insect-specific plant reactions. Trends Plant Sci. 17, 1–20 (2012).
    Article  Google Scholar 

    57.
    Williams, C. E., Collier, C. C., Nemacheck, J. A., Liang, C. Z. & Cambron, S. E. A lectin-like wheat gene responds systemically to attempted feeding by avirulent first-instar Hessian fly larvae. J. Chem. Ecol. 28, 1411–1428. https://doi.org/10.1023/a:1016200619766 (2002).
    CAS  Article  PubMed  Google Scholar 

    58.
    Herrera-Vasquez, A., Salinas, P. & Holuigue, L. Salicylic acid and reactive oxygen species interplay in the transcriptional control of defense genes expression (vol 6, 171, 2015). Frontiers in Plant Science 8, https://doi.org/10.3389/fpls.2017.00964 (2017).

    59.
    Hatchett, J. H., Kreitner, G. L. & Elzinga, R. J. Larval Mouthparts and Feeding Mechanism of the Hessian Fly (Diptera: Cecidomyiidae). Ann. Entomol. Soc. Am. 83, 1137–1147. https://doi.org/10.1093/aesa/83.6.1137 (1990).
    Article  Google Scholar 

    60.
    Schotzko, D. J. & Bosque-Perez, N. A. Relationship between Hessian fly infestation density and early seedling growth of resistant and susceptible wheat. J. Agric. Urban Entomol. 19, 95–107 (2002).
    Google Scholar 

    61.
    Ratcliffe, R. H. et al. Biotype composition of Hessian fly (Diptera: Cecidomyiidae) populations from the southeastern, midwestern, and northwestern United States and virulence to resistance genes in wheat. J Econ. Entomol. 93, 1319–1328 (2000).
    CAS  Article  Google Scholar 

    62.
    Song, S., Gong, W., Zhu, B. & Huang, X. Wavelength selection and spectral discrimination for paddy rice, with laboratory measurements of hyperspectral leaf reflectance. ISPRS J. Photogram. Remote Sens. 66, 672–682. https://doi.org/10.1016/j.isprsjprs.2011.05.002 (2011).
    ADS  Article  Google Scholar 

    63.
    Dechant, B., Cuntz, M., Vohland, M., Schulz, E. & Doktor, D. Estimation of photosynthesis traits from leaf reflectance spectra: correlation to nitrogen content as the dominant mechanism. Remote Sens. Environ. 196, 279–292. https://doi.org/10.1016/j.rse.2017.05.019 (2017).
    ADS  Article  Google Scholar 

    64.
    Ollinger, S. V. Sources of variability in canopy reflectance and the convergent properties of plants. New Phytol. 189, 375–394. https://doi.org/10.1111/j.1469-8137.2010.03536.x (2011).
    CAS  Article  PubMed  Google Scholar 

    65.
    Almeida Trapp, M., De Souza, G. D., Rodrigues-Filho, E., Boland, W. & Mithöfer, A. Validated method for phytohormone quantification in plants. Frontiers in Plant Science 5, https://doi.org/10.3389/fpls.2014.00417 (2014).

    66.
    Davis, T. S., Bosque-Pérez, N. A., Popova, I. & Eigenbrode, S. D. Evidence for additive effects of virus infection and water availability on phytohormone induction in a staple crop. Frontiers in Ecology and Evolution 3, https://doi.org/10.3389/fevo.2015.00114 (2015). More

  • in

    The contribution of water radiolysis to marine sedimentary life

    Radiation experiments
    We experimentally quantified radiolytic hydrogen (H2) production in (i) pure water, (ii) seawater, and (iii) seawater-saturated sediment. We irradiated these materials with α- or γ-radiation for fixed time intervals and then determined the concentrations of H2 produced. Sediment samples were slurried with natural seawater to achieve a slurry porosity (φ) of ~0.83, which is the average porosity of abyssal clay in the South Pacific Gyre34. The seawater source is described below. To avoid microbiological uptake of radiolytic H2 during the course of the experiment, seawater and marine sediment slurries were pre-treated with HgCl2 (0.05% solution) or NaN3 (0.1% wt/vol). To ensure that addition of these chemicals did not impact radiolytic H2 yields, irradiation experiments with pure water plus HgCl2 or NaN3 were also conducted. HgCl2 or NaN3 addition had no statistically significant impact on H2 yields5,6,10.
    Experimental samples were irradiated in 250 mL borosilicate vials. A solid-angle 137Cs source (beam energy of 0.67 MeV) was used for the γ-irradiation experiments at the Rhode Island Nuclear Science Center (RINSC). The calculated dose rate for sediment slurries was 2.19E−02 Gy h−1 accounting for the (i) source activity, (ii) distance between the source and the samples, (iii) sample vial geometry, and (iv) attenuation coefficient of γ-radiation through air, borosilicate, and sediment slurry. 210Po (5.3 MeV decay−1)-plated silver strips with total activities of 250 μCi were used for the α-irradiation experiments. For α-irradiation of each sediment slurry, a 210Po-plated strip was placed inside the borosilicate vial and immersed in the slurry. Calculated total absorbed doses were 4 Gy and 3 kGy for γ-irradiation and α-irradiation experiments, respectively.
    The settling time of sediment grains in the slurries (1 week) was long compared to the time span of each experiment (tens of minutes to an hour for α-experiments, hours to days for γ-experiments). Therefore, we assumed that the suspension was homogenous during the course of each experiment.
    H2 concentrations were measured by quantitative headspace analysis via gas chromatography. For headspace analysis, 30 mL of N2 was first injected into the sample vial. To avoid over-pressurization of the sample during injection, an equivalent amount of water was allowed to escape through a separate needle. The vials were then vigorously shaken for 5 min to concentrate the H2 into the headspace. Finally, a 500-μL-headspace subsample was injected into a reduced gas analyzer (Peak Performer 1, PP1). The reduced gas analyzer was calibrated using a 1077 ppmv H2 primary standard (Scott-Marrin, Inc.). A gas mixer was used to dilute the H2 standard with N2 gas to obtain various H2 concentrations and produce a five-point linear calibration curve (0.7, 2, 5, 20, and 45 ppm). H2 concentrations of procedural blanks consisting of sample vials filled with non-irradiated deionized 18-MΩ water were also determined. The concentration detection limit obtained using this protocol was 0.8–1 nM H2. Relative error was less than 5%. Radiation experiments were performed at a minimum in triplicate.
    Sample selection and experimental radiolytic H2 yields, G(H2)
    Millipore Milli-Q system water was used for our pure-water experiments. For seawater experiments, we used bottom water collected in the Hudson Canyon (water depth, 2136 m) by RV Endeavor expedition EN534. Salinity of North Atlantic bottom water in the vicinity of the Hudson Canyon (34.96 g kg−1) is similar to that of mean open-ocean bottom water (34.70 g kg−1)44,45.
    The 20 sediment samples used for the experiments were collected by scientific coring expeditions in three ocean basins (expedition KN223 to the North Atlantic46, expedition KN195-3 to the Equatorial and North Pacific47, International Ocean Discovery Program (IODP) Expedition 329 to the South Pacific Gyre34, MONA expedition to the Guaymas Basin48, expedition EN32 to the Gulf of Mexico49, and expedition EN20 to the Venezuela Basin50). To capture the dominant sediment types present in the global ocean, we selected samples typical of five common sediment types [abyssal clay (11 samples), nannofossil-bearing clay or calcareous marl (2 samples), clay-bearing diatom ooze (3 samples), calcareous ooze (2 samples), and lithogenous sediment (2 samples)]. The locations, lithological descriptions, and mineral compositions of the samples are given in Supplementary Tables 1,  2,  3, and Supplementary Fig. 1. Additional chemical and physical descriptions of the sediment samples used in the radiation experiments can be found in the expedition reports for the expeditions on which the samples were collected34,46.
    Energy-normalized radiolytic H2 yields are commonly expressed as G(H2)-values (molecules H2 per 100 eV absorbed)1. As shown in Supplementary Fig. 2, for all irradiated samples (pure water, seawater, and marine sediment slurries), H2 production increased linearly with absorbed α- and γ-ray-dose. We calculated G(H2)-values for each sample and radiation type (α or γ) as the slope of the least-square regression line of radiolytic H2 concentration versus absorbed dose (Supplementary Fig. 2). The error on the yields is less than 10% for each sample. G(H2)-values for each sample and radiation type (α or γ) are reported in Supplementary Table 3.
    Although radiolytic OH• is known to react with dissolved organic matter51, total organic content does not appear to significantly impact radiolytic H2 production, since the most organic-rich sediment (e.g., Guaymas Basin and Gulf of Mexico sediment) did not yield particularly high H2 (Supplementary Table 3).
    Calculated radiolytic production rates of H2 and oxidants in the cored sediment of individual sites
    We calculated radiolytic H2 production rates (PH2, in molecules H2 cm−3 yr−1) for the cored sediment column at nine sites with oxic subseafloor sediment in the North Pacific, South Pacific, and North Atlantic; and seven sites with anoxic subseafloor sediment in the Bering Sea, South Pacific, Equatorial Pacific, and Peru Margin (see Supplementary Fig. 3 for site locations). For these calculations, we used the following equation from Blair et al.2:

    $$P_{{mathrm{H}}_2} = {sum} A _{{mathrm{m}},i}rho left( {1 – {{upvarphi}} } right)E_i{mathrm{G}}({mathrm{H}}_2)_i$$
    (1)

    where i is alpha, beta, or gamma radiation; Am is radioactivity per mass solid; φ is porosity; ρ is density solid; (E_i) is decay energy; and ({mathrm{G}}({mathrm{H}}_2)_i) is radiolytic yield.
    We calculated radiolytic oxidant production rates for these sediment columns from the H2 production rates. Because H2 production and oxidant production are stochiometrically balanced in water radiolysis [2H2O → H2 + H2O2], the calculated radiolytic H2 production rates (in electron equivalents) are equal to radiolytic oxidant production rates (in electron equivalents).
    The in situ γ- and α-radiation dosages in marine sediment are, respectively, 13 and 15 orders of magnitude lower than the dosage used in our experiments. Because the measured G(H2) for pure water in our γ-irradiation experiment (dose rate = 2.19E-02 Gy h−1) is statistically indistinguishable from previously published G(H2) values at much higher dose rates (ca. 1.00E+3 Gy h−1)5, we infer that the γ-irradiation G(H2) value is constant with dose rate over five orders of magnitude. Similarly, our experimental pure water H2 yields following α-particle irradiation from a 210Po-source (dose rate of 2.55E+03 Gy h−1) are indistinguishable from the yield obtained by Crumière et al.6 [G(H2) = 1.30 ± 0.13] for air-saturated deionized water exposed to a cyclotron-generated He2+ particle beam at higher dose rate (dose rate 1.62E+05 Gy h−1). The close similarity in H2 yields obtained in both experiments implies that (i) radiolytic H2 yield from α-particle irradiation is identical to that from cyclotron-generated He2+ particle irradiation, and (ii) this yield is constant over a two-orders-of-magnitude range dose rate. Therefore, we use our experimentally determined α- and γ-irradiation G(H2) values for the low radiation dose rate found in the subseafloor. Because the G(H2) of β irradiation has not been experimentally determined for water-saturated materials, we assume that the G(H2) of β-radiation matches the G(H2) of γ-radiation for the same sediment types. In pure water, their G(H2) values differ by only 17%1. Because β radiation, on average, contributes only 11% of the total radiolytic H2 production from the U, Th series and K decay in marine sediment, these estimates of total H2 production differ by only 2–5% relative to estimates where the G(H2) of β radiation is assumed equal to that for pure water or for α radiation of the same sediment types.
    To calculate H2 production rates for the entire sediment column at seven South Pacific sites and two North Atlantic sites, we measured downcore sediment profiles of U, Th, and K (i) 187 sediment samples from IODP Expedition 329 Sites U1365, U1366, U1367, U1368, U1369, U1370, and U137134,52, and (ii) 40 samples from KN223 expedition Sites 11 and 12 (ref. 46). Total U and Th (ppm) and K2O (wt%) for these sites are reported in the EarthChem SedDB data repository. We measured U, Th, and K abundances using standard atomic emission and mass spectrometry techniques (i.e. ICP-ES and ICP-MS) in the Analytical Geochemistry Facilities at Boston University. Sample preparation, analytical protocol, and data are reported in Dunlea et al.52. The precision for each element is ~2% of the measured value, based on three separate digestions of a homogenized in-house standard of deep-sea sediment.
    To calculate H2 production rates for the sediment columns at North Pacific coring Sites EQP10 and EQP11 (ref. 47), we used radioactive element content data from Kyte et al.53, who measured chemical concentrations at high resolution in bulk sediment in core LL44-GPC3. Because Site EQP11 was cored at the same location as LL44-GPC3 (ref. 53) and the sediment retrieved at all three sites is homogeneous abyssal clay, we assume the radioactive element abundances measured in core LL44-GPC3 to be representative of Sites EQP10 and EQP11 (ref. 47). Calculated radiolytic H2 production rates for South Pacific sites are listed in Supplementary Table 4 and for North Atlantic and North Pacific sites in Supplementary Table 5.
    For Bering Sea Sites U1343 and U1345 (ref. 54), sedimentary U, Th, and K content measurements are unavailable. Since sediment recovered at these two sites is primarily siliciclastic with a varying amount of diatom-rich clay, we use U, Th, and K concentration values reported for upper continental crust by Li and Schoonmaker for these Bering Sea sites55. Finally, we calculate downhole radiolytic H2 production rates for ODP Leg 201 Sites 1225, 1226, 1227, and 1230 (ref. 35). Sediment compositions for these sites include nannofossil-rich calcareous ooze (Site 1225), alternation of nannofossil (calcareous) ooze and diatom ooze (Site 1226), and siliciclastic with diatom-rich clay intervals (Sites 1227 and 1230). Because sedimentary U, Th, and K measurements are not available for Leg 201 sites, we used average U, Th, and K concentration values measured in North Atlantic46 and South Pacific Sites34,52 with corresponding lithologies.
    We use isotopic abundance values reported in Erlank et al.56 to calculate the abundance of 238U, 235U, 232Th, and 40K from the measured ICP-MS values of total U, Th, and K concentration. We then converted radionuclide concentrations to activities using Avogadro’s number and each isotope’s decay constant2. We refer to Blair et al. for a detailed explanation of activities and radiolytic yield calculations2.
    Calculation of global radiolytic H2 and oxidant production rates in marine sediment
    We calculated global radiolytic H2 production in ocean sediment by applying Eq. (1) (ref. 2) globally. As with the rates at individual sites, we calculated global radiolytic oxidant production (in electron equivalents) from global H2 production and the stochiometry of water radiolysis [2H2O → H2 + H2O2].
    Our global radiolytic H2 production calculation spatially integrates calculations of sedimentary porewater radiolysis rates that are based on (i) our experimentally constrained radiolytic H2 yields for the principal marine sediment types, (ii) measured radioactive element content of sediment cores in three ocean basins (North Atlantic46, North Pacific53, and South Pacific34,52), and (iii) global distributions of sediment lithology57, sediment porosity58, and sediment column length59,60.
    To generate the global map of radiolytic H2 production, we created global maps of seafloor U, Th, and K concentrations, density, G(H2)-α values, and G(H2)-γ-and-β by assigning each grid cell in our compiled seafloor lithology map (Supplementary Fig. 4) its lithology-specific set of input variables (Supplementary Table 6). Because our model assumes that lithology is constant with depth, U, Th, and K content, grain density, and G(H2)-values are constant with depth.
    The G(H2)-values (α, β, and γ radiation), radioactive element content (sedimentary U, Th, and K concentration), density, porosity, and sediment thickness are determined as follows.
    Radiolytic yield [G(H2)] for α,β-&-γ radiation
    Radiolytic yields for the main seafloor lithologies are obtained by averaging experimentally derived yields for the respective lithologies (Supplementary Table 6). We assume that G(H2)-β values equal G(H2)-γ values.
    Sediment lithology
    For these calculations of radiolytic chemical production, we generally used seafloor lithologies and assumed that sediment type is constant with sediment depth. For seafloor lithology, the geographic database of global bottom sediment types57 was compiled into five lithologic categories: abyssal clay, calcareous ooze, siliceous ooze, calcareous marl, and lithogenous (Supplementary Fig. 4). Some areas of the seafloor are not described in the database57. These include (i) high-latitude regions (as the seafloor lithology database extends from 70°N to 50°S)57 and (ii) some discrete areas located along continental margins (e.g., Mediterranean Sea, Timor Sea, South China Sea, Supplementary Fig. 4). We used other data sources to identify seafloor lithologies for these regions. We added an opal belt (siliceous ooze) in the Southern Ocean between 57°S and 66°S61,62. The geographic extent of this opal belt was based on DeMaster62 and Dutkiewicz et al.61. We defined the areas of the seafloor from 50°S to 57°S, from 66°S to 90°S, and in the Arctic Ocean as mostly composed of lithogenous material, based on (i) drillsite lithologies in the Southern Ocean [ODP: Site 695 (ref. 63), Site 694 (ref. 63), Site 1165 (ref. 64), Site 739 (ref. 65)], the Bering Sea and Arctic Ocean [International Ocean Discovery Program (IODP): Sites U1343 and U1345 (ref. 54), Site M0002 (ref. 66), ODP: Site 910 (ref. 67), Site 645 (ref. 68)] and between 50°S and 57°S [Deep Sea Drilling Project (DSDP): Site 326 (ref. 69), Ocean Drilling Program (ODP): Site 1138 (ref. 70), Site 1121 (ref. 71)], and Dutkiewicz et al.61.
    In the North and South Atlantic, sediment type can be very different at depth than at the seafloor. For these regions, we departed from our assumption that sediment lithology is the same at depth as at the seafloor. Subseafloor lithologies at ODP Sites [1063 (ref. 72), 951 (ref. 73), 925 (ref. 74), and 662 (ref. 75)] and IODP Sites [U1403 (ref. 76) and U1312 (ref. 77)] indicate that sediment in the Atlantic Ocean basin is generally 30–90% biogenic carbonate content and detrital clay78, even where the seafloor lithology is abyssal clay57. Therefore, regions in the Atlantic Ocean described as abyssal clay in the seafloor lithology database57 were characterized as calcareous marl for our calculations (Supplementary Fig. 4). Because abyssal clay catalyzes radiolytic H2 production at a higher rate than calcareous marl, this characterization may underestimate production of radiolytic H2 and radiolytic oxidants in these Atlantic regions.
    Radioactive element content
    For four of the five lithologic types in our global maps (abyssal clay, siliceous ooze, calcareous ooze, and calcareous marl), we average U, Th, and K concentrations from sites in the North Atlantic46, North Pacific53, and South Pacific34,52. The average U, Th, and K concentration values are consistent with data reported in Li and Schoonmaker55 for the characteristic U, Th, and K content found in abyssal clay and calcareous ooze. For lithogenous sediment, we use U, Th, and K concentration values reported for upper continental crust by Li and Schoonmaker55. Lithology-specific radioactive element values are given in Supplementary Table 6 and used to calculate Am,i in Eq. (1).
    Density
    Characteristic density values for calcite, quartz, terrigenous clay, and opal-rich sediment were extracted from the Proceedings of the Integrated Ocean Drilling Program Volume 320/321 and are assigned to calcareous ooze, lithogenous sediment, abyssal clay, and siliceous ooze, respectively79.
    Global porosity
    For global porosity, we use a seafloor porosity data set by Martin et al.58 and accounted for sediment compaction with depth by using separate sediment compaction length scales for continental-shelf (0–200 m water depth; c0 = 0.5 × 10−3), continental-margin (200–2500 m; c0 = 1.7 × 10−3), and abyssal sediment ( >3500 m; c0 = 0.85 × 10−3)80,81. Once the porosity was 0.1%, the depth integration was halted.
    Global sediment thickness
    We calculated global depth-integrated radiolytic H2 production by summing the seafloor production rates over sediment depth in one-meter intervals (Fig. 3 in main text). Sediment thickness is from Whittaker et al., supplemented with Laske and Masters where needed82,83.
    Ocean depth
    For porosity calculations, water depths were determined using the General Bathymetric Chart of the Oceans84, resampled to a 5-arc minute grid, i.e. the resolution of the Naval Oceanographic Office’s Bottom Sediment Type (BTS) database “Enhanced dataset”57.
    Dissolved H2 concentration profiles
    H2 concentrations from South Pacific Sites U1365, U1369, U1370, and U1371, and the measurement protocol, are described in ref. 1. H2 concentrations from North Atlantic KN223 Sites 11, 12, and 15, and North Pacific Site EQP11 were determined using the same protocol and are posted on SedDB (see “Data availability”). The detection limit for H2 ranged between 1 and 5 nM H2, depending on site, and is displayed as gray vertical lines in Fig. 2 of the main text. H2 concentrations for Equatorial Pacific Site 1225 and Peru Trench Site 1230 were measured by the “headspace equilibration technique”, which measures steady-state H2 levels reached following laboratory incubation of the sediment samples85,86.
    For comparison to these measured H2 concentrations, we use diffusion-reaction calculations to quantify what in situ H2 concentrations would be in the absence of H2-consuming reactions. The results of these calculations are represented as solid circles (•) in Fig. 2 of the main text. Temporal changes in H2 concentration due to diffusive processes and radiolytic H2 production in situ are expressed by Eq. (2):

    $$frac{{partial {mathrm{H}}_2(x,t)}}{{partial t}} = frac{D}{{varphi F}}frac{{partial ^2{mathrm{H}}_2(x,t)}}{{partial x^2}} + P(x)$$
    (2)

    with
    D: the diffusion coefficient of H2(aq) at in situ temperature
    (varphi): porosity
    F: formation factor
    x: depth
    Z: sediment column thickness
    ({mathrm{H}}_2): hydrogen concentration
    P: radiolytic H2 production rate
    t : time.
    With constant radiolytic H2 production, P(x) = P with depth,
    and at steady-state,

    $$frac{{partial ^2{mathrm{H}}_2(x)}}{{partial x^2}} = – frac{{Pvarphi F}}{D}.$$
    (3)

    We integrate Eq. (3) over the length x twice,

    $${mathrm{H}}_2(x) = – frac{1}{2}frac{{Pvarphi F}}{D}x^2 + Ax + B$$
    (4)

    where A and B in Eq. (4) are constants of integration. We use two boundary conditions to derive the value of these constants.
    Boundary condition 1: concentration of H2 at the sediment-water interface, x = 0, is zero due to diffusive loss to the overlying water column.
    Boundary condition 2: concentration of H2 at the basement–sediment-water interface, x = Z, is zero due to diffusive loss to the underlying basement.
    With these boundary conditions, (A = frac{1}{2}frac{{Pvarphi F}}{D}Z) and B = 0
    and

    $${mathrm{H}}_2(x) = frac{1}{2}frac{{Pvarphi F}}{D}(xZ – x^2).$$
    (5)

    In cases where we expect radiolytic H2 production rates to significantly vary with depth due to changes in lithology, we adapted the boundary conditions and applied a two-layer diffusion model to account for this variation.
    Calculation of Gibbs Energies for the Knallgas reaction
    For H2 concentrations above the detection limits at South Pacific IODP Expedition 329 sites (Supplementary Fig. 5)34, we quantified in situ Gibbs energies (ΔGr) of the Knallgas reaction (H2 + ½O2 → H2O). In situ ΔGr values depend on pressure (P), temperature (T), ionic strength, and chemical concentrations, all of which are explicitly accounted for in our calculations:

    $$Delta G_{mathrm{r}} = Delta G^circ _{mathrm{r}}left( {T,P} right) + 2.3,RT,{mathrm{log}}_{10}Q$$
    (6)

    where:
    ΔGr: in situ Gibbs energy of reaction (kJ mol H2−1)
    ΔG°r(T,P): Gibbs energy of reaction under in situ T and P conditions (kJ mol H2−1)
    R: gas constant (8.314 kJ−1 mol K−1)
    Q: activity quotient of compounds involved in the reaction.
    We use the measured composition of the sedimentary pore fluid to determine values of Q.
    For a more complete overview of in situ Gibbs energy-of-reaction calculations in subseafloor sediment, see Wang et al.87.
    Calculation of organic oxidation rates (net rates of O2 reduction and DIC production)
    We calculated the vertical distribution of net O2 reduction rates at nine sites where the sediment is oxic from seafloor to basement and the vertical distribution of DIC production rates at seven sites where the subseafloor sediment is anoxic (see Supplementary Fig. 3 for site locations). Dissolved O2 concentrations are from Røy et al.47 and D’Hondt et al.88. DIC concentrations are from ODP Leg 201 (ref. 35), and the Proceedings of the IODP Expedition 323 (Sites U12343, U1345)54 and IODP Expedition 329 (Site U1371 (ref. 34)).
    The net rates are calculated using the MatLab program and numerical procedures of Wang et al.89, modified by using an Akima spline, rather than a 5-point running mean, to generate a best-fit line to the chemical concentration data. Details of the calculation protocol for O2 production rates and DIC production rates are respectively described in the supplementary information of D’Hondt et al.88 and in Walsh et al.90. The DIC reaction rates and their first standard deviations calculated for the seven sites are given in Supplementary Table 7.
    To facilitate comparisons of radiolytic chemical rates to net DIC production rates, rates are converted to electron equivalents (2 electrons per H2, 4 electrons per O2, 4 electrons per organic C oxidized).
    Estimation of sediment ages
    We estimated sediment ages for Sites U1343 and U1343 using the sediment-age model of Takahashi et al.54, which is based on biostratigraphic and magnetostratigraphic data. Because detailed chronostratigraphic data are not available for the remaining sites (Equatorial Pacific sites (1225 and 1226), Peru Trench Site 1230 and Peru Basin Site 1231, South Pacific sites U1365, U1366, U1367, U1369, U1370, and U1371, North Pacific sites EQP9 and EQP10, and North Atlantic sites KN223-11 and KN223-12), we used the mean sediment accumulation rate for each of these sites (Supplementary Fig. 3) to convert its sediment depth (in meters below seafloor) to sediment age (in millions of years, Ma). Mean sediment accumulation rate was calculated by dividing sediment thickness by basement age91 (Supplementary Table 8). For Sites 1225, 1226, 1230, 1231, U1365, U1366, U1367, U1369, U1370, and U1371, sediment thickness was determined by drilling to basement34,35. For Sites EQP9, EQP10, KN223-11, and KN223-12, sediment thicknesses were determined from acoustic basement reflection data. More