More stories

  • in

    Declining severe fire activity on managed lands in Equatorial Asia

    Andela, N. et al. A human-driven decline in global burned area. Science 356, 1356 (2017).CAS 
    Article 

    Google Scholar 
    Sloan, S., Locatelli, B., Wooster, M. J. & Gaveau, D. L. A. Fire activity in Borneo driven by industrial land conversion and drought during El Niño periods, 1982–2010. Glob. Environ. Change 47, 95–109 (2017).Article 

    Google Scholar 
    Kelley, D. I. et al. How contemporary bioclimatic and human controls change global fire regimes. Nat. Clim. Change 9, 690–96 (2019).Article 

    Google Scholar 
    Jolly, W. M. et al. Climate-induced variations in global wildfire danger from 1979 to 2013. Nat. Commun. 6, 7537 (2015).Ward, D. S., Shevliakova, E., Malyshev, S. & Rabin, S. Trends and variability of global fire emissions due to historical anthropogenic activities. Glob. Biogeochem. Cycles 32, 122–42 (2018).CAS 
    Article 

    Google Scholar 
    Earl, N. & Simmonds, I. Spatial and temporal variability and trends in 2001–2016 global fire activity. J. Geophys. Res. Atmos. 123, 2524–36 (2018).Article 

    Google Scholar 
    Giglio, L., Randerson, J. T. & van der Werf, G. R. Analysis of daily, monthly, and annual burned area using the fourth-generation Global Fire Emissions Database (GFED4). J. Geophys. Res. Biogeosci. 118, 317–28 (2013).Article 

    Google Scholar 
    Doerr, S. H. & Santín, C. Global trends in wildfire and its impacts: perceptions versus realities in a changing world. Philos. Trans. R. Soc. B Biol. Sci. 371, 20150345 (2016).Article 

    Google Scholar 
    van Lierop, P., Lindquist, E., Sathyapala, S. & Franceschini, G. Global forest area disturbance from fire, insect pests, diseases and severe weather events. Forest Ecol. Manag. 352, 78–88 (2015).Article 

    Google Scholar 
    Zheng, B. et al. Increasing forest fire emissions despite the decline in global burned area. Sci. Adv. 7, eabh2646 (2021).Article 

    Google Scholar 
    Andela, N. & van der Werf, G. R. Recent trends in African fires driven by cropland expansion and El Niño to La Niña transition. Nat. Clim. Change 4, 791–95 (2014).Article 

    Google Scholar 
    Van der Werf, G. R. et al. Global fire emissions and the contribution of deforestation, savanna, forest, agricultural, and peat fires (1997–2009). Atmos. Chem. Phys. 10, 11707–35 (2010).Article 
    CAS 

    Google Scholar 
    Balch, J. K. et al. Negative fire feedback in a transitional forest of southeastern Amazonia. Glob. Change Biol. 14, 2276–87 (2008).Article 

    Google Scholar 
    Cochrane, M. A. & Laurance, W. F. Synergisms among fire, land use, and climate change in the Amazon. Ambio 37, 522–27 (2008).Article 

    Google Scholar 
    Gaveau, D. L. A. et al. Major atmospheric emissions from peat fires in Southeast Asia during non-drought years: evidence from the 2013 Sumatran fires. Sci. Rep. 4, 6112 (2014).Vadrevu, K. P. et al. Trends in vegetation fires in South and Southeast Asian countries. Sci. Rep. 9, 7422 (2019).Article 
    CAS 

    Google Scholar 
    Sloan, S., Tacconi, L. & Cattau, M. E. Fire prevention in managed landscapes: recent successes and challenges in Indonesia. Mitig. Adapt. Strateg. Glob. Change 26, Article 32 (2021).Article 

    Google Scholar 
    Gaveau, D. L. A., Descales, A., Salim, M. A., Shields, D. & Sloan, S. Refined burned-area mapping protocol using Sentinel-2 data increases estimate of 2019 Indonesian burning. Earth Syst. Sci. Data, https://doi.org/10.5194/essd-2021-113, (2021).Randerson, J. T., Chen, Y., van der Werf, G. R., Rogers, B. M. & Morton, D. C. Global burned area and biomass burning emissions from small fires. J. Geophys. Res. Biogeosci. 117, G04012 (2012).Article 
    CAS 

    Google Scholar 
    Field, R. D., van der Werf, G. R. & Shen, S. S. P. Human amplification of drought-induced biomass burning in Indonesia since 1960. Nat. Geosci. 2, 185–88 (2009).CAS 
    Article 

    Google Scholar 
    Huijnen, V. et al. Fire carbon emissions over maritime Southeast Asia in 2015 largest since 1997. Sci. Rep. 6, 26886 (2016).CAS 
    Article 

    Google Scholar 
    Tacconi, L. Preventing fires and haze in Southeast Asia. Nat. Clim. Change 6, 640–43 (2016).Article 

    Google Scholar 
    Koplitz, S. N. et al. Public health impacts of the severe haze in Equatorial Asia in September–October 2015: demonstration of a new framework for informing fire management strategies to reduce downwind smoke exposure. Environ. Res. Lett. 11, 094023 (2016).Article 

    Google Scholar 
    Kiely, L. et al. Air quality and health impacts of vegetation and peat fires in Equatorial Asia during 2004–2015. Environ. Res. Lett.15, 094054 (2020).Article 

    Google Scholar 
    Crippa, P. et al. Population exposure to hazardous air quality due to the 2015 fires in Equatorial Asia. Sci. Rep. 6, 37074 (2016).CAS 
    Article 

    Google Scholar 
    Glauber, A. J. & Gunawan, I. The Cost of Fire: An Economic Analysis of Indonesia’s 2015 Fire Crisis. (The World Bank, Washington, D.C., (2016).Tan, Z. D., Carrasco, L. R. & Taylor, D. Spatial correlates of forest and land fires in Indonesia. Int. J. Wildland Fire 29, 1088–99 (2020).Article 

    Google Scholar 
    Marlier, M. E. et al. Fire emissions and regional air quality impacts from fires in oil palm, timber, and logging concessions in Indonesia. Environ. Res. Lett. 10, 085005 (2015).Article 
    CAS 

    Google Scholar 
    Vetrita, Y. & Cochrane, M. A. Fire frequency and related land-use and land-cover changes in Indonesia’s peatlands. Remote Sens. 12, 5 (2020).Nikonovas, T., Spessa, A., Doerr, S. H., Clay, G. D. & Mezbahuddin, S. Near-complete loss of fire-resistant primary tropical forest cover in Sumatra and Kalimantan. Commun. Earth Environ. 1, 65 (2020).Article 

    Google Scholar 
    Field, R. Biomass burning in Indonesia: Signs of Progress in 2019?, http://www.columbia.edu/~rf2426/index_files/20200128.Field.GSFC.NoOz.pdf, January, NASA Goddard Space Flight Center, (2019).Watts, J. et al. Incentivising compliance: evaluating the effectiveness of targeted village incentives for reducing forest and peat fires. Forest Policy Econ. 108, 101956 (2019).Wijedasa, L. et al. Carbon emissions from peat forests will continue to increase despite emission-reduction schemes. Glob. Change Biol. 24, 4598–613 (2018).Article 

    Google Scholar 
    Sloan, S., Meyfroidt, P., Rudel, T. K. & Bongers, F. & Chazdon Robin, L. The forest transformation: Planted tree cover and regional dynamics of tree gains and losses. Glob. Environ. Change 59, 101988 (2019).Article 

    Google Scholar 
    Albar, I., Jaya, I. N. S., Saharjo, B. H., Kuncahyo, B. & Vadrevu, K. P. Spatio-temporal analysis of land and forest fires in Indonesia using MODIS active fire dataset, in Land-Atmospheric Research Applications in South and Southeast Asia (eds K P Vadrevu et al.), p. 105-27 (Springer International Publishing, 2018).Miettinen, J., Shi, C. & Liew, S. C. Fire distribution in Peninsular Malaysia, Sumatra and Borneo in 2015 with special emphasis on peatland fires. Environ. Manage. 60, 747–57 (2017).Article 

    Google Scholar 
    Fanin, T. & van der Werf, G. R. Precipitation–fire linkages in Indonesia (1997–2015). Biogeosciences 14, 3995–4008 (2017).Article 

    Google Scholar 
    Wiggins, E. B. et al. Smoke radiocarbon measurements from Indonesian fires provide evidence for burning of millennia-aged peat. Proc. Natl. Acad. Sci. USA 115, 12419 (2018).CAS 
    Article 

    Google Scholar 
    Page, S. E. et al. The amount of carbon released from peat and forest fires in Indonesia during 1997. Nature 420, 61–65 (2002).CAS 
    Article 

    Google Scholar 
    Lohberger, S., Stängel, M., Atwood, E. C. & Siegert, F. Spatial evaluation of Indonesia’s 2015 fire-affected area and estimated carbon emissions using Sentinel-1. Glob. Change Biol. 24, 644–54 (2018).Article 

    Google Scholar 
    van der Werf, G. R. et al. Global fire emissions estimates during 1997–2016. Earth Syst. Sci. Data 9, 697–720 (2017).Article 

    Google Scholar 
    Field, R. D. et al. Indonesian fire activity and smoke pollution in 2015 show persistent nonlinear sensitivity to El Niño-induced drought. Proc. Natl Acad. Sci. USA 113, 9204–09 (2016).CAS 
    Article 

    Google Scholar 
    Austin, K. G. et al. Shifting patterns of oil palm driven deforestation in Indonesia and implications for zero-deforestation commitments. Land Use Policy 69, 41–48 (2017).Article 

    Google Scholar 
    Pan, X., Chin, M., Ichoku, C. & Field, R. Connecting Indonesian fires and drought with the type of El Niño and phase of the Indian Ocean Dipole during 1979–2016. J. Geophys. Res. Atmos. 123, (2018).van der Werf, G. R. et al. Climate regulation of fire emissions and deforestation in Equatorial Asia. Proc. Natl Acad. Sci. USA 105, 20350–55 (2008).Article 

    Google Scholar 
    Wooster, M. J., Roberts, G., Perry, G. L. W. & Kaufman, Y. J. Retrieval of biomass combustion rates and totals from fire radiative power observations: FRP derivation and calibration relationships between biomass consumption and fire radiative energy release. J. Geophys. Rese. Atmos. 110, (2005).Spessa, A. et al. Seasonal forecasting of fires over Kalimantan, Indonesia. Nat. Hazards Earth Syst. Sci. 15, 429–42 (2015).Article 

    Google Scholar 
    Siegert, F., Ruecker, G., Hinrichs, A. & Hoffmann, A. A. Increased damage from fires in logged forests during droughts caused by El Niño. Nature 414, 437–40 (2001).CAS 
    Article 

    Google Scholar 
    Fernandes, K. et al. Heightened fire probability in Indonesia in non-drought conditions: the effect of increasing temperatures. Environ. Res. Lett. 12, 054002 (2017).Article 

    Google Scholar 
    Herawati, H. & Santoso, H. Tropical forest susceptibility to and risk of fire under changing climate: a review of fire nature, policy and institutions in Indonesia. Forest Policy Econ. 13, 227–33 (2011).Article 

    Google Scholar 
    Nepstad, D. et al. Slowing Amazon deforestation through public policy and interventions in beef and soy supply chains. Science 344, 1118–23 (2014).CAS 
    Article 

    Google Scholar 
    Dennis, R. A Review of Fire Projects In Indonesia, 1982-1998. (CIFOR, Bogor, Indonesia, 1999).de Groot, W. J., Field, R. D., Brady, M. A., Roswintiarti, O. & Mohamad, M. Development of the Indonesian and Malaysian fire danger rating systems. Mitig. Adapt. Strateg. Glob. Change 12, 165 (2006).Article 

    Google Scholar 
    Clough, Y. et al. Land-use choices follow profitability at the expense of ecological functions in Indonesian smallholder landscapes. Nat. Commun. 7, 13137 (2016).CAS 
    Article 

    Google Scholar 
    Bissonnette, J.-F. & De Koninck, R. The return of the plantation? Historical and contemporary trends in the relation between plantations and smallholdings in Southeast Asia. J. Peasant Stud. 44, 918–38 (2017).Article 

    Google Scholar 
    Gaveau, D. L. A. et al. Slowing deforestation in Indonesia follows declining oil palm expansion and lower oil prices. PLOS ONE 17, e0266178 (2022).Svatoňová, T., Herák, D. & Kabutey, A. Financial profitability and sensitivity analysis of palm oil plantation in Indonesia. Acta Univ. Agric. Silvic. Mendelianae Brunensis 63, 1365–73 (2015).Article 

    Google Scholar 
    Gaveau, D. L. A. et al. Rapid conversions and avoided deforestation: examining four decades of industrial plantation expansion in Borneo. Scientific Reports 6, (2016).Simamora, A. P. Govt says no to converting peatland into plantations, The Jakarta Post. August (2010).Satriastanti, F. E. Jokowi bans new oil palm and mining concessions, Mongabay.com April (2016).Sloan, S., Edwards, D. P. & Laurance, W. F. Does Indonesia’s REDD+ moratorium on new concessions spare imminently-threatened forests? Conserv. Lett. 5, 222–31 (2012).Article 

    Google Scholar 
    Busch, J. et al. Reductions in emissions from deforestation from Indonesia’s moratorium on new oil palm, timber, and logging concessions. Proc. Natl Acad Sci USA 112, 1328–33 (2015).CAS 
    Article 

    Google Scholar 
    Forsyth, T. Public concerns about transboundary haze: a comparison of Indonesia, Singapore, and Malaysia. Glob. Environ. Change 25, 76–86 (2014).Article 

    Google Scholar 
    Carbon Conservation. Fire Free Village Program – Review 2017. (Carbon Conservation, Singapore, (2017).Gaveau, D. L. A. et al. Overlapping land claims limit the use of satellites to monitor no-deforestation committments and no-burning compliance. Conserv. Lett. 10, 257–64 (2017).Article 

    Google Scholar 
    EarthData. MODIS Collection 6 Active-Fire Detections standard scientific data (MCD14ML), NASA EarthData, https://earthdata.nasa.gov/firms (2019).Giglio, L., Schroeder, W. & Justice, C. O. The Collection 6 MODIS active fire detection algorithm and fire products. Remote Sens. Environ. 178, 31–41 (2016).Article 

    Google Scholar 
    Sloan, S., Cattau, M.E. Discrete Fire Events, their Severity, and their Ignitions, as Derived from MODIS MCD 14ML Active-Fire Detection Data for Indonesia, 2002-2019. Sean Sloan and Megan E. Cattau, Datadryad.org. (2022).Cattau, M. E. et al. Sources of anthropogenic fire ignitions on the peat-swamp landscape in Kalimantan, Indonesia. Glob. Environ. Change 39, 205–19 (2016).Article 

    Google Scholar 
    Wooster, M. J., Perry, G. L. W. & Zoumas, A. Fire, drought and El Niño relationships on Borneo during the pre-MODIS era (1980–2000). Biogeosciences 9, 317–40 (2012).Article 

    Google Scholar 
    Tansey, K., Beston, J., Hoscilo, A., Page, S. E. & Paredes Hernández, C. U. Relationship between MODIS fire hot spot count and burned area in a degraded tropical peat swamp forest in Central Kalimantan, Indonesia. J. Geophys. Res. 113, (2008).Oom, D., Silva, P. C., Bistinas, I. & Pereira, J. M. C. Highlighting biome-specific sensitivity of fire size distributions to time-gap parameter using a new algorithm for fire event individuation. Remote Sens. 8, 663 (2016).Schroeder, W. et al. Validation of GOES and MODIS active fire detection products using ASTER and ETM plus data. Remote Sens. Environ. 112, 2711–26 (2008).Article 

    Google Scholar 
    Hantson, S., Padilla, M., Corti, D. & Chuvieco, E. Strengths and weaknesses of MODIS hotspots to characterize global fire occurrence. Remote Sens. Environ. 131, 152–59 (2013).Article 

    Google Scholar 
    Tanpipat, V., Honda, K. & Nuchaiya, P. MODIS hotspot validation over Thailand. Remote Sens. 1, 1043–54 (2009).Article 

    Google Scholar 
    Liew, S. C., Shen, C., Low, J., Lim, A. & Kwoh, L. K. The 24th Asian Conference on Remote Sensing and 2003 International Symposium on Remote Sensing (ACRS2003). p. 671-73 (Asian Association on Remote Sensing), November 3–7.Fornacca, D., Ren, G. & Xiao, W. Performance of three MODIS fire products (MCD45A1, MCD64A1, MCD14ML), and ESA Fire_CCI in a mountainous area of northwest Yunnan, China, characterized by frequent small fires. Remote Sens. 9, 1131 (2017).Article 

    Google Scholar 
    Schroeder, W., Oliva, P., Giglio, L. & Csiszar, I. A. The New VIIRS 375m active fire detection data product: algorithm description and initial assessment. Remote Sens. Environ. 143, 85–96 (2014).Article 

    Google Scholar 
    Giglio, L., Boschetti, L., Roy, D. P., Humber, M. L. & Justice, C. O. The Collection 6 MODIS burned area mapping algorithm and product. Remote Sens. Environ. 217, 72–85 (2018).Article 

    Google Scholar 
    Roy, D. P., Boschetti, L., Justice, C. O. & Ju, J. The Collection 5 MODIS burned area product — Global evaluation by comparison with the MODIS active fire product. Remote Sens. Environ. 112, 3690–707 (2008).Article 

    Google Scholar 
    Miettinen, J., Langner, A. & Siegert, F. Burnt area estimation for the year 2005 in Borneo using multi-resolution satellite imagery. Int. J. Wildland Fire 16, 45–53 (2007).Luo, R., Hui, D., Miao, N., Liang, C. & Wells, N. Global relationship of fire occurrence and fire intensity: a test of intermediate fire occurrence-intensity hypothesis. J. Geophys. Res. Biogeosci. 122, 1123–36 (2017).Article 

    Google Scholar 
    Andela, N. et al. The Global Fire Atlas of individual fire size, duration, speed, and direction. Earth Syst. Sci. Data 11, 529–52 (2019).Article 

    Google Scholar 
    Andela, N., Morton, D. C., Giglio, L. & Randerson, J. T. Global Fire Atlas with Characteristics of Individual Fires, 2003-2016, ORNL Distributed Active Archive Center, https://doi.org/10.3334/ORNLDAAC/1642, https://daac.ornl.gov/cgi-bin/dsviewer.pl?ds_id=1642 (2019).Field, R. D. & Shen, S. S. P. Predictability of carbon emissions from biomass burning in Indonesia from 1997 to 2006. J. Geophys. Res. Biogeosci. 113, G04024 (2008).Article 
    CAS 

    Google Scholar 
    Fuller, D. O. & Murphy, K. The ENSO-fire dynamic in insular Southeast Asia. Clim. Change 74, 435–55 (2006).Article 

    Google Scholar 
    Field, R. D. et al. Development of a global fire weather database. Nat. Hazards Earth Syst. Sci. 15, 1407–23 (2015).Article 

    Google Scholar 
    Huffman, G. J. GPM IMERG Final Precipitation gridded data, L3 1 month 0.1 degree x 0.1 degree, version 06B. NASA Precipitation Processing System, Goddard Earth Sciences Data and Information Services Center (GES DISC). https://storm-pps.gsfc.nasa.gov/storm/; https://pmm.nasa.gov/data-access/downloads/gpm (2019).Huffman, G. J. et al. The TRMM Multisatellite Precipitation Analysis (TMPA): quasi-global, multiyear, combined-sensor precipitation estimates at fine scales. J. Hydrometeorol. 8, 38–55 (2007).Article 

    Google Scholar 
    Funk, C. et al. The climate hazards infrared precipitation with stations—a new environmental record for monitoring extremes. Sci. Data 2, 150066 (2015).Article 

    Google Scholar 
    Hsu, J., Huang, W.-R., Liu, P.-Y. & Li, X. Validation of CHIRPS precipitation estimates over taiwan at multiple timescales. Remote Sens. 13, 254 (2021).Rozante, J. R., Vila, D. A., Barboza Chiquetto, J., Fernandes, A. D. A. & Souza Alvim, D. Evaluation of TRMM/GPM blended daily products over Brazil. Remote Sens. 10, 882 (2018).Prakash, S., Mitra, A. K., Pai, D. S. & AghaKouchak, A. From TRMM to GPM: how well can heavy rainfall be detected from space? Adv. Water Resour. 88, 1–7 (2016).Article 

    Google Scholar 
    Ma, Q. et al. Performance evaluation and correction of precipitation data using the 20-year IMERG and TMPA precipitation products in diverse subregions of China. Atmos. Res. 249, 105304 (2021).Article 

    Google Scholar 
    Nwachukwu, P. N., Satge, F., Yacoubi, S. E., Pinel, S. & Bonnet, M.-P. From TRMM to GPM: how reliable are satellite-based precipitation data across Nigeria? Remote Sens. 12, 3964 (2020).Popovych, V. F. & Dunaieva, I. A. Assessment of the GPM IMERG and CHIRPS precipitation estimations for the steppe part of the Crimea. Meteorol. Hydrol. Water Manage 9, (2021).Navarro, A. et al. Assessment of IMERG precipitation estimates over Europe. Remote Sens. 11, 2470 (2019).Dezfuli, A. K. et al. Validation of IMERG precipitation in Africa. J. Hydrometeorol. 18, 2817–25 (2017).Article 

    Google Scholar 
    Efron, B. & Tibshirani, R. J. An Introduction to the Bootstrap. (Chapman and Hall, Boca Raton, FL, USA, 1993).Pérez-Hoyos, A., Rembold, F., Kerdiles, H. & Gallego, J. Comparison of global land cover datasets for cropland monitoring. Remote Sens. 9, 1118 (2017).ESA. Annual land-cover product, 1992 to 2019/present, based on MERIS 300-m and ancillary SPOT, AVHRR, Sentinel-3 and PROB-V satellite data. European Space Agency (ESA) European Centre for Medium-Range Weather Forecasts (ECMFW) Copernicus Climate Change Service (C3S) Climate Change Initiative (CCI), https://cds.climate.copernicus.eu/cdsapp#!/dataset/satellite-land-cover?tab=overview; http://maps.elie.ucl.ac.be/CCI/viewer/download.php; http://www.esa-landcover-cci.org/ (2020).Defourny, P. Product User Guide and Specification: ICDR Land Cover 2016 to 2019 (Version 2.1.1 of ESA Coperninus Climate Change Intitiative Annual 300-m Land-Cover Classifications). (Universitie Catholique du Lovain, Louvain, Belgium, (2020).Vetrita, Y. & Cochrane, M. A. Annual Burned Area from Landsat, Mawas, Central Kalimantan, Indonesia, 1997-2015, ORNL Distributed Active Archive Center, https://doi.org/10.3334/ORNLDAAC/1708, https://daac.ornl.gov/CMS/guides/Annual_Burned_Area_Maps.html; https://daac.ornl.gov/cgi-bin/dataset_lister.pl?p=33 (2019). More

  • in

    Ninety years of coastal monitoring reveals baseline and extreme ocean temperatures are increasing off the Finnish coast

    IPCC, 2014, Climate Change 2014: Synthesis Report. Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change.Bindoff, N. L. et al. Changing Ocean, Marine Ecosystems, and Dependent Communities. IPCC Spec. Rep. Ocean Cryosph. a Chang. Clim. [H.-O. Pörtner, D.C. Roberts, V. Masson-Delmotte, P. Zhai, M. Tignor, E. Poloczanska, K. Mintenbeck, A. Alegría, M. Nicolai, A. Okem, J. Petzold, B. Rama, N.M. Weyer (eds.)]. Press 447–588 (2019).Cheng, L. et al. Upper Ocean Temperatures Hit Record High in 2020. Adv. Atmos. Sci. 38, 523–530 (2021).Article 

    Google Scholar 
    Smale, D. A. et al. Marine heatwaves threaten global biodiversity and the provision of ecosystem services. Nat. Clim. Chang. 9, 306–312 (2019).Article 

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

    Google Scholar 
    Garrabou, J. et al. Mass mortality in Northwestern Mediterranean rocky benthic communities: Effects of the 2003 heat wave. Glob. Chang. Biol. 15, 1090–1103 (2009).Article 

    Google Scholar 
    Frölicher, T. L. & Laufkötter, C. Emerging risks from marine heat waves. Nat. Commun. 9, 2015–2018 (2018).Article 

    Google Scholar 
    Oliver, E. C. J. et al. Longer and more frequent marine heatwaves over the past century. Nat. Commun. https://doi.org/10.1038/s41467-018-03732-9 (2018).Article 

    Google Scholar 
    Garcia-Herrera, R., Díaz, J., Trigo, R. M., Luterbacher, J. & Fischer, E. M. A review of the european summer heat wave of 2003. Crit. Rev. Environ. Sci. Technol. 40, 267–306 (2010).Article 

    Google Scholar 
    Marbà, N., Jordà, G., Agustí, S., Girard, C. & Duarte, C. M. Footprints of climate change on Mediterranean Sea biota. Front. Mar. Sci. 2, 56 (2015).Holbrook, N. J. et al. Keeping pace with marine heatwaves. Nat. Rev. Earth Environ. https://doi.org/10.1038/s43017-020-0068-4 (2020). in press.Article 

    Google Scholar 
    Oliver, E. C. J., Wernberg, T., Benthuysen, J., Chen, K. & Eds. Advances in Understanding Marine Heatwaves and Their Impacts. Lausanne: Frontiers Media SA. vol. 7 (2020).Smale, D. A. & Wernberg, T. Satellite-derived SST data as a proxy for water temperature in nearshore benthic ecology. Mar. Ecol. Prog. Ser. 387, 27–37 (2009).Article 

    Google Scholar 
    Schlegel, R. W., Oliver, E. C. J., Wernberg, T. & Smit, A. J. Nearshore and offshore co-occurrence of marine heatwaves and cold-spells. Prog. Oceanogr. 151, 189–205 (2017).Article 

    Google Scholar 
    Rutgersson, A., Jaagus, J., Schenk, F. & Stendel, M. Observed changes and variability of atmospheric parameters in the Baltic Sea region during the last 200 years. Clim Res. 61, 177–190 (2014).Liblik, T. & Lips, U. Stratification has strengthened in the baltic sea – an analysis of 35 years of observational data. Front. Earth Sci. 7, 1–15 (2019).Article 

    Google Scholar 
    Reusch, T. B. H. et al. The Baltic Sea as a time machine for the future coastal ocean. Sci. Adv. 4, eaar8195 (2018).Hu, S. et al. Observed strong subsurface marine heatwaves in the tropical western Pacific Ocean. Environ. Res. Lett. 16, 104024 (2021).Scannell, H. A., Johnson, G. C., Thompson, L., Lyman, J. M. & Riser, S. C. Subsurface Evolution and Persistence of Marine Heatwaves in the Northeast Pacific. Geophys. Res. Lett. 47, 1–10 (2020).Article 

    Google Scholar 
    Schaeffer, A. & Roughan, M. Subsurface intensification of marine heatwaves off southeastern Australia: The role of stratification and local winds. Geophys. Res. Lett. 44, 5025–5033 (2017).Article 

    Google Scholar 
    WMO, Guide to Climatological Practices. (2018).Hobday, A. J. et al. Categorizing and naming marine heatwaves. Oceanography 31, 162–173 (2018).Article 

    Google Scholar 
    Zanna, L., Khatiwala, S., Gregory, J. M., Ison, J. & Heimbach, P. Global reconstruction of historical ocean heat storage and transport. Proc. Natl. Acad. Sci. U. S. A. 116, 1126–1131 (2019).CAS 
    Article 

    Google Scholar 
    Reynolds, R. W. et al. Daily high-resolution-blended analyses for sea surface temperature. J. Clim. 1, 5473–5496 (2007).Veneranta, L., Vanhatalo, J. & Urho, L. Detailed temperature mapping–Warming characterizes archipelago zones. Estuar. Coast. Shelf Sci. 182, 123–135 (2016).Article 

    Google Scholar 
    Merkouriadi, I. & Leppäranta, M. Long-term analysis of hydrography and sea-ice data in Tvärminne, Gulf of Finland, Baltic Sea. Clim. Change 124, 849–859 (2014).CAS 
    Article 

    Google Scholar 
    Woolway, R. I. et al. Lake heatwaves under climate change. Nature 589, 402–407 (2021).CAS 
    Article 

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

    Google Scholar 
    Rey, J., Rohat, G., Perroud, M., Goyette, S. & Kasparian, J. Shifting velocity of temperature extremes under climate change. Environ. Res. Lett. 15, 034027 (2020).Oliver, E. C. J. et al. Marine Heatwaves. Ann. Rev. Mar. Sci. 13, 313–342 (2021).Article 

    Google Scholar 
    Bennett, J. M. et al. The evolution of critical thermal limits of life on Earth. Nat. Commun. 1–9 (2021) https://doi.org/10.1038/s41467-021-21263-8.Holbrook, N. J. et al. A global assessment of marine heatwaves and their drivers. Nat. Commun. 10, 1–13 (2019).CAS 
    Article 

    Google Scholar 
    Kniebusch, M., Meier, H. E. M., Neumann, T. & Börgel, F. Temperature variability of the baltic sea since 1850 and attribution to atmospheric forcing variables. J. Geophys. Res. Ocean. 124, 4168–4187 (2019).Article 

    Google Scholar 
    Merkouriadi, I. & Leppäranta, M. Influence of sea ice on the seasonal variability of hydrography and heat content in Tvärminne, Gulf of Finland. Ann. Glaciol. 56, 274–284 (2015).Article 

    Google Scholar 
    Haapala, J. Upwelling and its influence on nutrient concentration in the coastal area of the Hanko Peninsula, entrance of the Gulf of Finland. Estuarine, Coastal and Shelf Science 38, 507–521 (1994).CAS 
    Article 

    Google Scholar 
    Sorte, C. J. B., Fuller, A. & Bracken, M. E. S. Impacts of a simulated heat wave on composition of a marine community. Oikos 119, 1909–1918 (2010).Article 

    Google Scholar 
    Pansch, C. et al. Heat waves and their significance for a temperate benthic community: A near-natural experimental approach. Glob. Chang. Biol. 24, 4357–4367 (2018).Article 

    Google Scholar 
    Morón Lugo, S. C. et al. Warming and temperature variability determine the performance of two invertebrate predators. Sci. Rep. 10, 1–14 (2020).Article 

    Google Scholar 
    Humborg, C. et al. High emissions of carbon dioxide and methane from the coastal Baltic Sea at the end of a summer heat wave. Front. Mar. Sci. 6, 1–14 (2019).Article 

    Google Scholar 
    Laakso, L. et al. 100 Years of atmospheric and marine observations at the Finnish Utö Island in the Baltic Sea. Ocean Sci. 14, 617–632 (2018).Article 

    Google Scholar 
    Høyer, J. L. & Karagali, I. Sea surface temperature climate data record for the North Sea and Baltic Sea. J. Clim. 29, 2529–2541 (2016).Article 

    Google Scholar 
    Schlegel, R. W. & Smit, A. J. heatwaveR: A central algorithm for the detection of heatwaves and cold-spells. J. Open Source Softw. 3, 821 (2018).Article 

    Google Scholar 
    Schlegel, R. W., Oliver, E. C. J., Hobday, A. J. & Smit, A. J. Detecting Marine Heatwaves With Sub-Optimal Data. Front. Mar. Sci. 6, 1–14 (2019).Article 

    Google Scholar  More

  • in

    Climate change increases global risk to urban forests

    Liu, Z., He, C., Zhou, Y. & Wu, J. How much of the world’s land has been urbanized, really? A hierarchical framework for avoiding confusion. Landsc. Ecol. 29, 763–771 (2014).
    Google Scholar 
    The World’s Cities in 2018: Data Booklet (UN, 2018).Miller, R. W., Hauer, R. J. & Werner, L. P. Urban Forestry: Planning and Managing Urban Greenspaces 3rd edn (Waveland Press, 2015).Escobedo, F. J., Kroeger, T. & Wagner, J. E. Urban forests and pollution mitigation: analyzing ecosystem services and disservices. Environ. Pollut. 159, 2078–2087 (2011).CAS 

    Google Scholar 
    Keeler, B. L. et al. Social-ecological and technological factors moderate the value of urban nature. Nat. Sustain. 2, 29 (2019).
    Google Scholar 
    Petri, A. C., Koeser, A. K., Lovell, S. T. & Ingram, D. How green are trees?—using life cycle assessment methods to assess net environmental benefits. J. Environ. Hortic. 34, 101–110 (2016).CAS 

    Google Scholar 
    Bastin, J.-F. et al. The global tree restoration potential. Science 365, 76–79 (2019).CAS 

    Google Scholar 
    IPCC Climate Change 2021: The Physical Science Basis (eds Masson-Delmotte, V. et al.) (Cambridge Univ. Press, 2021).Van Mantgem, P. J. et al. Widespread increase of tree mortality rates in the western United States. Science 323, 521–524 (2009).
    Google Scholar 
    Nowak, D. J. & Greenfield, E. J. Declining urban and community tree cover in the United States. Urban For. Urban Green. 32, 32–55 (2018).
    Google Scholar 
    Easterling, D. R. et al. Climate extremes: observations, modeling, and impacts. Science 289, 2068–2074 (2000).CAS 

    Google Scholar 
    Zscheischler, J. et al. Future climate risk from compound events. Nat. Clim. Change 8, 469–477 (2018).
    Google Scholar 
    Yan, P. & Yang, J. Performances of urban tree species under disturbances in 120 cities in China. Forests 9, 50 (2018).
    Google Scholar 
    Hilbert, D., Roman, L., Koeser, A. K., Vogt, J. & Van Doorn, N. S. Urban tree mortality: a literature review. Arboric. Urban For. 45, 167–200 (2019).
    Google Scholar 
    Young, R. F. & McPherson, E. G. Governing metropolitan green infrastructure in the United States. Landsc. Urban Plan. 109, 67–75 (2013).
    Google Scholar 
    Esperon-Rodriguez, M. et al. Assessing climate risk to support urban forests in a changing climate. Plants People Planet https://doi.org/10.1002/ppp3.10240 (2022).Esperon-Rodriguez, M. et al. Assessing the vulnerability of Australia’s urban forests to climate extremes. Plants People Planet 1, 387–397 (2019).Gallagher, R. V., Allen, S. & Wright, I. J. Safety margins and adaptive capacity of vegetation to climate change. Sci. Rep. 9, 8241 (2019).
    Google Scholar 
    Bertrand, R. et al. Changes in plant community composition lag behind climate warming in lowland forests. Nature 479, 517–520 (2011).CAS 

    Google Scholar 
    Bertrand, R. et al. Ecological constraints increase the climatic debt in forests. Nat. Commun. 7, 12643 (2016).Richard, B. et al. The climatic debt is growing in the understory of temperate forests: stand characteristics matter. Global Ecol. Biogeogr. 30, 1474–1487 (2021).IPCC Climate Change 2001: The Scientific Basis (eds Houghton, J. T. et al.) (Cambridge Univ. Press, 2001).Dawson, T. P., Jackson, S. T., House, J. I., Prentice, I. C. & Mace, G. M. Beyond predictions: biodiversity conservation in a changing climate. Science 332, 53–58 (2011).CAS 

    Google Scholar 
    Foden, W. B. et al. Climate change vulnerability assessment of species. WIREs Clim. Change 10, e551 (2019).
    Google Scholar 
    Pacifici, M. et al. Assessing species vulnerability to climate change. Nat. Clim. Change 5, 215–224 (2015).
    Google Scholar 
    Reisinger, A. et al. The Concept of Risk in the IPCC Sixth Assessment Report: A Summary of Cross-Working Group Discussions (IPCC, 2020).Chen, C. et al. University of Notre Dame Global Adaptation Index: Country Index Technical Report (ND-GAIN, 2015).McPherson, E. G., Berry, A. M. & van Doorn, N. S. Performance testing to identify climate-ready trees. Urban For. Urban Green. 29, 28–39 (2018).
    Google Scholar 
    Soberón, J. & Peterson, A. T. Interpretation of models of fundamental ecological niches and species’ distributional areas. Biodivers. Inform. 2 https://doi.org/10.17161/bi.v2i0.4 (2005).Pulliam, H. R. On the relationship between niche and distribution. Ecol. Lett. 3, 349–361 (2000).
    Google Scholar 
    Ordóñez, C. & Duinker, P. Assessing the vulnerability of urban forests to climate change. Environ. Rev. 22, 311–321 (2014).
    Google Scholar 
    Gallagher, R. V., Beaumont, L. J., Hughes, L. & Leishman, M. R. Evidence for climatic niche and biome shifts between native and novel ranges in plant species introduced to Australia. J. Ecol. 98, 790–799 (2010).
    Google Scholar 
    Smith, I. A., Dearborn, V. K. & Hutyra, L. R. Live fast, die young: accelerated growth, mortality, and turnover in street trees. PLoS ONE 14, e0215846 (2019).
    Google Scholar 
    Hirabayashi, Y., Kanae, S., Emori, S., Oki, T. & Kimoto, M. Global projections of changing risks of floods and droughts in a changing climate. Hydrol. Sci. J. 53, 754–772 (2008).
    Google Scholar 
    Van der Veken, S., Hermy, M., Vellend, M., Knapen, A. & Verheyen, K. Garden plants get a head start on climate change. Front. Ecol. Environ. 6, 212–216 (2008).
    Google Scholar 
    Ballinas, M. & Barradas, V. L. Transpiration and stomatal conductance as potential mechanisms to mitigate the heat load in Mexico City. Urban For. Urban Green. 20, 152–159 (2016).
    Google Scholar 
    Di Baldassarre, G. et al. Water shortages worsened by reservoir effects. Nat. Sustain. 1, 617 (2018).
    Google Scholar 
    Hoekstra, A. Y. & Mekonnen, M. M. The water footprint of humanity. Proc. Natl Acad. Sci. USA 109, 3232–3237 (2012).CAS 

    Google Scholar 
    Manoli, G. et al. Magnitude of urban heat islands largely explained by climate and population. Nature 573, 55–60 (2019).CAS 

    Google Scholar 
    Kim, D.-H., Doyle, M. R., Sung, S. & Amasino, R. M. Vernalization: winter and the timing of flowering in plants. Annu. Rev. Cell Dev. Biol. 25, 277–299 (2009).CAS 

    Google Scholar 
    Kummu, M. & Varis, O. The world by latitudes: a global analysis of human population, development level and environment across the north–south axis over the past half century. Appl. Geogr. 31, 495–507 (2011).
    Google Scholar 
    Vogt, J. et al. Citree: a database supporting tree selection for urban areas in temperate climate. Landsc. Urban Plan. 157, 14–25 (2017).
    Google Scholar 
    Paquette, A. et al. Praise for diversity: a functional approach to reduce risks in urban forests. Urban For. Urban Green. 62, 127157 (2021).
    Google Scholar 
    Esperon-Rodriguez, M. et al. Functional adaptations and trait plasticity of urban trees along a climatic gradient. Urban For. Urban Green. 54, 126771 (2020).
    Google Scholar 
    Hirons, A. D. et al. Using botanic gardens and arboreta to help identify urban trees for the future. Plants People Planet 3, 182–193 (2021).
    Google Scholar 
    Watkins, H., Hirons, A., Sjöman, H., Cameron, R. & Hitchmough, J. D. Can trait-based schemes be used to select species in urban forestry? Front. Sustain. Cities 3 https://doi.org/10.3389/frsc.2021.654618 (2021).Populated Places (Natural Earth, accessed 2018); http://www.naturalearthdata.com/downloads/Ossola, A. et al. The Global Urban Tree Inventory: a database of the diverse tree flora that inhabits the world’s cities. Glob. Ecol. Biogeogr. 29, 1907–1914 (2020).
    Google Scholar 
    Sabatini, F., Lenoir, J. & Bruelheide, H. sPlotOpen—An Environmentally-Balanced, Open-Access, Global Dataset of Vegetation Plots (iDiv, 2021); https://doi.org/10.25829/idiv.3474-40-3292Sabatini, F. M. et al. sPlotOpen—an environmentally balanced, open-access, global dataset of vegetation plots. Global Ecol. Biogeogr. 30, 1740–1764 (2021).Zizka, A. et al. CoordinateCleaner: standardized cleaning of occurrence records from biological collection databases. Methods Ecol. Evol. 10, 744–751 (2019).
    Google Scholar 
    R Core Team. R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, 2021).Taxonstand: Taxonomic standardization of plant species names. R package version 2.4 https://cran.r-project.org/web/packages/Taxonstand/Taxonstand.pdf (2021).Kelso, N. & Patterson, T. World Urban Areas, LandScan, 1:10 Million (2012) (North American Cartographic Information Society, 2012).Karger, D. N. et al. Climatologies at high resolution for the earth’s land surface areas. Sci. Data 4, 170122 (2017).
    Google Scholar 
    O’Donnell, M. S. & Ignizio, D. A. Bioclimatic Predictors for Supporting Ecological Applications in the Conterminous United States (USGS, 2012).Field, C. et al. IPCC, 2014: Climate Change 2014: Impacts, Adaptation, and Vulnerability. Part A: Global and Sectoral Aspects. Contribution of Working Group II to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change (Cambridge Univ. Press, 2014).Meinshausen, M. et al. The RCP greenhouse gas concentrations and their extensions from 1765 to 2300. Clim. Change 109, 213–241 (2011).CAS 

    Google Scholar 
    Zhao, L. et al. Global multi-model projections of local urban climates. Nat. Clim. Change 11, 152–157 (2021).
    Google Scholar 
    Huang, K., Li, X., Liu, X. & Seto, K. C. Projecting global urban land expansion and heat island intensification through 2050. Environ. Res. Lett. 14, 114037 (2019).
    Google Scholar 
    Alavipanah, S., Wegmann, M., Qureshi, S., Weng, Q. & Koellner, T. The role of vegetation in mitigating urban land surface temperatures: a case study of Munich, Germany during the warm season. Sustainability 7, 4689–4706 (2015).
    Google Scholar 
    Corburn, J. Cities, climate change and urban heat island mitigation: localising global environmental science. Urban Stud. 46, 413–427 (2009).
    Google Scholar 
    Baston, D., ISciences, L.L., Baston, M.D. Package ‘exactextractr’. terra. R package version 0.8.2 (2022).Hijmans, R. J. et al. raster: Geographic data analysis and modeling. R package version 2.3-33 http://cran.r-project.org/web/packages/raster/index.html (2016).Bates, D., Mächler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1–48 (2015).
    Google Scholar 
    Bivand, R. et al. maptools: Tools for handling spatial objects. R package version 08, 23 https://cran.r-project.org/web/packages/maptools/ (2013). More

  • in

    Evaluation of animal and plant diversity suggests Greenland’s thaw hastens the biodiversity crisis

    Species occurrence recordsWe compiled data on the distribution of 21,252 endemic species of any of the twelve megadiverse countries from four tetrapod (5,757) and four vascular plant groups (15,389) (amphibians, reptiles, birds, mammals, lycophytes, ferns, gymnosperms, and flowering plants). Species occurrence records were obtained from the Global Biodiversity Information Facility (GBIF)27, the International Union of Conservation of Nature (IUCN)28, and BirdLife60,61. We only modeled species with at least 25 unique records at a 5 arc-minute resolution (~10 km at the equator). In many cases, the processing of the IUCN polygons resulted in species with thousands of occurrence records. In these cases, we randomly chose a maximum of 500 records per species. The greater the number of observed records, more problems can be associated with spatial bias in the modeling62. In the case of records coming from IUCN polygons, more records require more computing time and these do not necessarily provide more information into the modeling given that their distribution is quite homogeneous.For tetrapods, we first explored the possibility of using occurrence records from GBIF, but data for megadiverse countries were scarce. Consequently, we decided to use the distribution polygons provided by the IUCN for amphibians, reptiles, and mammals (terrestrial and freshwater species)28, and the distribution polygons provided by BirdLife60. We based this decision on the fact that ecological niche modeling using IUCN polygons has been proven to give robust results20. For the IUCN polygons, we retained species that have been categorized as “extant”, “possibly extinct”, “probably extant”, “possibly extant”, and “presence uncertain”, discarding species considered to be “extinct”. In addition, we did not model species reported by the IUCN as “introduced”, “vagrant”, or those in the “assisted colonization” category; for mammals and birds, we only considered the distribution of “resident” species. Depending on the taxonomic group, and given the information available, we used different approaches to identify species endemic to any of twelve megadiverse countries: Australia, Brazil, China, Colombia, Ecuador, India, Indonesia, Madagascar, Mexico, Peru, Philippines, and Venezuela. For birds, we used BirdLife to identify species listed as “breeding endemic” and then choose the corresponding IUCN polygons. To identify the rest of endemic species in the other groups, we used a 0.08333° buffer around each country to select the IUCN polygons that fall completely within the country limits. We converted all selected species polygons into unique records at a 5 min resolution (~10 km at the equator).For vascular plants, we used geographic occurrence data obtained from the Global Biodiversity Information Facility by querying all records under “Tracheophyta” (we only considered “Preserved Specimens” in our search). Plants records were taxonomically homogenized and cleaned following the procedures described in ref. 63 using Kew’s Plants of the World database64 as the source of taxonomic information. Mostly, we identified endemic species as those with all occurrence records restricted to any given megadiverse country. For countries in which data for vascular plants were scarce or absent (e.g., India), we complemented occurrence information with polygons from the IUCN (although IUCN data for plants remains limited) following the procedure described for tetrapods.Climatic dataWe used the 19 bioclimatic variables available at WorldClim v.2 (Fick 2017) as the baseline (present-day) climatic conditions (1970–2000) (annual mean temperature, mean diurnal range, isothermality, temperature seasonality, the maximum temperature of the warmest month, minimum temperature of the coldest month, temperature annual range, mean annual range, mean temperature of wettest quarter, mean temperature of driest quarter, mean temperature of warmest quarter, mean temperature of coldest quarter, annual precipitation, precipitation of wettest month, precipitation of driest month, precipitation seasonality, precipitation of wettest quarter, precipitation of driest quarter, precipitation of warmest quarter and precipitation of coldest quarter). From this baseline scenario, bioclimatic variables start to vary because of climate change. We used bioclimatic variables derived from the IPSL-CM5-LR ocean-atmospheric model under five scenarios: (i) the high-emissions RCP 8.5 W/m2; and (ii) melting scenarios consisting of four different experiments of freshwater discharge into the North Atlantic from Greenland’s meltwater (see DeFrance16 for details). We acknowledge that using a single GCM does not allow us to estimate inter-GCM variability in the resulting distribution models; however, the melting scenarios do only exist for IPSL-CM5-LR GCM. We applied as control scenario RCP 8.5 because melting scenarios would have been more complicated to support with lower emission scenarios. In addition, we are using well-designed opportunity experiments from ref. 11 and wanted to be consistent with their choice of RCP 8.5. Also, these experiments are based on CMIP5, which shows similar climate impact fingerprints than CMIP665. This might be explained by the fact that CMIP5 and CMIP6 are still relatively close, and that the main climatic effects of the AMOC are already well-represented by the climate dynamics in CMIP5.The four melting scenarios are equivalent to a sea-level rise of 0.5, 1.0, 1.5, and 3.0 meters above the current sea level, and these are named accordingly: Melting 0.5, Melting 1.0, Melting 1.5., and Melting 3.0. These AMOC scenarios are experiments that were superimposed to the RCP 8.5 scenario adding 0.11, 0.22, 0.34, and 0.68 Sv (1 Sv = 106 m3/s) coming from a freshwater release that starts in 2020 and finishes in 2070 (Anthoff et al.14). We obtained debiased bioclimatic variables11 under the five future scenarios for three consecutive time horizons: T1: 2030 (2030–2060); T2: 2050 (2050–2080); and T3: 2070 (2070–2100). The time horizons evaluated represent short, medium, and long terms in order to help decision-makers order conservation priorities.Ecological niche modelingAt their most basic, the algorithms used to construct species distribution models relate species occurrence records with climatic variables to create a climatic profile that can be projected onto other time periods and geographic regions66. The resulting models have proven useful in evaluating the impacts of climate change on biodiversity and to identify varying levels of vulnerability among species32,67,68. Here, we employed a multi-algorithm (ensemble) approach to construct species distribution models as implemented in the “biomod2” package67 in R69 (Supplementary Fig. 33). The underlying philosophy of ensemble modeling is that each model carries a true “signal” about the climate-occurrence relationships we aim to capture, but it also carries “noise” created by biases and uncertainties in the data and model structure32,67. By combining models created with different algorithms, ensemble models aim at capturing the true “signal” while controlling for algorithm-derived model differences; therefore, model uncertainty is accounted for during model construction (see Supplementary Material for further detail).Prior to modeling, we reduced the number of bioclimatic variables per species by estimating collinearity among present-day bioclimatic variables. We employed the “corrSelect” function of the package fuzzySim70 in R69, using a Pearson correlation threshold of 0.8 and variance inflation factors as criteria to select variables. Given the number of species evaluated and the ecological information scarcity, we did not select a set of variables based on ecological knowledge by each of the species modeled. Instead, for the variables pre-selection, we used the statistical approach described above that has been proven to give models with good performance71,72. We used seven algorithms with a good predictive performance (evaluated with the TSS and ROC statistics; Supplementary Fig. 1): Maxent (MAXENT.Phillips), Generalized Additive Models (GAM), Classification Trees Analysis (CTA), Artificial Neural Networks (ANN), Surface Range Envelope (SRE), Flexible Discriminant Analysis (FDA), and Random Forest (RF). Because occurrence datasets consisted of presence-only data, for each model, we randomly generated 10,000 pseudo-absences within the model calibration area; we gave presences and absences the same importance during the calibration process (BIOMOD’s prevalence = 0.5). For each species, we selected a calibration area (i.e., the accessible area or M)73 using a spatial intersection between a 4° buffer around species occurrences and the terrestrial ecoregions occupied by the species73 (Supplementary Fig. 33). The projected M (i.e., the area accessible for species in future scenarios) was defined using a 2° buffer around the present-day calibration area (M). By limiting the M, we incorporated information about dispersal and ecological limitations of each species into the modeling66. We did this to take into account a more realistic dispersal scenario given the velocity with which climatic changes are happening and because there are geographic and ecological barriers, which is the reason why we used ecoregions to limit our M. We assumed climatic niche conservatism across time; and inside the projected M we also assumed full dispersal. Consequently, inside the projected M, the evaluated species can win or lose suitable climatic conditions.We calibrated each algorithm using a random sample of 70% of occurrence records and evaluated the resulting models using the remaining 30% of records. To validate the predictive power of the ecological niche models, we used the True Skill Statistics (TSS) and the Receiver Operating Characteristics (ROC) and performed 10 replicates for every model, providing a tenfold internal cross-validation. To account for uncertainty, we constructed the ensemble models (seven algorithms × ten replicates) using a total consensus rule, where models from different algorithms were assembled using a weighted mean of replicates with an evaluation threshold of AUC  > 0.7 (Supplementary Fig. 1). However, as shown by the distribution of validation statistic in Supplementary Fig. 1, most ensemble models presented a very good predictive power (AUC  > 0.8). In some cases, modeling issues in some insular species required that we change the calibration area (M) to the entire country.We used the resulting ensemble models to project the potential distribution of each species under both current and future climatic conditions (Supplementary Fig. 34). We then examined the frequency in which different bioclimatic variables appeared to have the highest contribution during model construction for each species. The algorithms used (Maxent, GAM, CTA, ANN, SRE, FDA, and RF) identify these variables by iteratively testing combinations of all the available variables (i.e., those selected based on low correlation values) until reaching a set of variables that was most informative on the distribution of species; this set of variables had the highest predictive power of species occurrence. For every species, we retrieved the two variables with the largest model contribution (Supplementary Figs. 34 and 35).Species geographic rangeWe converted ensemble probability maps into binary maps of presence/absence using the TSS threshold; these binary maps reflect the distribution of climatic suitability of species, where values of 0 and 1 represent grid cells with non-suitable and suitable climates, respectively. In order to approximate the vulnerability of individual species to climate change, we estimated the temporal changes in the extent of the area of climatic suitability (geographic range) for every species relative to the present-day distribution. We estimated species’ geographic ranges by identifying and counting those grid cells with suitable climatic conditions (values of 1) in the present-day and under future scenarios. We then estimated the proportion of range changes through time, quantifying the proportion of grid cells either lost or gained for each species. This allowed us to estimate the proportion of species (by country and group) projected to have a complete loss of geographic ranges in the future.Species richness, differences in species richness, potential species hotspots (PSH), and temporal dissimilarityWe used binary maps to construct presence-absence matrices (PAM), which contain information on the presence (values of 1) or absence (values of 0) of species across grid cells. Using these PAMs, we estimated species richness (SR) as the sum of species present in each grid cell; to visualize SR across space, we generated 16 species richness maps corresponding to the present-day and the four future scenarios at each of the three temporal horizons. We used these maps to estimate and visualize temporal differences in species richness (ΔSR) over time by subtracting the estimated SR in the future from the current SR, for every grid cell; for visualization, we standardized SR per country to the range 0–1. We assumed full dispersal ability of species in all analyses, meaning that all suitable areas in the future had the same probability of being occupied, irrespective of the distance to the present-day distribution.By calculating species richness (SR) across grid cells, we defined Potential Species Hotspots (PSH) within each country as those grid cells with the highest levels of SR. For this, we defined the PSH by calculating the maximum present-day species richness (maxSR) observed in each country and then identified grid cells with richness values above a threshold of maxSR*0.6. Considering only those grid cells with a SR above this threshold, we estimated the geographic extent of PSH across time periods and scenarios and estimated changes to the extent of PSH relative to present-day conditions. Given that we use the threshold to define PSHs, we tested two additional thresholds (20 and 90%) to define and quantify the extent of PSHs. However, these additional results agree with the general trend. We chose not to base our threshold on the distribution of SR values (i.e., quantiles, median) due to the high proportion of grid cells with SR  More

  • in

    Selection, drift and community interactions shape microbial biogeographic patterns in the Pacific Ocean

    Nelson G. From Candolle to croizat: comments on the history of biogeography. J Hist Biol. 1978;11:269–305.PubMed 
    Article 
    CAS 

    Google Scholar 
    Lomolino MV, Riddle BR, Whittaker RJ, Brown JH. Biogeography. Sunderland, MA: Sinauer Associates; 2005. p. 752Wang J, Soininen J, Zhang Y, Wang B, Yang X, Shen J. Contrasting patterns in elevational diversity between microorganisms and macroorganisms. J Biogeogr. 2011;38:595–603.Article 

    Google Scholar 
    Treseder KK, Maltz MR, Hawkins BA, Fierer N, Stajich JE, Mcguire KL. Evolutionary histories of soil fungi are reflected in their large-scale biogeography. Ecol Lett. 2014;17:1086–93.PubMed 
    Article 

    Google Scholar 
    Meyer KM, Memiaghe H, Korte L, Kenfack D, Alonso A, Bohannan BJM. Why do microbes exhibit weak biogeographic patterns? ISME J. 2018;12:1404–13.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Lindström ES, Langenheder S. Local and regional factors influencing bacterial community assembly. Environ Microbiol Rep. 2012;4:1–9.PubMed 
    Article 

    Google Scholar 
    Ghiglione JF, Galand PE, Pommier T, Pedrós-Alió C, Maas EW, Bakker K, et al. Pole-to-pole biogeography of surface and deep marine bacterial communities. Proc Natl Acad Sci USA 2012;109:17633–8.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Sul WJ, Oliver TA, Ducklow HW, Amaral-Zettlera LA, Sogin ML. Marine bacteria exhibit a bipolar distribution. Proc Natl Acad Sci USA 2013;110:2342–7.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Sunagawa S, Coelho LP, Chaffron S, Kultima JR, Labadie K, Salazar G, et al. Structure and function of the global ocean microbiome. Science. 2015;348:1261359.PubMed 
    Article 
    CAS 

    Google Scholar 
    de Vargas C, Audic S, Henry N, Decelle J, Mahé F, Logares R, et al. Eukaryotic plankton diversity in the sunlit ocean. Science. 2015;348:1261605.PubMed 
    Article 
    CAS 

    Google Scholar 
    Milici M, Tomasch J, Wos-Oxley ML, Decelle J, Jáuregui R, Wang H. et al. Bacterioplankton biogeography of the Atlantic ocean: a case study of the distance-decay relationship. Front Microbiol. 2016;7:Article 590.PubMed 

    Google Scholar 
    Raes EJ, Bodrossy L, Van De Kamp J, Bissett A, Ostrowski M, Brown MV, et al. Oceanographic boundaries constrain microbial diversity gradients in the south pacific ocean. Proc Natl Acad Sci USA 2018;115:8266–75.Article 
    CAS 

    Google Scholar 
    Wu W, Lu HP, Sastri A, Yeh YC, Gong GC, Chou WC, et al. Contrasting the relative importance of species sorting and dispersal limitation in shaping marine bacterial versus protist communities. ISME J. 2018;12:485–94.PubMed 
    Article 

    Google Scholar 
    Vellend M. Conceptual synthesis in community ecology. Q Rev Biol. 2010;85:183–206.PubMed 
    Article 

    Google Scholar 
    Hanson CA, Fuhrman JA, Horner-Devine MC, Martiny JBH. Beyond biogeographic patterns: Processes shaping the microbial landscape. Nat Rev Microbiol. 2012;10:497–506.PubMed 
    Article 
    CAS 

    Google Scholar 
    Nemergut DR, Schmidt SK, Fukami T, O’Neill SP, Bilinski TM, Stanish LF, et al. Patterns and processes of microbial community assembly. Microbiol Mol Biol Rev. 2013;77:342–56.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Stegen JC, Lin X, Fredrickson JK, Chen X, Kennedy DW, Murray CJ, et al. Quantifying community assembly processes and identifying features that impose them. ISME J. 2013;7:2069–79.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Schmidt TSB, Matias Rodrigues JF, Von Mering C. A family of interaction-adjusted indices of community similarity. ISME J. 2017;11:791–807.PubMed 
    Article 

    Google Scholar 
    Zhou J, Ning D. Stochastic community assembly: does it matter in microbial ecology? Microbiol Mol Biol Rev. 2017;81:1–32.Article 

    Google Scholar 
    Djurhuus A, Port J, Closek CJ, Yamahara KM, Romero-maraccini O, Walz KR. et al. Evaluation of filtration and DNA extraction methods for environmental DNA biodiversity assessments across multiple trophic levels. Front Mar Sci. 2017;4:Article 314.Article 

    Google Scholar 
    Wang ZB, Sun YY, Li Y, Chen XL, Wang P, Ding HT, et al. Significant bacterial distance-decay relationship in continuous, well-connected southern ocean surface water. Micro Ecol. 2020;80:73–80.Article 
    CAS 

    Google Scholar 
    Dlugosch L, Pohlein A, Wemheuer B, Pfeiffer B, Badewien T, Daniel R, et al. Significance of gene variants for the functional biogeography of the near-surface Atlantic Ocean microbiome. Nat Commun. 2022;13:456.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Lozupone C, Knight R. UniFrac: A new phylogenetic method for comparing microbial communities. Appl Environ Microbiol. 2005;71:8228–35.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Logares R, Deutschmann IM, Junger PC, Giner CR, Krabberød AK, Schmidt TSB, et al. Disentangling the mechanisms shaping the surface ocean microbiota. Microbiome. 2020;8:55.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Doblin MA, Petrou K, Sinutok S, Seymour JR, Messer LF, Brown MV, et al. Nutrient uplift in a cyclonic eddy increases diversity, primary productivity and iron demand of microbial communities relative to a western boundary current. PeerJ. 2016;4:e1973.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Polovina JJ, Howell E, Kobayashi DR, Seki MP. The transition zone chlorophyll front, a dynamic global feature defining migration and forage habitat for marine resources. Prog Oceanogr. 2001;49:469–83.Article 

    Google Scholar 
    Karl DM, Church MJ. Ecosystem structure and dynamics in the north pacific subtropical gyre: new views of an old ocean. Ecosystems. 2017;20:433–57.Article 

    Google Scholar 
    Mestre M, Ruiz-González C, Logares R, Duarte CM, Gasol JM, Sala MM. Sinking particles promote vertical connectivity in the ocean microbiome. Proc Natl Acad Sci USA 2018;115:6799–807.Article 
    CAS 

    Google Scholar 
    Balmonte JP, Simon M, Giebel HA, Arnosti C. A sea change in microbial enzymes: Heterogeneous latitudinal and depth-related gradients in bulk water and particle-associated enzymatic activities from 30°S to 59°N in the Pacific Ocean. Limnol Oceanogr. 2021;66:3489–507.Article 
    CAS 

    Google Scholar 
    Giebel H-A, Arnosti C, Badewien TH, Bakenhus I, Balmonte JP, Billerbeck S. et al. Microbial growth and organic matter cycling in the Pacific Ocean along a latitudinal transect between subarctic and subantarctic waters. Front Mar Sci. 2021;8:Article 764383.Article 

    Google Scholar 
    Milici M, Tomasch J, Wos-Oxley ML, Wang H, Jáuregui R, Camarinha-Silva A, et al. Low diversity of planktonic bacteria in the tropical ocean. Sci Rep. 2016;6:19054.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Longhurst AR. Ecological geography of the sea. San Diego, USA: Academic Press; 2007.Parada AE, Needham DM, Fuhrman JA. Every base matters: Assessing small subunit rRNA primers for marine microbiomes with mock communities, time series and global field samples. Environ Microbiol. 2016;18:1403–14.PubMed 
    Article 
    CAS 

    Google Scholar 
    Milke F, Sanchez-Garcia S, Dlugosch L, McNichol J, Fuhrman J, Simon M. et al. Composition and biogeography of pro- and eukaryotic communities in the Atlantic Ocean: primer choice matters. Front Microbiol. 2022;13:Article 895875.PubMed 
    Article 

    Google Scholar 
    Vaulot D, Geisen S, Mahé F, Bass D. pr2-primers: An 18S rRNA primer database for protists. Mol Ecol Resour. 2022;22:168–79.PubMed 
    Article 
    CAS 

    Google Scholar 
    Yeh YC, McNichol J, Needham DM, Fichot EB, Berdjeb L, Fuhrman JA. Comprehensive single-PCR 16S and 18S rRNA community analysis validated with mock communities, and estimation of sequencing bias against 18S. Environ Microbiol. 2021;23:3240–50.PubMed 
    Article 
    CAS 

    Google Scholar 
    Quast C, Pruesse E, Yilmaz P, Gerken J, Schweer T, Yarza P, et al. The SILVA ribosomal RNA gene database project: Improved data processing and web-based tools. Nucleic Acids Res. 2013;41:590–6.Article 
    CAS 

    Google Scholar 
    Guillou L, Bachar D, Audic S, Bass D, Berney C, Bittner L, et al. The Protist Ribosomal Reference database (PR2): a catalog of unicellular eukaryote Small Sub-Unit rRNA sequences with curated taxonomy. Nucleic Acids Res. 2013;41:597–604.Article 
    CAS 

    Google Scholar 
    Bolyen E, Rideout JR, Dillon MR, Bokulich NA, Abnet CC, Al-Ghalith GA, et al. Reproducible, interactive, scalable and extensible microbiome data science using QIIME 2 (Nature Biotechnology, (2019), 37, 8, (852-857), 10.1038/s41587-019-0209-9). Nat Biotechnol. 2019;37:1091.PubMed 
    Article 
    CAS 

    Google Scholar 
    Callahan BJ, McMurdie PJ, Rosen MJ, Han AW, Johnson AJA, Holmes SP. DADA2: High-resolution sample inference from Illumina amplicon data. Nat Methods. 2016;13:581–3.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Friedman J, Alm EJ. Inferring correlation networks from genomic survey data. PLoS Comput Biol. 2012;8:e1002687.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Bodenhofer U, Bonatesta E, Horejš-Kainrath C, Hochreiter S. Msa: an R package for multiple sequence alignment. Bioinformatics. 2015;31:3997–9.PubMed 
    CAS 

    Google Scholar 
    Kaufman L, Rousseeuw PJ. Finding groups in data: an introduction to cluster analysis. Hoboken NJ, USA: John Wiley & Sons; 2009.Pruesse E, Peplies J, Glöckner FO. SINA: Accurate high-throughput multiple sequence alignment of ribosomal RNA genes. Bioinformatics. 2012;28:1823–9.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Price MN, Dehal PS, Arkin AP. FastTree 2 – approximately maximum-likelihood trees for large alignments. PLoS ONE. 2010;5:e9490.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Losos JB. Phylogenetic niche conservatism, phylogenetic signal and the relationship between phylogenetic relatedness and ecological similarity among species. Ecol Lett. 2008;11:995–1003.PubMed 
    Article 

    Google Scholar 
    Stegen JC, Lin X, Konopka AE, Fredrickson JK. Stochastic and deterministic assembly processes in subsurface microbial communities. ISME J. 2012;6:1653–64.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Fine PVA, Kembel SW. Phylogenetic community structure and phylogenetic turnover across space and edaphic gradients in western Amazonian tree communities. Ecography. 2011;34:552–65.Article 

    Google Scholar 
    Chase JM, Kraft NJB, Smith KG, Vellend M, Inouye BD. Using null models to disentangle variation in community dissimilarity from variation in α-diversity. Ecosphere. 2011;2:1–11.Article 

    Google Scholar 
    NASA Goddard Space Flight Center, Ocean Ecology Laboratory OBPG. Moderate-resolution Imaging Spectroradiometer (MODIS) aqua chlorophyll data. https://oceancolor.gsfc.nasa.gov/data/10.5067/AQUA/MODIS/L3B/CHL/2018/. Accessed 13 Nov 2020.Pommier T, Douzery EJP, Mouillot D. Environment drives high phylogenetic turnover among oceanic bacterial communities. Biol Lett. 2012;8:562–6.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Giovannoni SJ, Cameron Thrash J, Temperton B. Implications of streamlining theory for microbial ecology. ISME J. 2014;8:1553–65.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Sañudo-Wilhelmy SA, Gómez-Consarnau L, Suffridge C, Webb EA. The role of B vitamins in marine biogeochemistry. Ann Rev Mar Sci. 2014;6:339–67.PubMed 
    Article 

    Google Scholar 
    Morris JJ, Lenski RE, Zinser ER. The black queen hypothesis: evolution of dependencies through adaptive gene loss. MBio. 2012;3:e00036–12.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Carini P, Campbell EO, Morré J, Sañudo-Wilhelmy SA, Cameron Thrash J, Bennett SE, et al. Discovery of a SAR11 growth requirement for thiamin’s pyrimidine precursor and its distribution in the Sargasso Sea. ISME J. 2014;8:1727–38.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Wienhausen G, Bruns S, Sultana S, Dlugosch L, Groon L, Wilkes H, et al. The overlooked role of a biotin precursor for marine bacteria – desthiobiotin as an escape route for biotin auxotrophy. ISME J. 2022. https://doi.org/10.1038/s41396-022-01304-w.Biller SJ, Coe A, Chisholm SW. Torn apart and reunited: Impact of a heterotroph on the transcriptome of Prochlorococcus. ISME J. 2016;10:2831–43.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Sokolovskaya OM, Shelton AN, Taga ME. Sharing vitamins: cobamides unveil microbial interactions. Science. 2020;369:eaba0165.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Wienhausen G, Dlugosch L, Jarling R, Wilkes H, Giebel H-A, Simon M. Availability of vitamin B12 and its lower ligand intermediate a-ribazole impact prokaryotic and protist communities in oceanic systems. ISME J. 2022;16:2002–14.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Reintjes G, Arnosti C, Fuchs B, Amann R. Selfish, sharing and scavenging bacteria in the Atlantic Ocean: a biogeographical study of bacterial substrate utilisation. ISME J. 2019;13:1119–32.PubMed 
    Article 
    CAS 

    Google Scholar 
    Bertrand EM, McCrow JP, Moustafa A, Zheng H, McQuaid JB, Delmont TO, et al. Phytoplankton-bacterial interactions mediate micronutrient colimitation at the coastal Antarctic sea ice edge. Proc Natl Acad Sci USA 2015;112:9938–43.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Amin SA, Hmelo LR, Van Tol HM, Durham BP, Carlson LT, Heal KR, et al. Interaction and signalling between a cosmopolitan phytoplankton and associated bacteria. Nature. 2015;522:98–101.PubMed 
    Article 
    CAS 

    Google Scholar 
    Shibl AA, Isaac A, Ochsenkühn MA, Cárdenas A, Fei C, Behringer G, et al. Diatom modulation of select bacteria through use of two unique secondary metabolites. Proc Natl Acad Sci USA 2020;117:27445–55.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Villarino E, Watson JR, Chust G, Woodill AJ, Klempay B, Jonsson B, et al. Global beta diversity patterns of microbial communities in the surface and deep ocean. Glob Ecol Biogeogr. 2022;00:1–14.
    Google Scholar 
    Cravatte S, Kestenare E, Marin F, Dutrieux P, Firing E. Subthermocline and intermediate zonal currents in the tropical Pacific Ocean: Paths and vertical structure. J Phys Oceanogr. 2017;47:2305–24.Article 

    Google Scholar 
    Cho BC, Azam F. Major role of bacteria in biogeochemical fluxes in the ocean’s interior. Nature. 1988;332:441–3.Article 
    CAS 

    Google Scholar 
    Salazar G, Cornejo-Castillo FM, Benítez-Barrios V, Fraile-Nuez E, Álvarez-Salgado XA, Duarte CM, et al. Global diversity and biogeography of deep-sea pelagic prokaryotes. ISME J. 2016;10:596–608.PubMed 
    Article 

    Google Scholar 
    Delmont TO, Kiefl E, Kilinc O, Esen OC, Uysal I, Rappé MS, et al. Single-amino acid variants reveal evolutionary processes that shape the biogeography of a global SAR11 subclade. Elife. 2019;8:e46497.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Hillebrand H. On the generallity of the latutinal diversity gradient. Am Nat. 2004;163:192–211.PubMed 
    Article 

    Google Scholar  More

  • in

    Genic distribution modelling predicts adaptation of the bank vole to climate change

    Davis, M. B. & Shaw, R. G. Range shifts and adaptive responses to Quaternary climate change. Science 292, 673–679 (2001).CAS 
    PubMed 
    Article 

    Google Scholar 
    Parmesan, C. Ecological and evolutionary responses to recent climate change. Annu. Rev. Ecol. Evol. Syst. 37, 637–669 (2006).Article 

    Google Scholar 
    Hewitt, G. The genetic legacy of the Quaternary ice ages. Nature 405, 907–913 (2000).CAS 
    PubMed 
    Article 

    Google Scholar 
    Williams, J. E. & Blois, J. L. Range shifts in response to past and future climate change: can climate velocities and species’ dispersal capabilities explain variation in mammalian range shifts? J. Biogeogr. 45, 2175–2189 (2018).Article 

    Google Scholar 
    Parmesan, C. & Yohe, G. A globally coherent fingerprint of climate change impacts across natural systems. Nature 421, 37–42 (2003).CAS 
    PubMed 
    Article 

    Google Scholar 
    Thomas, C. D. Climate, climate change and range boundaries. Divers. Distrib. 16, 488–495 (2010).Article 

    Google Scholar 
    Bradshaw, A. D. & McNeilly, T. Evolutionary response to global climatic change. Ann. Bot. 67, 5–14 (1991).Article 

    Google Scholar 
    Harter, D. E. V. et al. Impacts of global climate change on the floras of oceanic islands—projections, implications and current knowledge. Perspect. Plant Ecol. Evol. Syst. 17, 160–183 (2015).Article 

    Google Scholar 
    Veron, S., Haevermans, T., Govaerts, R., Mouchet, M. & Pellens, R. Distribution and relative age of endemism across islands worldwide. Sci. Rep. 9, 1–12 (2019).Article 
    CAS 

    Google Scholar 
    Román-Palacios, C. & Wiens, J. J. Recent responses to climate change reveal the drivers of species extinction and survival. Proc. Natl Acad. Sci. USA 117, 4211–4217 (2020).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Jump, A. S. & Peñuelas, J. Running to stand still: adaptation and the response of plants to rapid climate change. Ecol. Lett. 8, 1010–1020 (2005).PubMed 
    Article 

    Google Scholar 
    Freeman, B. G., Scholer, M. N., Ruiz-Gutierrez, V. & Fitzpatrick, J. W. Climate change causes upslope shifts and mountaintop extirpations in a tropical bird community. Proc. Natl Acad. Sci. USA 115, 11982–11987 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Gilbert, K. J. & Whitlock, M. C. The genetics of adaptation to discrete heterogeneous environments: frequent mutation or large-effect alleles can allow range expansion. J. Evol. Biol. 30, 591–602 (2017).CAS 
    PubMed 
    Article 

    Google Scholar 
    Christmas, M. J., Breed, M. F. & Lowe, A. J. Constraints to and conservation implications for climate change adaptation in plants. Conserv. Genet. 17, 305–320 (2015).Article 
    CAS 

    Google Scholar 
    Barrett, R. D. H. & Schluter, D. Adaptation from standing genetic variation. Trends Ecol. Evol. 23, 38–44 (2008).PubMed 
    Article 

    Google Scholar 
    Lai, Y. T. et al. Standing genetic variation as the predominant source for adaptation of a songbird. Proc. Natl Acad. Sci. USA 116, 2152–2157 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Hoban, S. et al. Finding the genomic basis of local adaptation: Pitfalls, practical solutions, and future directions. Am. Nat. 188, 379–397 (2016).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Hoffmann, A. A. & Sgrò, C. M. Climate change and evolutionary adaptation. Nature 470, 479–485 (2011).CAS 
    PubMed 
    Article 

    Google Scholar 
    Catullo, R. A., Llewelyn, J., Phillips, B. L. & Moritz, C. C. The potential for rapid evolution under anthropogenic climate change. Curr. Biol. 29, R996–R1007 (2019).CAS 
    PubMed 
    Article 

    Google Scholar 
    Botkin, D. B. et al. Forecasting the effects of global warming on biodiversity. BioScience 57, 227–236 (2007).Article 

    Google Scholar 
    Wiens, J. A., Stralberg, D., Jongsomjit, D., Howell, C. A. & Snyder, M. A. Niches, models, and climate change: assessing the assumptions and uncertainties. Proc. Natl Acad. Sci. USA 106, 19729–19736 (2009).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Smith, A. B., Godsoe, W., Rodríguez-Sánchez, F., Wang, H. H. & Warren, D. Niche estimation above and below the species level. Trends Ecol. Evol. 34, 260–273 (2019).PubMed 
    Article 

    Google Scholar 
    Waldvogel, A.-M. et al. Evolutionary genomics can improve prediction of species’ responses to climate change. Evol. Lett. 4, 4–18 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Razgour, O. et al. An integrated framework to identify wildlife populations under threat from climate change. Mol. Ecol. Resour. 18, 18–31 (2018).PubMed 
    Article 

    Google Scholar 
    Razgour, O. et al. Considering adaptive genetic variation in climate change vulnerability assessment reduces species range loss projections. Proc. Natl Acad. Sci. USA 116, 10418–10423 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Aguirre-Liguori, J. A., Ramírez-Barahona, S., Tiffin, P. & Eguiarte, L. E. Climate change is predicted to disrupt patterns of local adaptation in wild and cultivated maize. Proc. R. Soc. B 286, 20190486 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Evans, T. G., Diamond, S. E. & Kelly, M. W. Mechanistic species distribution modelling as a link between physiology and conservation. Conserv. Physiol. 3, cov056 (2015).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Hall, S. J. G. Haemoglobin polymorphism in the bank vole, Clethrionomys glareolus, in Britain. J. Zool. 187, 153–160 (1979).Article 

    Google Scholar 
    Kotlík, P. et al. Adaptive phylogeography: functional divergence between haemoglobins derived from different glacial refugia in the bank vole. Proc. R. Soc. B 281, 20140021 (2014).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Searle, J. B. et al. The Celtic fringe of Britain: Insights from small mammal phylogeography. Proc. R. Soc. B 276, 4287–4294 (2009).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Escalante, M. A., Horníková, M., Marková, S. & Kotlík, P. Niche differentiation in a postglacial colonizer, the bank vole Clethrionomys glareolus. Ecol. Evol. 11, 8054–8070 (2021).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Reischl, E., Dafre, A. L., Franco, J. L. & Wilhelm Filho, D. Distribution, adaptation and physiological meaning of thiols from vertebrate hemoglobins. Comp. Biochem. Physiol. Part C. Toxicol. Pharmacol. 146, 22–53 (2007).Article 
    CAS 

    Google Scholar 
    Storz, J. F. & Wheat, C. W. Integrating evolutionary and functional approaches to infer adaptation at specific loci. Evolution 64, 2489–2509 (2010).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Rossi, R. et al. Different metabolizing ability of thiol reactants in human and rat blood. Biochemical and pharmacological implications. J. Biol. Chem. 276, 7004–7010 (2001).CAS 
    PubMed 
    Article 

    Google Scholar 
    Vitturi, D. A. et al. Antioxidant functions for the hemoglobin β93 cysteine residue in erythrocytes and in the vascular compartment in vivo. Free Radic. Biol. Med. 55, 119–129 (2013).CAS 
    PubMed 
    Article 

    Google Scholar 
    Petersen, A. G. et al. Hemoglobin polymerization via disulfide bond formation in the hypoxia-tolerant turtle Trachemys scripta: Implications for antioxidant defense and O2 transport. Am. J. Physiol. Regul. Integr. Comp. Physiol. 314, R84–R93 (2018).PubMed 
    Article 
    CAS 

    Google Scholar 
    Paital, B. et al. Longevity of animals under reactive oxygen species stress and disease susceptibility due to global warming. World J. Biol. Chem. 7, 110–127 (2016).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Jacobs, P. J., Oosthuizen, M. K., Mitchell, C., Blount, J. D. & Bennett, N. C. Heat and dehydration induced oxidative damage and antioxidant defenses following incubator heat stress and a simulated heat wave in wild caught four-striped field mice Rhabdomys dilectus. PLoS One 15, e0242279 (2020).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Kotlík, P., Marková, S., Horníková, M., Escalante, M. A. & Searle, J. B. The bank vole (Clethrionomys glareolus) as a model system for adaptive phylogeography in the European theater. Front. Ecol. Evol. 10, 866605 (2022).Article 

    Google Scholar 
    Strážnická, M., Marková, S., Searle, J. B. & Kotlík, P. Playing hide-and-seek in beta-globin genes: Gene conversion transferring a beneficial mutation between differentially expressed gene guplicates. Genes 9, 492 (2018).PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Stocker, T. Climate Change 2013: the Physical Science Basis: Working Group I Contribution to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change (Cambridge University Press, 2013).Araújo, M. B., Pearson, R. G., Thuiller, W. & Erhard, M. Validation of species-climate impact models under climate change. Glob. Chang. Biol. 11, 1504–1513 (2005).Article 

    Google Scholar 
    Peterson, A. T., Papeş, M. & Soberón, J. Rethinking receiver operating characteristic analysis applications in ecological niche modeling. Ecol. Modell. 213, 63–72 (2008).Article 

    Google Scholar 
    Warren, D. L., Glor, R. E. & Turelli, M. Environmental niche equivalency versus conservatism: quantitative approaches to niche evolution. Evolution 62, 2868–2883 (2008).PubMed 
    Article 

    Google Scholar 
    Warren, D. L. et al. ENMTools 1.0: an R package for comparative ecological biogeography. Ecography 44, 504–511 (2021).Article 

    Google Scholar 
    Mayes, J. & Wheeler, D. Regional weather and climates of the British Isles—part 1: introduction. Weather 68, 3–8 (2013).Article 

    Google Scholar 
    Kotlík, P., Marková, S., Konczal, M., Babik, W. & Searle, J. B. Genomics of end-Pleistocene population replacement in a small mammal. Proc. R. Soc. B 285, 20172624 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Capblancq, T., Fitzpatrick, M. C., Bay, R. A., Exposito-Alonso, M. & Keller, S. R. Genomic prediction of (mal)adaptation across current and future climatic landscapes. Annu. Rev. Ecol. Evol. Syst. 51, 245–269 (2020).Article 

    Google Scholar 
    Benito Garzón, M., Robson, T. M. & Hampe, A. ΔTraitSDMs: species distribution models that account for local adaptation and phenotypic plasticity. N. Phytol. 222, 1757–1765 (2019).Article 

    Google Scholar 
    Wisz, M. S. et al. Effects of sample size on the performance of species distribution models. Divers. Distrib. 14, 763–773 (2008).Article 

    Google Scholar 
    Phillips, S. J., Dudík, M. & Schapire, R. E. A maximum entropy approach to species distribution modeling. in Twenty-first International Conference on Machine Learning – ICML ’04 9, 83 (ACM Press, 2004).Hijmans, R. J., Cameron, S. E., Parra, J. L., Jones, P. G. & Jarvis, A. Very high resolution interpolated climate surfaces for global land areas. Int. J. Climatol. 25, 1965–1978 (2005).Article 

    Google Scholar 
    Zeng, Y., Low, B. W. & Yeo, D. C. J. Novel methods to select environmental variables in MaxEnt: A case study using invasive crayfish. Ecol. Modell. 341, 5–13 (2016).Article 

    Google Scholar 
    Warren, D. L. & Seifert, S. N. Ecological niche modeling in Maxent: the importance of model complexity and the performance of model selection criteria. Ecol. Appl. 21, 335–342 (2011).PubMed 
    Article 

    Google Scholar 
    Warren, D. L., Glor, R. E. & Turelli, M. ENMTools: a toolbox for comparative studies of environmental niche models. Ecography 33, 607–611 (2010).Article 

    Google Scholar 
    Gent, P. R. et al. The community climate system model version 4. J. Clim. 24, 4973–4991 (2011).Article 

    Google Scholar 
    Dufresne, J. L. et al. Climate change projections using the IPSL-CM5 Earth System Model: from CMIP3 to CMIP5. Clim. Dyn. 40, 2123–2165 (2013).Article 

    Google Scholar 
    Watanabe, S. et al. MIROC-ESM 2010: model description and basic results of CMIP5-20c3m experiments. Geosci. Model Dev. 4, 845–872 (2011).Article 

    Google Scholar 
    Giorgetta, M. A. et al. Climate and carbon cycle changes from 1850 to 2100 in MPI-ESM simulations for the Coupled Model Intercomparison Project phase 5. J. Adv. Model. Earth Syst. 5, 572–597 (2013).Article 

    Google Scholar 
    Schoener, T. W. The anolis lizards of Bimini: resource partitioning in a complex fauna. Ecology 49, 704–726 (1968).Article 

    Google Scholar  More

  • in

    Putting pesticides on the map for pollinator research and conservation

    Overall strategyThe aim of this project was to synthesize publicly available data on land use, pesticide use, and toxicity to generate a ‘toolkit’ of data resources enabling improved landscape-scale research on pesticide-pollinator interactions. The main outcomes are several novel datasets covering ten major crops or crop groups in each of the 48 contiguous U.S. states:

    I)

    Average application rate (kg/ha/yr) of >500 common pesticide active ingredients (1997–2017),

    II)

    Aggregate bee toxic load (honey bee lethal doses/ha/yr) of all insecticides combined (1997–2014), (Note that this dataset ends in 2014 because after that year, data on seed-applied pesticides were excluded29, and these contribute significantly to bee toxic load21)

    III)

    Reclass tables relating these pesticide-use indicators to land use/land cover classes to enable the creation of maps predicting annual pesticide loading at 30–56 m resolution.

    An overview of the steps, inputs, and outcomes are provided in Fig. 1.Fig. 1Overview of the data synthesis workflow described in this paper.Full size imageData inputsA summary of input datasets is provided in Table 1.Table 1 Data inputs used in this study.Full size tablePesticide dataPesticide use data were last downloaded from the USGS National Pesticide Synthesis Project30,31 in June 2020. This dataset reports total kg applied of 508 common pesticide active ingredients by combinations of state, crop group, and year for the contiguous U.S. from 1992–2017 (crop groups explained in Table 2). The data are derived primarily from farmer surveys conducted by a private firm (Kynetec). For California, USGS obtains data from the state’s pesticide use reporting program32. USGS then aggregates and standardizes both data sources into a common national dataset that is released to the public and was used in this effort. The USGS dataset includes both a ‘high’ and a ‘low’ estimate of pesticide use, varying based on the treatment of missing values in the source data31. Because previous work on this dataset suggested that the ‘low’ estimate more closely matches independent pesticide estimates33, we used the ‘low’ estimate throughout, but assess the influence of this choice on the resulting estimates (see Technical Validation). While we focus on the ‘low’ estimate for the data and outputs presented in this manuscript, the workflow we developed can accommodate both the low and high estimates.Table 2 USGS crop categories in pesticide source data, based on metadata from USGS30,31 and personal communication with USGS staff scientists.Full size tableCrop area dataTo translate pesticide use estimates into average application rates, it was necessary to divide total kg of pesticide applied by the land area to which it was potentially applied. Crop area data were last downloaded from the Quick Stats Database of the USDA34 in May 2020, using data files downloaded from the ‘developer’ page. This USDA dataset contains crop acreage estimates generated from two sources: the Census of Agriculture (Census), which is comprehensive but conducted only once every five years35 and the crop survey conducted by the National Agricultural Statistics Service (NASS), which is an annual survey based on a representative sample of farmers in major production regions for a more limited subset of crops36.Honey bee toxicity dataTranslating insecticide application rates into estimates of bee toxic load (honey bee lethal doses/ha/yr) required toxicity values for each insecticide active ingredient in the USGS dataset. We used LD50 values for the honey bee (Apis mellifera) because this is the standard terrestrial insect species used in regulatory procedures, and so has the most comprehensive data available. This species is also of particular concern as an important provider of pollination services to agriculture. As previously reported21, the LD50 values were derived from two sources, the ECOTOX database37 of the U.S. Environmental Protection Agency (US-EPA), and the Pesticide Properties Database (PPDB)038. ECOTOX was queried in July 2017, by searching for all LD50 values for the honey bee (Apis mellifera) that were generated under laboratory conditions. Acute contact and oral LD50 values for the honey bee were recorded manually from the PPDB in June 2018.Land cover dataMapping pesticides to the landscape requires land use/land cover data indicating where crops are grown. We used the USDA Cropland Data Layer (CDL)39, a land cover dataset at 30–56 m resolution produced through remote sensing. This dataset is available starting in 2008 for states in the contiguous U.S., with some states (primarily in the Midwest and Mid-South) available back to the early 2000s.Data preparationRelating datasetsA major challenge in this data synthesis effort was relating the various data sources to each other, given that each dataset has unique nomenclature and organization. We created the following keys (summarized in Table 3) to facilitate joining datasets:

    I)

    USGS-USDA crop keys – Using documentation and metadata associated with the USGS pesticide dataset31,33,40, we created keys relating the USGS surveyed crop names (‘ePest’ crops) and the ten USGS crop categories to the large number of corresponding crop acreage data items in the Census and NASS datasets. For annual crops and hay crops we used ‘harvested acres,’ and for tree crops we used ‘acres bearing & non-bearing.’ These choices were made to maximize data availability and to correspond as closely as possible to the crop acreage from which the pesticide data were derived31. A separate key was developed for California because California pesticide data derives from different source data and covers a larger range of crops.

    II)

    USGS-CASRN compound key – Using USGS documentation as well as background information on pesticide active ingredients38,41, we generated keys relating USGS active ingredient names to chemical abstracts service (CAS) registry numbers to facilitate matching compounds to the ECOTOX and PPDB databases.

    III)

    USGS compound-category key – In this key we classified active ingredients into major groups (insecticides, fungicides, nematicides, etc.) and into mode-of-action classes on the basis of information from pesticide databases and resistance action committees38,41,42,43,44.

    IV)

    USGS-USDA compound key – To facilitate our data validation effort, we generated a key relating USGS compound names to USDA compound names, on the basis of information from several pesticide databases38,41.

    V)

    USGS-CDL land use-land cover keys – Using documentation from the USGS pesticide dataset describing the crop composition of each of the ten crop categories31, we created a key that matches these categories to land cover classes in the CDL. A separate key was developed for California given the differences in surveyed crops in this state, noted above.

    Table 3 Keys generated to relate datasets.Full size tableProcessing crop area dataBecause of differences in the crops included in pesticide use estimates, crop acreage data were processed separately for California and for all other states, and then re-joined, as follows: Acreage data were first filtered to include only data at the state level, reporting total annual acreage for states in the contiguous U.S. after 1996. Acreage data were joined to the appropriate USGS-USDA crop key and only those crops represented in the pesticide dataset were retained. We then generated an acreage dataset with single rows for each combination of crop, state, and year using data from the Census when available (1997, 2002, 2007, 2012, 2017), data from NASS in non-Census years, and temporal interpolation to fill in remaining missing values (i.e. linear interpolation between values in the same state and crop in the nearest surrounding years). This process was repeated for California, using acreage data for only that state in combination with the CA crop key. Finally, acreage data in the two datasets were recombined, converted to hectares, and summed by USGS crop group.Processing honey bee toxicity dataProcessing for the honey bee toxicity data has been described in detail elsewhere21. Briefly, toxicity values were categorized as contact, oral, or other and standardized where possible into µg/bee. Records were retained if they represented acute exposure (4 days or less) for adult bees representing contact or oral LD50 values in µg/bee. To generate a consensus list of contact and oral LD50 values for all insecticides reported in the USGS dataset, we gave preference to point estimates and estimates generated through U.S. or E.U. regulatory procedures, taking a geometric mean if multiple such estimates were available. Unbounded estimates (“greater than” or “less than” some value) were only used when point estimates were unavailable, using the minimum (for “less than”) or the maximum (for “greater than”). If values for a compound were unavailable in both datasets, we used the median toxicity value for the insecticide mode-of-action group. And finally, in rare cases (n = 1/148 compounds for contact toxicity and 8/148 compounds for oral toxicity) we were still left without a toxicity estimate for a particular insecticide. In those cases, we used the median value for all insecticides.Data synthesisCompound-specific application rates for state-crop-year combinationsUSGS data on pesticide application were joined to data on crop area. Average pesticide application rates were calculated by dividing kg applied by crop area (ha) for each combination of compound, crop group, state, and year.Aggregate insecticide application rates for state-crop-year combinationsThe dataset from the previous step was filtered to include only insecticides, and then joined to LD50 data by compound name. Bee toxic load associated with each insecticide active ingredient was calculated by dividing the application rate by the contact or oral LD50 value (µg/bee) to generate a number of lethal doses applied per unit area. These values were then summed across compounds to generate estimates of kg and bee toxic load per ha for combinations of crop group, state, and year.Missing values were estimated using temporal interpolation, where possible (i.e. linear interpolation between values in the same state and crop group in the nearest surrounding years). This dataset ends in 2014 because after that year seed-applied pesticides were excluded from the source data29, and they constitute a major contribution to bee toxic load21.We focused bee toxic load on insecticides for three reasons. First, quality of LD50 data is highest for insecticides and uneven for fungicides and herbicides. Point estimates make up the majority of LD50 values for insecticides, whereas  100 µg/bee”, increasing the uncertainty of downstream estimates). Second, insecticides tend to have greater acute toxicity toward insects than fungicides and herbicides (median [IQR] LD50 = 100 [44–129] µg/bee for fungicides, 100 [75–112] µg/bee for herbicides, and 1.36 [0.16–12] µg/bee for insecticides). As a result, insecticides account for > 95% of bee toxic load nationally, even when herbicides and fungicides are included (and even though insecticides make up only 6.5% of pesticides applied on a weight basis). Third, focusing these values on insecticides increases their interpretability, reflecting efforts directed toward insect pest management, rather than a mix of insect, weed, and fungal pest management (which often have distinct dynamics and constraints for farmers).While we chose to include only insecticides in this aggregate value, users are welcome to adjust the workflow to include fungicides and herbicides if desired. To this end, we provide our best estimates for LD50 values for fungicides and herbicides in the USGS dataset (Table 4).Table 4 Data outputs generated by this study.Full size tableReclassification tablesTo generate reclassification tables for the CDL, the pesticide datasets described above were joined by crop group to CDL land use categories. The output of these processes was a set of reclassification tables for combinations of compound, state, and year. Also generated was a set of reclassification tables for aggregate insecticide use for combinations of state and year.Of the 131 land use categories in the CDL, 16 represent two crops grown sequentially in the same year (double crops, found on ~2% of U.S. cropland in 201245), which required a modified accounting in our workflow. Pesticide use practices on double crops are not well described, but one study suggested that pesticide expenditures on soybean grown after wheat were similar to pesticide expenditures in soybean grown alone46. Therefore, we assumed that pesticide use on double crops would be additive (e.g. for a wheat-soybean double crop, the annual pesticide use estimate was generated by summing pesticide use associated with wheat and soybean).Missing values in the reclassification tables resulted from several distinct issues. Some values were missing because a particular crop was not included in the underlying pesticide use survey (e.g. oats was not included in the Kynetec survey), or because the land use category was not a crop at all (e.g. deciduous forest). These two issues were indicated with values of ‘1’ in columns called ‘unsurveyed’ and ‘noncrop,’ respectively. For double crops, a value of 0.5 in the ‘unsurveyed’ column indicates that one of the crops was surveyed and the other was not. For compound-specific datasets, missing values may reflect that a given compound was not used in a state-crop group-year combination. For the aggregate insecticide dataset, even after interpolation there were some missing values, usually when a state had very little area of a particular crop or crop group.Finally, missing data for double crops were treated slightly differently in the aggregate vs. compound-specific reclassification tables. For the aggregate insecticide dataset, estimates for double crops were only included if estimates were available for both crops; otherwise the value was reported as missing. For the compound-specific datasets, estimates for double crops were included if there was an estimate for at least one of the crops, since specific compounds may be used in one crop but not another. More

  • in

    Genomic adaptation of the picoeukaryote Pelagomonas calceolata to iron-poor oceans revealed by a chromosome-scale genome sequence

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

    Google Scholar 
    Boyce, D. G., Lewis, M. R. & Worm, B. Global phytoplankton decline over the past century. Nature 466, 591–596 (2010).CAS 
    PubMed 
    Article 

    Google Scholar 
    Henson, S. A., Cael, B. B., Allen, S. R. & Dutkiewicz, S. Future phytoplankton diversity in a changing climate. Nat. Commun. 12, 5372 (2021).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Vaulot, D., Eikrem, W., Viprey, M. & Moreau, H. The diversity of small eukaryotic phytoplankton (≤3 μm) in marine ecosystems. FEMS Microbiol. Rev. 32, 795–820 (2008).CAS 
    PubMed 
    Article 

    Google Scholar 
    Agawin, N. S. R., Duarte, C. M. & Agustí, S. Nutrient and temperature control of the contribution of picoplankton to phytoplankton biomass and production. Limnol. Oceanogr. 45, 591–600 (2000).CAS 
    Article 

    Google Scholar 
    Morán, X. A. G., López-Urrutia, Á., Calvo-Díaz, A. & Li, W. K. W. Increasing importance of small phytoplankton in a warmer ocean. Glob. Change Biol. 16, 1137–1144 (2010).Article 

    Google Scholar 
    Li, W. K. W., McLaughlin, F. A., Lovejoy, C. & Carmack, E. C. Smallest algae thrive as the arctic ocean freshens. Science 326 https://doi.org/10.1126/science.1179798 (2009).Benner, I., Irwin, A. J. & Finkel, Z. V. Capacity of the common Arctic picoeukaryote Micromonas to adapt to a warming ocean. Limnol. Oceanography Lett. 5, 221–227 (2020).Sunda, W. G. & Huntsman, S. A. Iron uptake and growth limitation in oceanic and coastal phytoplankton. Mar. Chem. 50, 189–206 (1995).CAS 
    Article 

    Google Scholar 
    Raven, J. A. The twelfth Tansley Lecture. Small is beautiful: the picophytoplankton. Funct. Ecol. 12, 503–513 (1998).Article 

    Google Scholar 
    Morel, F. M. M. & Price, N. M. The biogeochemical cycles of trace metals in the oceans. Science 300, 944–947 (2003).CAS 
    PubMed 
    Article 

    Google Scholar 
    Gao, X., Bowler, C. & Kazamia, E. Iron metabolism strategies in diatoms. J. Exp. Bot. 72, 2165–2180 (2021).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Caputi, L. et al. Community-level responses to iron availability in open ocean plankton ecosystems. Glob. Biogeochemical Cycles 33, 391–419 (2019).CAS 
    Article 

    Google Scholar 
    Carradec, Q. et al. A global ocean atlas of eukaryotic genes. Nat. Commun. 9, 373 (2018).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Morrissey, J. et al. A novel protein, ubiquitous in marine phytoplankton, concentrates iron at the cell surface and facilitates uptake. Curr. Biol. 25, 364–371 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    Moore, C. M. et al. Processes and patterns of oceanic nutrient limitation. Nat. Geosci. 6, 701–710 (2013).CAS 
    Article 

    Google Scholar 
    Kumar, A. & Bera, S. Revisiting nitrogen utilization in algae: a review on the process of regulation and assimilation. Bioresour. Technol. Rep. 12, 100584 (2020).Article 

    Google Scholar 
    Smith, S. R. et al. Evolution and regulation of nitrogen flux through compartmentalized metabolic networks in a marine diatom. Nat. Commun. 10, 4552 (2019).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Berg, G. M., Glibert, P. M., Lomas, M. W. & Burford, M. A. Organic nitrogen uptake and growth by the chrysophyte Aureococcus anophagefferens during a brown tide event. Mar. Biol. 129, 377–387 (1997).CAS 
    Article 

    Google Scholar 
    Andersen, R. A., Saunders, G. W., Paskind, M. P. & Sexton, J. P. Ultrastructure and 18s rRNA gene sequence for Pelagomonas calceolata gen. et sp. nov. and the description of a new algal class, the pelagophyceae classis nov. J. Phycol. 29, 701–715 (1993).CAS 
    Article 

    Google Scholar 
    Choi, C. J. et al. Seasonal and geographical transitions in eukaryotic phytoplankton community structure in the Atlantic and Pacific Oceans. Front. Microbiol. 11, 542372 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Duerschlag, J. et al. Niche partitioning by photosynthetic plankton as a driver of CO2-fixation across the oligotrophic South Pacific Subtropical Ocean. ISME J 1–12 https://doi.org/10.1038/s41396-021-01072-z (2021).Worden, A. Z. et al. Global distribution of a wild alga revealed by targeted metagenomics. Curr. Biol. 22, R675–R677 (2012).CAS 
    PubMed 
    Article 

    Google Scholar 
    Dimier, C. é, Brunet, C., Geider, R. & Raven, J. Growth and photoregulation dynamics of the picoeukaryote Pelagomonas calceolata in fluctuating light. Limnol. Oceanogr. 54, 823–836 (2009).CAS 
    Article 

    Google Scholar 
    Dupont, C. L. et al. Genomes and gene expression across light and productivity gradients in eastern subtropical Pacific microbial communities. ISME J. 9, 1076–1092 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    Kang, Y. et al. Transcriptomic responses of four pelagophytes to nutrient (N, P) and light stress. Front. Mar. Sci. 8, 636699 (2021).Huff, J. T., Zilberman, D. & Roy, S. W. Mechanism for DNA transposons to generate introns on genomic scales. Nature 538, 533–536 (2016).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Waterhouse, R. M. et al. BUSCO applications from quality assessments to gene prediction and phylogenomics. Mol. Biol. Evol. 35, 543–548 (2018).CAS 
    PubMed 
    Article 

    Google Scholar 
    Nambiar, M. & Smith, G. R. Repression of harmful meiotic recombination in centromeric regions. Semin Cell Dev. Biol. 54, 188–197 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Pessia, E. et al. Evidence for widespread GC-biased gene conversion in eukaryotes. Genome Biol. Evol. 4, 675–682 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Chi, J., Mahé, F., Loidl, J., Logsdon, J. & Dunthorn, M. Meiosis gene inventory of four ciliates reveals the prevalence of a synaptonemal complex-independent crossover pathway. Mol. Biol. Evol. 31, 660–672 (2014).CAS 
    PubMed 
    Article 

    Google Scholar 
    Ramesh, M. A., Malik, S.-B. & Logsdon, J. M. A phylogenomic inventory of meiotic genes; evidence for sex in Giardia and an early eukaryotic origin of meiosis. Curr. Biol. 15, 185–191 (2005).CAS 
    PubMed 

    Google Scholar 
    Schurko, A. M. & Logsdon, J. M. Using a meiosis detection toolkit to investigate ancient asexual ‘scandals’ and the evolution of sex. Bioessays 30, 579–589 (2008).CAS 
    PubMed 
    Article 

    Google Scholar 
    Ibarbalz, F. M. et al. Global trends in marine plankton diversity across kingdoms of life. Cell 179, 1084–1097.e21 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Frémont, P. et al. Restructuring of plankton genomic biogeography in the surface ocean under climate change. Nat. Clim. Chang. 12, 393–401 (2022).Article 

    Google Scholar 
    Ward, D. M. & Kaplan, J. Ferroportin-mediated iron transport: expression and regulation. Biochim Biophys. Acta 1823, 1426–1433 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Gobler, C. J., Lonsdale, D. J. & Boyer, G. L. A review of the causes, effects, and potential management of harmful brown tide blooms caused by Aureococcus anophagefferens (Hargraves et sieburth). Estuaries 28, 726–749 (2005).Article 

    Google Scholar 
    Agusti, S., Lubián, L. M., Moreno-Ostos, E., Estrada, M. & Duarte, C. M. Projected changes in photosynthetic picoplankton in a warmer subtropical ocean. Front. Mar. Sci. 5, 506 (2019).Article 

    Google Scholar 
    Anderson, S. I., Barton, A. D., Clayton, S., Dutkiewicz, S. & Rynearson, T. A. Marine phytoplankton functional types exhibit diverse responses to thermal change. Nat. Commun. 12, 6413 (2021).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Martin, J. H. et al. Testing the iron hypothesis in ecosystems of the equatorial Pacific Ocean. Nature 371, 123–129 (1994).CAS 
    Article 

    Google Scholar 
    Shi, D., Xu, Y., Hopkinson, B. M. & Morel, F. M. M. Effect of ocean acidification on iron availability to marine phytoplankton. Science 327, 676–679 (2010).CAS 
    PubMed 
    Article 

    Google Scholar 
    McQuaid, J. B. et al. Carbonate-sensitive phytotransferrin controls high-affinity iron uptake in diatoms. Nature 555, 534–537 (2018).CAS 
    PubMed 
    Article 

    Google Scholar 
    Turnšek, J. et al. Proximity proteomics in a marine diatom reveals a putative cell surface-to-chloroplast iron trafficking pathway. eLife 10, e52770 (2021).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Urzica, E. I. et al. Systems and trans-system level analysis identifies conserved iron deficiency responses in the plant lineage[W][OA]. Plant Cell 24, 3921–3948 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Mao, X. et al. Diversity, prevalence, and expression of cyanase genes (cynS) in planktonic marine microorganisms. ISME J. 16, 602–605 (2022).CAS 
    PubMed 
    Article 

    Google Scholar 
    Ou, L., Cai, Y., Jin, W., Wang, Z. & Lu, S. Understanding the nitrogen uptake and assimilation of the Chinese strain of Aureococcus anophagefferens (Pelagophyceae). Algal Res. 34, 182–190 (2018).Article 

    Google Scholar 
    Shu, C. J., Ulrich, L. E. & Zhulin, I. B. The NIT domain: a predicted nitrate-responsive module in bacterial sensory receptors. Trends Biochem Sci. 28, 121–124 (2003).CAS 
    PubMed 
    Article 

    Google Scholar 
    Wu, S. Q., Chai, W., Lin, J. T. & Stewart, V. General nitrogen regulation of nitrate assimilation regulatory gene nasR expression in Klebsiella oxytoca M5al. J. Bacteriol. 181, 7274–7284 (1999).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Durand, N. C. et al. Juicer provides a one-click system for analyzing loop-resolution Hi-C experiments. Cell Syst. 3, 95–98 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Li, R., Li, Y., Kristiansen, K. & Wang, J. SOAP: short oligonucleotide alignment program. Bioinformatics 24, 713–714 (2008).CAS 
    PubMed 
    Article 

    Google Scholar 
    Alberti, A. et al. Viral to metazoan marine plankton nucleotide sequences from the Tara Oceans expedition. Sci. Data 4, 170093 (2017).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Kopylova, E., Noé, L. & Touzet, H. SortMeRNA: fast and accurate filtering of ribosomal RNAs in metatranscriptomic data. Bioinformatics 28, 3211–3217 (2012).CAS 
    PubMed 
    Article 

    Google Scholar 
    Kim, D., Song, L., Breitwieser, F. P. & Salzberg, S. L. Centrifuge: rapid and sensitive classification of metagenomic sequences. Genome Res. https://doi.org/10.1101/gr.210641.116 (2016).Vurture, G. W. et al. GenomeScope: fast reference-free genome profiling from short reads. Bioinformatics 33, 2202–2204 (2017).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Vaser, R. & Šikić, M. Yet another de novo genome assembler. BioRxiv. https://doi.org/10.1101/656306 (2019).Liu, H. et al. SMARTdenovo: a de novo assembler using long noisy reads. Gigabyte 2021, 1–9 (2021).Article 

    Google Scholar 
    Kolmogorov, M., Yuan, J., Lin, Y. & Pevzner, P. A. Assembly of long, error-prone reads using repeat graphs. Nat. Biotechnol. 37, 540–546 (2019).CAS 
    PubMed 
    Article 

    Google Scholar 
    Ruan, J. & Li, H. Fast and accurate long-read assembly with wtdbg2. Nat. Methods 17, 155–158 (2020).CAS 
    PubMed 
    Article 

    Google Scholar 
    Wick, R. R., Schultz, M. B., Zobel, J. & Holt, K. E. Bandage: interactive visualization of de novo genome assemblies. Bioinformatics 31, 3350–3352 (2015).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Vaser, R., Sović, I., Nagarajan, N. & Šikić, M. Fast and accurate de novo genome assembly from long uncorrected reads. Genome Res 27, 737–746 (2017).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Aury, J.-M. & Istace, B. Hapo-G, haplotype-aware polishing of genome assemblies with accurate reads. NAR Genomics Bioinform. 3, lqab034 (2021).Benson, G. Tandem repeats finder: a program to analyze DNA sequences. Nucleic Acids Res. 27, 573–580 (1999).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Morgulis, A., Gertz, E. M., Schäffer, A. A. & Agarwala, R. A fast and symmetric DUST implementation to mask low-complexity DNA sequences. J. Comput Biol. 13, 1028–1040 (2006).CAS 
    PubMed 
    Article 

    Google Scholar 
    Smit, A. F. A., Hubley, R. & Green, P. RepeatMasker. http://repeatmasker.org/ (2013).Price, A. L., Jones, N. C. & Pevzner, P. A. De novo identification of repeat families in large genomes. Bioinformatics 21, i351–i358 (2005).CAS 
    PubMed 
    Article 

    Google Scholar 
    Quinlan, A. R. & Hall, I. M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Pedersen, B. S. & Quinlan, A. R. Mosdepth: quick coverage calculation for genomes and exomes. Bioinformatics 34, 867–868 (2018).CAS 
    PubMed 
    Article 

    Google Scholar 
    Schulz, M. H., Zerbino, D. R., Vingron, M. & Birney, E. Oases: robust de novo RNA-seq assembly across the dynamic range of expression levels. Bioinformatics 28, 1086–1092 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Zerbino, D. R. & Birney, E. Velvet: algorithms for de novo short read assembly using de Bruijn graphs. Genome Res. 18, 821–829 (2008).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Li, H. et al. The sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Marchler-Bauer, A. et al. CDD: NCBI’s conserved domain database. Nucleic Acids Res. 43, D222–D226 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    Niang, G. et al. METdb: A genomic reference database for marine species. F1000Research, https://doi.org/10.7490/f1000research.1118000.1 (2020).Kent, W. J. BLAT–the BLAST-like alignment tool. Genome Res. 12, 656–664 (2002).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Altschul, S. F., Gish, W., Miller, W., Myers, E. W. & Lipman, D. J. Basic local alignment search tool. J. Mol. Biol. 215, 403–410 (1990).CAS 
    PubMed 
    Article 

    Google Scholar 
    Birney, E., Clamp, M. & Durbin, R. GeneWise and genomewise. Genome Res. 14, 988–995 (2004).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Stanke, M. et al. AUGUSTUS: ab initio prediction of alternative transcripts. Nucleic Acids Res. 34, W435–W439 (2006).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Dubarry, M. et al. Gmove a tool for eukaryotic gene predictions using various evidences. F1000Research, https://doi.org/10.7490/f1000research.1111735.1 (2016).Sibbald, S. J., Lawton, M. & Archibald, J. M. Mitochondrial genome evolution in pelagophyte algae. Genome Biol. Evol. 13, evab018 (2021).Quevillon, E. et al. InterProScan: protein domains identifier. Nucleic Acids Res. 33, W116–W120 (2005).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Buchfink, B., Reuter, K. & Drost, H.-G. Sensitive protein alignments at tree-of-life scale using DIAMOND. Nat. Methods 18, 366–368 (2021).CAS 
    PubMed 
    PubMed Central 
    Article 

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

    Google Scholar 
    Delmont, T. O. et al. Functional repertoire convergence of distantly related eukaryotic plankton lineages abundant in the sunlit ocean. Cell Genomics 2, 100123 (2022).CAS 
    Article 

    Google Scholar 
    Li, H. & Durbin, R. Fast and accurate short read alignment with Burrows–Wheeler transform. Bioinformatics 25, 1754–1760 (2009).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Pesant, S. et al. Open science resources for the discovery and analysis of Tara Oceans data. Sci. Data 2, 150023 (2015).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Aumont, O., Ethé, C., Tagliabue, A., Bopp, L. & Gehlen, M. PISCES-v2: an ocean biogeochemical model for carbon and ecosystem studies. Geoscientific Model Dev. 8, 2465–2513 (2015).CAS 
    Article 

    Google Scholar 
    Clayton, S. et al. Biogeochemical versus ecological consequences of modeled ocean physics. Biogeosciences 14, 2877–2889 (2017).CAS 
    Article 

    Google Scholar 
    Ravindra, K., Rattan, P., Mor, S. & Aggarwal, A. N. Generalized additive models: building evidence of air pollution, climate change and human health. Environ. Int. 132, 104987 (2019).CAS 
    PubMed 
    Article 

    Google Scholar 
    Günther, F. & Fritsch, S. neuralnet: training of neural networks. R. J. 2, 30–38 (2010).Article 

    Google Scholar 
    Gobler, C. J. et al. Niche of harmful alga Aureococcus anophagefferens revealed through ecogenomics. Proc. Natl Acad. Sci. USA 108, 4352–4357 (2011).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Guo, L. et al. Genome assembly of Nannochloropsis oceanica provides evidence of host nucleus overthrow by the symbiont nucleus during speciation. Commun. Biol. 2, 1–12 (2019).CAS 
    Article 

    Google Scholar 
    Bowler, C. et al. The Phaeodactylum genome reveals the evolutionary history of diatom genomes. Nature 456, 239–244 (2008).CAS 
    PubMed 
    Article 

    Google Scholar 
    Armbrust, E. V. et al. The genome of the diatom thalassiosira pseudonana: ecology, evolution, and metabolism. Science 306, 79–86 (2004).CAS 
    PubMed 
    Article 

    Google Scholar 
    Worden, A. Z. et al. Green evolution and dynamic adaptations revealed by genomes of the marine picoeukaryotes micromonas. Science 324, 268–272 (2009).CAS 
    PubMed 
    Article 

    Google Scholar 
    Palenik, B. et al. The tiny eukaryote Ostreococcus provides genomic insights into the paradox of plankton speciation. PNAS 104, 7705–7710 (2007).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Moreau, H. et al. Gene functionalities and genome structure in Bathycoccus prasinos reflect cellular specializations at the base of the green lineage. Genome Biol. 13, R74 (2012).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Read, B. A. et al. Pan genome of the phytoplankton Emiliania underpins its global distribution. Nature 499, 209–213 (2013).CAS 
    PubMed 
    Article 

    Google Scholar  More