More stories

  • in

    Summer warming explains widespread but not uniform greening in the Arctic tundra biome

    1.
    Arctic Monitoring and Assessment Programme. Snow, Water, Ice and Permafrost in the Arctic (SWIPA) 2017 (Arctic Monitoring and Assessment Programme (AMAP), 2017).
    2.
    Chapin, F. S. 3rd et al. Role of land-surface changes in arctic summer warming. Science 310, 657–660 (2005).
    ADS  CAS  PubMed  Google Scholar 

    3.
    Tape, K. D., Christie, K., Carroll, G. & O’donnell, J. A. Novel wildlife in the Arctic: the influence of changing riparian ecosystems and shrub habitat expansion on snowshoe hares. Glob. Change Biol. 22, 208–219 (2016).
    ADS  Google Scholar 

    4.
    Downing, A. & Cuerrier, A. A synthesis of the impacts of climate change on the First Nations and Inuit of Canada. Indian J. Tradit. Knowl. 10, 57–70 (2011).
    Google Scholar 

    5.
    National Academies of Sciences. Understanding Northern Latitude Vegetation Greening and Browning: Proceedings of a Workshop (The National Academies Press, 2019).

    6.
    Bjorkman, A. D. et al. Plant functional trait change across a warming tundra biome. Nature 562, 57–62 (2018).
    ADS  CAS  PubMed  Google Scholar 

    7.
    Elmendorf, S. C. et al. Plot-scale evidence of tundra vegetation change and links to recent summer warming. Nat. Clim. Change 2, 453–457 (2012).
    ADS  Google Scholar 

    8.
    Gauthier, G. et al. Long-term monitoring at multiple trophic levels suggests heterogeneity in responses to climate change in the Canadian Arctic tundra. Philos. Trans. R. Soc. Ser. B 368, 20120482 (2013).
    Google Scholar 

    9.
    Myers-Smith, I. H. et al. Eighteen years of ecological monitoring reveals multiple lines of evidence for tundra vegetation change. Ecol. Monogr. 89, e01351 (2019).
    Google Scholar 

    10.
    Tape, K. D., Hallinger, M., Welker, J. M. & Ruess, R. W. Landscape heterogeneity of shrub expansion in Arctic Alaska. Ecosystems 15, 711–724 (2012).
    CAS  Google Scholar 

    11.
    Pattison, R. R., Jorgenson, J. C., Raynolds, M. K. & Welker, J. M. Trends in NDVI and Tundra Community Composition in the Arctic of NE Alaska Between 1984 and 2009. Ecosystems 18, 707–719 (2015).
    Google Scholar 

    12.
    Gamm, C. M. et al. Declining growth of deciduous shrubs in the warming climate of continental western Greenland. J. Ecol. 106, 640–654 (2018).
    CAS  Google Scholar 

    13.
    Forchhammer M. Sea-ice induced growth decline in Arctic shrubs. Biol. Lett. 13, 20170122 (2017).

    14.
    Street, L., Shaver, G., Williams, M. & Van Wijk, M. What is the relationship between changes in canopy leaf area and changes in photosynthetic CO2 flux in arctic ecosystems? J. Ecol. 95, 139–150 (2007).
    Google Scholar 

    15.
    Raynolds, M. K., Walker, D. A., Epstein, H. E., Pinzon, J. E. & Tucker, C. J. A new estimate of tundra-biome phytomass from trans-Arctic field data and AVHRR NDVI. Remote Sens. Lett. 3, 403–411 (2012).
    Google Scholar 

    16.
    Berner, L. T., Jantz, P., Tape, K. D. & Goetz, S. J. Tundra plant aboveground biomass and shrub dominance mapped across the North Slope of Alaska. Environ. Res. Lett. 13, 035002 (2018).
    ADS  Google Scholar 

    17.
    Bhatt, U. S. et al. Changing seasonality of panarctic tundra vegetation in relationship to climatic variables. Environ. Res. Lett. 12, 1–18 (2017).
    Google Scholar 

    18.
    Guay, K. C. et al. Vegetation productivity patterns at high northern latitudes: a multi-sensor satellite data assessment. Glob. Change Biol. 20, 3147–3158 (2014).
    ADS  Google Scholar 

    19.
    Pinzon, J. & Tucker, C. A non-stationary 1981–2012 AVHRR NDVI3g time series. Remote Sens. 6, 6929–6960 (2014).
    ADS  Google Scholar 

    20.
    Ju, J. & Masek, J. G. The vegetation greenness trend in Canada and US Alaska from 1984–2012 Landsat data. Remote Sens. Environ. 176, 1–16 (2016).
    ADS  Google Scholar 

    21.
    Karlsen, S. R., Anderson, H. B., Van der Wal, R. & Hansen, B. B. A new NDVI measure that overcomes data sparsity in cloud-covered regions predicts annual variation in ground-based estimates of high arctic plant productivity. Environ. Res. Lett. 13, 025011 (2018).
    ADS  Google Scholar 

    22.
    McManus, kM. et al. Satellite-based evidence for shrub and graminoid tundra expansion in northern Quebec from 1986 to 2010. Glob. Change Biol. 18, 2313–2323 (2012).
    ADS  Google Scholar 

    23.
    Frost, G. V., Epstein, H. & Walker, D. Regional and landscape-scale variability of Landsat-observed vegetation dynamics in northwest Siberian tundra. Environ. Res. Lett. 9, 025004 (2014).
    ADS  Google Scholar 

    24.
    Raynolds, M. K. & Walker, D. A. Increased wetness confounds Landsat-derived NDVI trends in the central Alaska North Slope region, 1985–2011. Environ. Res. Lett. 11, 085004 (2016).
    ADS  Google Scholar 

    25.
    Gorelick, N. et al. Google Earth Engine: planetary-scale geospatial analysis for everyone. Remote Sens. Environ. 202, 18–27 (2017).
    ADS  Google Scholar 

    26.
    Zhu, Z., Wang, S. & Woodcock, C. E. Improvement and expansion of the Fmask algorithm: cloud, cloud shadow, and snow detection for Landsats 4–7, 8, and Sentinel 2 images. Remote Sens. Environ. 159, 269–277 (2015).
    ADS  Google Scholar 

    27.
    Pastick, N. J. et al. Spatiotemporal remote sensing of ecosystem change and causation across Alaska. Glob. Change Biol. 25, 1171–1189 (2019).
    ADS  Google Scholar 

    28.
    Walker, D. et al. Phytomass, LAI, and NDVI in northern Alaska: relationships to summer warmth, soil pH, plant functional types, and extrapolation to the circumpolar Arctic. J. Geophys. Res. 108, 8169 (2003).
    Google Scholar 

    29.
    Lucht, W. et al. Climatic control of the high-latitude vegetation greening trend and Pinatubo effect. Science 296, 1687–1689 (2002).
    ADS  CAS  PubMed  Google Scholar 

    30.
    Fraser, R. H., Lantz, T. C., Olthof, I., Kokelj, S. V. & Sims, R. A. Warming-induced shrub expansion and lichen decline in the Western Canadian. Arct. Ecosyst. 17, 1151–1168 (2014).
    Google Scholar 

    31.
    Bonney, M. T., Danby, R. K. & Treitz, P. M. Landscape variability of vegetation change across the forest to tundra transition of central Canada. Remote Sens. Environ. 217, 18–29 (2018).
    ADS  Google Scholar 

    32.
    Cuerrier, A., Brunet, N. D., Gérin-Lajoie, J., Downing, A. & Lévesque, E. The study of Inuit knowledge of climate change in Nunavik, Quebec: a mixed methods approach. Hum. Ecol. 43, 379–394 (2015).
    Google Scholar 

    33.
    Forbes, B. C. & Stammler, F. Arctic climate change discourse: the contrasting politics of research agendas in the West and Russia. Polar Res. 28, 28–42 (2009).
    Google Scholar 

    34.
    Forbes, B. C., Fauria, M. M. & Zetterberg, P. Russian Arctic warming and ‘greening’ are closely tracked by tundra shrub willows. Glob. Change Biol. 16, 1542–1554 (2010).
    ADS  Google Scholar 

    35.
    Tape, K., Sturm, M. & Racine, C. The evidence for shrub expansion in Northern Alaska and the Pan-Arctic. Glob. Change Biol. 12, 686–702 (2006).
    ADS  Google Scholar 

    36.
    Ropars, P. & Boudreau, S. Shrub expansion at the forest–tundra ecotone: spatial heterogeneity linked to local topography. Environ. Res. Lett. 7, 015501 (2012).
    ADS  Google Scholar 

    37.
    Myers-Smith, I. H. et al. Complexity revealed in the greening of the Arctic. Nat. Clim. Change 10, 106–117 (2020).
    ADS  Google Scholar 

    38.
    Park, T. et al. Changes in growing season duration and productivity of northern vegetation inferred from long-term remote sensing data. Environ. Res. Lett. 11, 084001 (2016).
    ADS  Google Scholar 

    39.
    Riihimäki, H., Heiskanen, J. & Luoto, M. The effect of topography on arctic-alpine aboveground biomass and NDVI patterns. Int. J. Appl. Earth Obs. Geoinf. 56, 44–53 (2017).
    ADS  Google Scholar 

    40.
    Fraser, R. H., Olthof, I., Lantz, T. C. & Schmitt, C. UAV photogrammetry for mapping vegetation in the low-Arctic. Arct. Sci. 2, 79–102 (2016).
    Google Scholar 

    41.
    Berner, L. T., Beck, P. S. A., Bunn, A. G. & Goetz, S. J. Plant response to climate change along the forest-tundra ecotone in northeastern Siberia. Glob. Change Biol. 19, 3449–3462 (2013).
    Google Scholar 

    42.
    Myers-Smith, I. H. et al. Climate sensitivity of shrub growth across the tundra biome. Nat. Clim. Change 5, 887–891 (2015).
    ADS  Google Scholar 

    43.
    Bjorkman, A. D., Vellend, M., Frei, E. R. & Henry, G. H. Climate adaptation is not enough: warming does not facilitate success of southern tundra plant populations in the high Arctic. Glob. Change Biol. 23, 1540–1551 (2017).
    ADS  Google Scholar 

    44.
    Post, E. & Pedersen, C. Opposing plant community responses to warming with and without herbivores. Proc. Natl Acad. Sci. USA 105, 12353–12358 (2008).
    ADS  CAS  PubMed  Google Scholar 

    45.
    Yu, Q., Epstein, H., Engstrom, R. & Walker, D. Circumpolar arctic tundra biomass and productivity dynamics in response to projected climate change and herbivory. Glob. Change Biol. 23, 3895–3907 (2017).
    ADS  Google Scholar 

    46.
    Liljedahl, A. K. et al. Pan-Arctic ice-wedge degradation in warming permafrost and its influence on tundra hydrology. Nat. Geosci. 9, 312–318 (2016).
    ADS  CAS  Google Scholar 

    47.
    Perreault, N., Levesque, E., Fortier, D. & Lamarque, L. J. Thermo-erosion gullies boost the transition from wet to mesic tundra vegetation. Biogeosciences 13, 1237–1253 (2016).
    ADS  Google Scholar 

    48.
    Grant, R. F., Mekonnen, Z. A., Riley, W. J., Arora, B. & Torn, M. S. Mathematical modelling of Arctic Polygonal Tundra with Ecosys: 2. Microtopography determines how CO2 and CH4 exchange responds to changes in temperature and precipitation. J. Geophys. Res. 122, 3174–3187 (2017).
    CAS  Google Scholar 

    49.
    Phoenix, G. K. & Bjerke, J. W. Arctic browning: extreme events and trends reversing arctic greening. Glob. Change Biol. 22, 2960–2962 (2016).
    ADS  Google Scholar 

    50.
    Treharne, R., Bjerke, J. W., Tømmervik, H., Stendardi, L. & Phoenix, G. K. Arctic browning: Impacts of extreme climatic events on heathland ecosystem CO2 fluxes. Glob. Change Biol. 25, 489–503 (2018).
    ADS  Google Scholar 

    51.
    Forbes, B. C. et al. High resilience in the Yamal-Nenets social–ecological system, west Siberian Arctic, Russia. Proc. Natl Acad. Sci. USA 106, 22041–22048 (2009).
    ADS  CAS  PubMed  Google Scholar 

    52.
    Mekonnen, Z. A., Riley, W. J. & Grant, R. F. Accelerated nutrient cycling and increased light competition will lead to 21st century shrub expansion in North American Arctic tundra. J. Geophys. Res. 123, 1683–1701 (2018).
    CAS  Google Scholar 

    53.
    Rocha, A. V. et al. The footprint of Alaskan tundra fires during the past half-century: implications for surface properties and radiative forcing. Environ. Res. Lett. 7, 044039 (2012).
    ADS  Google Scholar 

    54.
    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).
    ADS  PubMed  PubMed Central  Google Scholar 

    55.
    Hu, F. S. et al. Arctic tundra fires: natural variability and responses to climate change. Front. Ecol. Environ. 13, 369–377 (2015).
    Google Scholar 

    56.
    Mack, M. C. et al. Carbon loss from an unprecedented Arctic tundra wildfire. Nature 475, 489–492 (2011).
    ADS  CAS  PubMed  Google Scholar 

    57.
    Jones, B. M. et al. Identification of unrecognized tundra fire events on the north slope of Alaska. J. Geophys. Res. 118, 1334–1344 (2013).
    Google Scholar 

    58.
    Loranty, M. M. et al. Siberian tundra ecosystem vegetation and carbon stocks four decades after wildfire. J. Geophys. Res. 119, 2144–2154 (2014).
    CAS  Google Scholar 

    59.
    Natali, S. M. et al. Large loss of CO2 in winter observed across the northern permafrost region. Nat. Clim. Change 9, 852–857 (2019).
    ADS  CAS  Google Scholar 

    60.
    Schuur, E. et al. Climate change and the permafrost carbon feedback. Nature 520, 171–179 (2015).
    ADS  CAS  PubMed  Google Scholar 

    61.
    Pearson, R. G. et al. Shifts in Arctic vegetation and associated feedbacks under climate change. Nat. Clim. Change 3, 673–677 (2013).
    ADS  Google Scholar 

    62.
    Loranty, M. M., Goetz, S. J. & Beck, P. S. A. Tundra vegetation effects on pan-Arctic albedo. Environ. Res. Lett. 6, 024014 (2011).
    ADS  Google Scholar 

    63.
    Loranty, M. M. et al. Reviews and syntheses: changing ecosystem influences on soil thermal regimes in northern high-latitude permafrost regions. Biogeosciences 15, 5287–5313 (2018).
    ADS  CAS  Google Scholar 

    64.
    Tape, K. D., Gustine, D. D., Ruess, R. W., Adams, L. G. & Clark, J. A. Range expansion of moose in Arctic Alaska linked to warming and increased shrub habitat. PLoS ONE 11, e0152636 (2016).
    PubMed  PubMed Central  Google Scholar 

    65.
    Tape, K. D., Jones, B. M., Arp, C. D., Nitze, I. & Grosse, G. Tundra be dammed: beaver colonization of the Arctic. Glob. Change Biol. 24, 4478–4488 (2018).
    ADS  Google Scholar 

    66.
    Joly, K., Jandt, R. R. & Klein, D. R. Decrease of lichens in Arctic ecosystems: the role of wildfire, caribou, reindeer, competition and climate in north‐western Alaska. Polar Res. 28, 433–442 (2009).
    Google Scholar 

    67.
    Macias-Fauria, M., Forbes, B. C., Zetterberg, P. & Kumpula, T. Eurasian Arctic greening reveals teleconnections and the potential for structurally novel ecosystems. Nat. Clim. Change 2, 613–618 (2012).
    ADS  Google Scholar 

    68.
    Wesche, S. D. & Chan, H. M. Adapting to the impacts of climate change on food security among Inuit in the Western Canadian Arctic. EcoHealth 7, 361–373 (2010).
    PubMed  Google Scholar 

    69.
    Kuhnlein, H. V. & Chan, H. M. Environment and contaminants in traditional food systems of northern indigenous peoples. Annu. Rev. Nutr. 20, 595–626 (2000).
    CAS  PubMed  Google Scholar 

    70.
    Tucker, C. J. Red and photographic infrared linear combinations for monitoring vegetation. Remote Sens. Environ. 8, 127–150 (1979).
    ADS  Google Scholar 

    71.
    Virtanen, R. et al. Where do the treeless tundra areas of northern highlands fit in the global biome system: toward an ecologically natural subdivision of the tundra biome. Ecol. Evol. 6, 143–158 (2016).
    PubMed  Google Scholar 

    72.
    Masek, J. G. et al. A Landsat surface reflectance dataset for North America, 1990-2000. IEEE Geosci. Remote Sens. Lett. 3, 68–72 (2006).
    ADS  Google Scholar 

    73.
    Vermote, E., Justice, C., Claverie, M. & Franch, B. Preliminary analysis of the performance of the Landsat 8/OLI land surface reflectance product. Remote Sens. Environ. 185, 46–56 (2016).
    ADS  PubMed  Google Scholar 

    74.
    Python Software Foundation. Python Language Software Version 3.7.3. https://www.python.org/ (2020).

    75.
    Foga, S. et al. Cloud detection algorithm comparison and validation for operational Landsat data products. Remote Sens. Environ. 194, 379–390 (2017).
    ADS  Google Scholar 

    76.
    Roy, D. P. et al. Characterization of Landsat-7 to Landsat-8 reflective wavelength and normalized difference vegetation index continuity. Remote Sens. Environ. 185, 57–70 (2016).
    ADS  PubMed  Google Scholar 

    77.
    Sulla-Menashe, D., Friedl, M. A. & Woodcock, C. E. Sources of bias and variability in long-term Landsat time series over Canadian boreal forests. Remote Sens. Environ. 177, 206–219 (2016).
    ADS  Google Scholar 

    78.
    Liaw, A. & Wiener, M. Classification and Regression by randomForest. R News 2, 18–22 (2002).
    Google Scholar 

    79.
    Wright, M. N. & Ziegler, A. Ranger: a fast implementation of random forests for high dimensional data in C++ and R. J. Stat. Softw. 77, 1–17 (2017).
    Google Scholar 

    80.
    Melaas, E. K. et al. Multisite analysis of land surface phenology in North American temperate and boreal deciduous forests from Landsat. Remote Sens. Environ. 186, 452–464 (2016).
    ADS  Google Scholar 

    81.
    Markham, B. L. & Helder, D. L. Forty-year calibrated record of earth-reflected radiance from Landsat: a review. Remote Sens. Environ. 122, 30–40 (2012).
    ADS  Google Scholar 

    82.
    Markham, B. et al. Landsat-8 operational land imager radiometric calibration and stability. Remote Sens. 6, 12275–12308 (2014).
    ADS  Google Scholar 

    83.
    Kendall, M. G. Rank Correlation Methods 4th edn (Charles Griffin, 1975).

    84.
    Sen, P. K. Estimates of the regression coefficient based on Kendall’s tau. J. Am. Stat. Assoc. 63, 1379–1389 (1968).
    MathSciNet  MATH  Google Scholar 

    85.
    Bronaugh, D. & Werner, A. zyp: Zhang + Yue-Pilon Trends Package. R Package Version 0.10-1.1. https://CRAN.R-project.org/package=zyp (2012).

    86.
    R Core Team. R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, 2020).

    87.
    Rohde, R. et al. A new estimate of the average Earth surface land temperature spanning 1753 to 2011. Geoinform. Geostat. 7, https://doi.org/10.4172/2327-4581.1000101 (2013).

    88.
    Hansen, J., Ruedy, R., Sato, M. & Lo, K. Global surface temperature change. Rev. Geophys. 48, RG4004 (2010).
    ADS  Google Scholar 

    89.
    Cowtan, K. & Way, R. G. Coverage bias in the HadCRUT4 temperature series and its impact on recent temperature trends. Q. J. R. Meteorol. Soc. 140, 1935–1944 (2014).
    ADS  Google Scholar 

    90.
    Harris, I., Jones, P. D., Osborn, T. J. & Lister, D. H. Updated high-resolution grids of monthly climatic observations—the CRU TS3.10 Dataset. Int. J. Climatol. 34, 623–642 (2014).
    Google Scholar 

    91.
    Willmott, C. J. & Matsuura, K. Terrestrial Air Temperature and Precipitation: Monthly Time Series (1900–2017) v. 5.01. http://climate.geog.udel.edu/~climate (University of Deleware, 2018).

    92.
    Breiman, L. Random Forests. Mach. Learn. 45, 5–32 (2001).
    MATH  Google Scholar 

    93.
    Abatzoglou, J. T., Dobrowski, S. Z., Parks, S. A. & Hegewisch, K. C. TerraClimate, a high-resolution global dataset of monthly climate and climatic water balance from 1958–2015. Sci. Data 5, 170191 (2018).
    PubMed  PubMed Central  Google Scholar 

    94.
    Obu, J. et al. ESA Permafrost Climate Change Initiative (Permafrost_cci): Permafrost Extent for the Northern Hemisphere, v1.0. https://doi.org/10.5285/c7590fe40d8e44169d511c70a60ccbcc (Centre for Environmental Data Analysis, 2019).

    95.
    Obu, J. et al. ESA Permafrost Climate Change Initiative (Permafrost_cci): Permafrost Ground Temperature for the Northern Hemisphere, v1.0. https://doi.org/10.5285/c7590fe40d8e44169d511c70a60ccbcc (Centre for Environmental Data Analysis, 2019).

    96.
    Obu, J. et al. ESA Permafrost Climate Change Initiative (Permafrost_cci): Permafrost Active Layer Thickness for the Northern Hemisphere, v1.0. https://doi.org/10.5285/1ee56c42cf6c4ef698693e00a63795f4 (Centre for Environmental Data Analysis, 2019).

    97.
    Olefeldt, D. et al. Arctic Circumpolar Distribution and Soil Carbon of Thermokarst Landscapes. https://doi.org/10.3334/ORNLDAAC/1332 (ORNL DAAC, 2015).

    98.
    Defourny, P. et al. Land Cover Climate Change Initiative—Product User Guide Version v2. http://maps.elie.ucl.ac.be/CCI/viewer/download/ESACCI-LC-Ph2-PUGv2_2.0.pdf (European Space Agency, 2017).

    99.
    Rizzoli, P. et al. Generation and performance assessment of the global TanDEM-X digital elevation model. ISPRS J. Photogramm. Remote Sens. 132, 119–139 (2017).
    ADS  Google Scholar 

    100.
    Kuhn, M. Building predictive models in R using the caret package. J. Stat. Softw. 28, 1–26 (2008).
    Google Scholar 

    101.
    Greenwell, B. M. pdp: an R package for constructing partial dependence plots. R. J. 9, 421–436 (2017).
    Google Scholar 

    102.
    Le Moullec, M., Buchwal, A., Wal, R., Sandal, L. & Hansen, B. B. Annual ring growth of a widespread high arctic shrub reflects past fluctuations in community-level plant biomass. J. Ecol. 107, 436–451 (2019).
    Google Scholar 

    103.
    Bunn, A. G. A dendrochronology program library in R (dplR). Dendrochronologia 26, 115–124 (2008).
    Google Scholar 

    104.
    Euskirchen, E., Bret-Harte, M. S., Scott, G., Edgar, C. & Shaver G. R. Seasonal patterns of carbon dioxide and water fluxes in three representative tundra ecosystems in northern Alaska. Ecosphere 3, https://doi.org/10.1890/ES1811-00202.00201 (2012).

    105.
    Euskirchen, E. S. et al. Interannual and seasonal patterns of carbon dioxide, water, and energy fluxes from ecotonal and thermokarst-impacted ecosystems on carbon-rich permafrost soils in Northeastern Siberia. J. Geophys. Res. 122, 2651–2668 (2017).
    CAS  Google Scholar 

    106.
    Baldocchi, D. et al. FLUXNET: a new tool to study the temporal and spatial variability of ecosystem-scale carbon dioxide, water vapor, and energy flux densities. Bull. Am. Meteorol. Soc. 82, 2415–2434 (2001).
    ADS  Google Scholar 

    107.
    Reichstein, M. et al. On the separation of net ecosystem exchange into assimilation and ecosystem respiration: review and improved algorithm. Glob. Change Biol. 11, 1424–1439 (2005).
    ADS  Google Scholar 

    108.
    Hijmans, R. J. raster: Geographic Analysis and Modeling. R package version 3.0-12. http://CRAN.R-project.org/package=raster (2019).

    109.
    Bivand, R., Keitt, T. & Rowlingson B. rgdal: Bindings for the ‘Geospatial’ Data Abstraction Library. R Package Version 1.4-8. https://CRAN.R-project.org/package=rgdal (2019).

    110.
    Bivand, R. & Lewin-Koh, N. maptools: Tools for Handling Spatial Objects. R Package Version 0.9.9. https://CRAN.R-project.org/package=maptools (2019).

    111.
    Dawle, M. & Srinivasan, A. data.table: Extension of ‘data.frame’. R Package Version 1.12.8. https://CRAN.R-project.org/package=data.table (2019).

    112.
    Wickham, H. & Francois, R. dplyr: A Grammar of Data Manipulation. R Package Version 0.8.5. https://CRAN.R-project.org/package=dplyr (2015).

    113.
    Wickham, H. & Henry, L. tidyr: Tidy Messy Data. R Package Version 1.0.2. https://CRAN.R-project.org/package=tidyr (2020).

    114.
    Sarkar, D. Lattice: Multivariate Data Visualization with R (Springer, 2008).

    115.
    Wickham, H. ggplot2: Elegant Graphics for Data Analysis (Springer, New York, 2016).

    116.
    Kassambara, A. ggpubr: ‘ggplot2’ Basde Publication Ready Plots. R Package Version 0.2.5. https://CRAN.R-project.org/package=ggpubr (2020). More

  • in

    Half of resources in threatened species conservation plans are allocated to research and monitoring

    Threatened species assessments
    We assessed the proportion of the proposed budget allocated to RM for a total of 2328 species, independently managed subspecies, or distinct populations (hereafter species): 700 in NZ, 361 in NSW, and 1267 freshwater and terrestrial species in the U.S. In all jurisdictions this included the most threatened listed species and/or those with recovery plans: species with Threatened and Endangered status in the U.S. with active recovery plans as of January 2017, species that met a series of criteria in NSW as of 2013 (e.g., excluding less threatened species that do not require any active intervention and those with a large geographic range17), and the most threatened species in New Zealand as of 2012, which included all species in the Threatened and At-Risk categories with declining populations42. In all three jurisdictions, species are listed for legal protection if they are at risk of extinction. Once listed, recovery planning (including proposed projects, management tasks, and budgets) documents are developed with the objective of securing species from extinction and recovering populations to a point that they can be de-listed. Although our dataset examining threatened species recovery planning is the most comprehensive to date, our data do not represent all spending on species—there are other activities for both management action and RM that occur at a sub-jurisdictional level or outside of government.
    Estimating resources allocated to RM vs action
    We gathered information on the planned costs of management tasks necessary to achieve recovery for threatened species from previously published recovery planning databases (details provided in refs. 15,16,43,44 and Supplementary Methods). Briefly, for NZ and NSW, a suite of management tasks had been developed during structured expert elicitation workshops, as part of a systematic prioritization exercise16,17. For the U.S., management tasks and their cost had been extracted from each species’ published recovery plans (Supplementary Methods15). These data represent an evolution of the implementation of a systematic and cost-effective approach to endangered species resource allocation (i.e., the Project Prioritization Protocol), beginning with NZ in 200916, and subsequently applied to NSW in 201317 and the U.S. in 201615.
    For each proposed management task we used the methods description to categorize tasks as research and monitoring or action based on the definitions in IUCN classification schemes (https://www.iucnredlist.org/resources/classification-schemes, Supplementary Table 145). For NZ and NSW, using previously published datasets we used a combination of the methods description field and 4 other columns that classified the management task methods into increasingly general categories16,17,43. We used keywords such as survey, monitor, surveillance, develop techniques, inventory, research, and develop plan to search for research and monitoring tasks. We reviewed the management tasks identified by these broad search terms to ensure only research and monitoring tasks were included. We also reviewed the management tasks that were not captured by search terms to ensure no research and monitoring tasks were excluded. For the US, the methods descriptions were too complex for keyword searches. Instead, the first author and a trained technician classified each management task manually. To ensure that management tasks were being classified similarly, the first 200 tasks were classified by both observers and any uncertainty was flagged for review together.
    For all jurisdictions, any methods descriptions that were vague, lacked context, or required further assumptions were excluded (2.6% of management tasks, U.S. only). Some management tasks (3.9%) were scored as both action and RM (e.g., translocate birds, action, and monitor the success of the release, RM; weed surveillance, RM, and control, action). For some management tasks, the distinction between action and RM was unclear. These tasks were discussed among the authors and the technician to reach a consensus. For example, ‘standard surveillance to detect invasive mammals’ in NZ could be considered an action, since it is required to detect and subsequently control invasions. However, we assigned it as RM because other management tasks clearly include an action component (e.g., ‘surveillance for invasive species and control if detected’) and other authors have categorized invasive species surveillance as monitoring46. Generally, management tasks to develop conservation plans are distinct from implementing plans and were thus scored as RM (K. Martin pers. comm.). Where we were unable to distinguish between RM and action, we scored as both action and RM.
    For a subset of 8050 management tasks (the first 207 species) in U.S. recovery plans, we further categorized the type of RM to explore common RM activities (Supplementary Table 1). Because we found that assigning management tasks into these 17 categories was challenging without making subjective judgement calls, we did not analyze specific tasks further.
    We estimated the cost of implementing each management task for each species following similar methods to those previously published, calculating costs over 50 years15, Supplementary Methods16,17. We calculated the proportion of the proposed budget allocated to RM for each species as the total cost of all management tasks scored as research or monitoring divided by the total cost of all management tasks. For management tasks that were scored as both action and RM, we multiplied the cost of the task by the average proportional difference between action and RM for each jurisdiction.
    Factors affecting the proportion allocated to RM
    We compared the characteristics of each species recovery plan with the proportion of proposed spending designated as RM. Characteristics available in recovery planning databases for all three jurisdictions included taxon, the estimated benefit of implementing all management tasks, and the total budget estimated for each species (Table 1, Supplementary Methods). The most general category shared among all jurisdictions was taxon, resulting in nine categories: amphibians, birds, bryophytes, fishes, fungus, invertebrates, mammals, reptiles, and vascular plants (set as a reference category). Lichens were removed from further analysis because there were only two species. For NZ and NSW, we extracted expert-elicited estimates of the benefit of implementing all management tasks, where experts were asked to consider the probability of species being secure in 50 years with and without the suite of management tasks16,17. Thus, benefit was calculated as the difference between the probability of security with and without the management tasks. For the U.S., in the absence of expert elicitation, the benefit of completing all management tasks in a recovery plan was approximated using information embedded in Recovery Priority Numbers (RPN). RPNs are an 18-category numeric rank for each species based on three categories of threat (high, moderate, and low), high or low recovery potential, and taxonomic distinctness monotypic genus, species, and subspecies,47. The limitations of using RPN to estimate the probability of persistence with or without management are discussed by Gerber et al.15 and Avery-Gomm48. To generate the total budget for each species, we used previously published total costs, which considered actions that benefited more than one species cost as shared among species projects15,16,17. In all further analysis, we removed species with a proposed budget of 0 (23 species in the U.S.) and extinct species (Guam broadbill—Myiagra freycineti and Eastern puma—Puma concolor couguar).
    We explored additional characteristics unique to U.S. recovery planning documents, using U.S. data only (Table 1). These included: the federal listing status, the number of species in the recovery plan (66% of plans include multiple species), the priority assigned to each management task (1: emergency measures needed to prevent extinction, 2: measures required to stabilize a species headed for extinction, and 3: needed to delist), the estimated management task duration in years, the fiscal year the management task was implemented, the management task status (ongoing, complete, planned, discontinued), the total estimated time to recovery, and an RPN, which we used to make a new factor called ‘recovery potential’ (one of six scores based on the RPN, where the highest had a high probability of recovery and low degree of threat and the lowest had a low probability of recovery and a high degree of threat). Federal listing status was collapsed from six into three categories: endangered, threatened, and not listed (including candidate species, species removed from ESA due to recovery, or populations considered as ‘non-essential, experimental’). Taxa were assigned to eight categories: amphibians, birds, fishes, invertebrates (set as a reference category), mammals, reptiles, and flowering and non-flowering plants.
    Quantitative analysis
    To examine what characteristics of recovery plans are associated with the proportion of the budget allocated to RM we used beta regression in the betareg package49 in R version 3.6.150. We fit two models—one including all data, with jurisdiction included as a covariate, and one including a wider suite of covariates only available for the U.S. (Table 1). All continuous covariates were standardized by subtracting the mean and dividing by the standard deviation to ensure the resulting parameter estimates would be comparable51. We standardized the total budget of each jurisdiction separately to account for each countries’ different currency and year the budget was estimated. To improve model fit we removed five species with total budgets over 5 million dollars (five times the median: Barton Springs salamander – Eurycea sosorum, Austin blind salamander—Eurycea waterlooensis, Indiana bat—Myotis sodalis, Bull trout – Salvelinus confluentus, Grizzly Bear—Ursus arctos horribilis). Our results are robust to the inclusion or exclusion of these species.
    Categorical covariates were converted to dummy variables. To select a reference category, we ran an initial model, using the category with the lowest mean proportion of budget RM as the reference. In this initial model, we selected the dummy variable with the highest variance inflation factor VIF in the car package52; as the reference in the final model. As a result, all VIF were 3). We excluded correlated covariates in successive models and chose the final model with the lowest Akaike’s Information Criterion (AIC53). The final model excluded the total proposed budget, which was correlated with the number of species in multi-species plans and the first fiscal year of the earliest RM. We consider any covariates where 95% confidence intervals around parameter estimates exclude zero to indicate a significant effect.
    Estimating species recovery outcomes
    To assess the relationship between the proportion of the budget allocated to RM and species recovery outcomes, we extracted a previously published index of recovery for U.S. listed species2 and developed similar indices based on annual and semi-annual reports from NZ and NSW (Supplementary Methods).
    To generate the U.S. recovery index, Gerber2 calculated sums of biennial status data from reports to Congress during 1989–2011 (total of 11 status reports30). For each species, reports included whether their status was extinct, declining (scored as −1), stable (scored as 0), improving (scored as +1), or unknown. These scores were summed, generating values from −11 to 11, indicating whether species are declining or improving more frequently.
    To develop recovery indices for NZ and NSW, we used similar reports through the New Zealand Threat Classification System and New South Wales Saving our Species annual report card over 4 and 5 assessment periods respectively. Assessments were annual in NSW and in NZ the periods between reports were on average every 4 years (Supplementary Methods). For each update or report card, we used a similar scoring (−1, 0, and +1) to indicate whether species were declining, stable, or improving between assessments (further details in Supplementary Methods). Note that in this analysis we were limited to a subset of the 2328 threatened species (78.5% of U.S. species, 13.5% of NZ species, and 14.7% of NSW species). Other studies have noted the limitations of recovery assessments28.
    Reporting summary
    Further information on research design is available in the Nature Research Reporting Summary linked to this article. More

  • in

    A network approach to elucidate and prioritize microbial dark matter in microbial communities

    Overall strategy to detect the relevance of Unknown taxa
    A pipeline based on network analysis was developed to detect and quantitatively measure the overall and individual impact of Unknown taxa on their environmental communities (Fig. 1). Briefly, Illumina 16S rRNA sequencing fastq files belonging to four distinctive aquatic extreme environments (i.e., hot springs, hypersaline, deep sea, and polar habitats that included both Arctic and Antarctic samples) were collected from public databases and 45 different BioProjects (Fig. 2a and Supplementary Dataset S1). We included different environment types to assess general and environment-specific patterns and chose to use extreme habitats as they comprise some of the harshest and relatively understudied habitats on Earth, and therefore, are likely to contain uncharacterized organisms.
    Fig. 1: Overview of the analysis pipeline.

    A minimum of 250 samples was retrieved for each of the four different extreme environments—hot springs (red), hypersaline (dark green), deep sea (turquoise), and polar (blue). Sequence reads were quality filtered, assigned to a taxonomy, and clustered to OTUs. At each classification level, any unassigned, ambiguous, or uncultured OTUs were designated as Unknowns, or “microbial dark matter” (MDM). For each environment, at each classification level, the direct co-occurrence relationship between all OTUs was mathematically modeled as a network. Networks were created for each environment, across all taxonomic classification levels, including all OTUs (Original, orange), excluding MDM (Without Unknowns, light green), and excluding an equal number of random Knowns (Bootstrap, blue). Network centrality metrics (i.e., degree, betweenness, and closeness) were calculated for each node, compared, and visualized as boxplots between these network types. Hub scores were calculated for each node in the Original network and networks were recreated, resizing by hub score, where the largest size node indicates a top hub species.

    Full size image

    Fig. 2: Summary of environmental 16S rRNA gene data.

    a Map of the sample sites used in this study. Circles symbolize sample locations and are color-coded by environment: hot springs (HS, red), hypersaline (HY, dark green), deep sea (DS, turquoise), and polar (PO, blue). b Summary of data used in this study. OTUs counts are provided at the genus level. c Proportion of “microbial dark matter” (MDM) OTUs for each environment labeled as unassigned (dark blue), uncultured (dark red), and ambiguous (yellow) after SILVA and UCLUST-based taxonomic assignment to OTU. d Venn diagram of shared OTUs in four extreme environments. Each pie chart depicts the proportion of unique OTUs that were Known (lighter shade) and Unknown (darker shade) for each environment, with the bottom-most pie chart showing combined data for all environments. e Prevalence curves indicate the number of unique OTUs consistently present at an increasing number of samples. Dotted lines signify the prevalence of MDM OTUs and solid lines signify the prevalence of Known OTUs.

    Full size image

    After quality filtering, reads were mapped to OTUs and annotated against the SILVA (v128) reference database [38] by an open-reference strategy, i.e., allowing the detection of Unknown OTUs [39]. Over two million Known and novel taxa were observed with the vast majority of the taxa annotated as Bacteria. Only the bacterial taxa or OTUs unclassified at the domain-level were targeted for downstream network analyses to demonstrate the feasibility of this approach across ecosystems. As the term “microbial dark matter” can have a broad meaning, here, we define Unknown taxa as uncultured, unassigned, or ambiguous by the reference database at each taxonomic classification level (e.g., phylum to genus). For each environment, networks reflecting across-samples co-occurrence relationships between all taxa, Known and Unknown, were constructed and referred to as the “Original” networks (Fig. 1). To assess the role of the Unknown taxa on network structure and properties, the Unknown nodes were removed from the ‘Original’ network and a new network, referred to as ‘Without Unknown’ was reconstructed. To ensure that changes in network properties were not caused just by the number of nodes, a third network, referred to as the “Bootstrap” network, was created where a random set of nodes of the same number as the Unknown OTUs was removed. The relevance of the Unknown taxa was assessed by comparing changes in degree, closeness, and betweenness scores between the three network types and by evaluating the frequency of Unknowns as top hubs within each of the “Original” environmental networks.
    A similar and significant fraction of Unknown taxa populates diverse environments
    To assess whether there were distinctive patterns or trends of Unknown taxa within the four targeted environments, data from each habitat type were mined from several geographical locations across the globe (Fig. 2a and Supplementary Dataset S1). Reads were collected from the online repositories National Center for Biotechnology Information Sequence Read Archive (NCBI SRA) and Joint Genome Institute Genomes Online Database (JGI Gold). For each environment, between 255 and 286 16S rRNA samples and between 51 and 57 million reads were included in the analysis, resulting in a total of 219,980,340 reads from 1086 samples (Fig. 2b).
    After processing, quality filtering, and OTU assignment steps were completed, there were 2102,595 unique OTUs totaling 164,896,127 amplicon read counts derived from these samples (Fig. 2b). The relative proportions of Unknown OTUs, which were designated as unassigned, ambiguous, or uncultured by the SILVA database, were compared between environments. Results indicated that all environments showed similar relative contributions of these three Unknown types, with unassigned and uncultured OTUs making up the majority of the Unknown component composition (Fig. 2c). Regardless of environment type, within each of the four habitats, between 25 and 38% of unique OTUs were cataloged as Unknown (Fig. 2b, d). Samples collected from polar habitats were significantly enriched (Fisher’s Exact Test p value < 0.05) in Unknown OTUs despite having the highest total read counts. The higher proportion of Unknown OTUs in the polar samples compared to the other habitats likely reflects the less-well-characterized biological diversity of these Arctic and Antarctic ecosystems. Next, the proportion of shared Known and Unknown OTUs between environments was evaluated. Most OTUs, regardless of assigned or unassigned taxonomic status, were environment-specific, with only 11,318 out of the 2,102,595 OTUs present in all four of the environments (Fig. 2d). The majority of shared OTUs were observed between the hypersaline and polar environments and between hypersaline and hot springs environments, possibly reflecting the widespread distribution of hypersaline habitats across diverse thermal zones. Unsurprisingly, polar and hot springs environmental samples shared the least number of OTUs (Fig. 2d). Given this low common OTU pool across environments, network analyses were applied to each environment independently. Last, the prevalence (i.e., the percentage of samples with nonzero counts in which an OTU was detected) of Known and Unknown OTUs was evaluated at the genus level to assess the consistency of OTU detection within each environment. The OTU matrix was sparse, with the majority of taxa observed in ≤50 (~20%) samples (Fig. 2e). However, prevalence curves were generally very similar for Known and Unknown OTUs in all four environments, indicating that Unknown OTUs are not necessarily rarer than already characterized species (Fig. 2e). Moreover, we confirmed that Unknown OTUs, like Known OTUs, were generally present and consistent across multiple studies within the same environment and did not tend to concentrate in any particular project (Supplementary Figs. S1–4). Consequently, these results indicated that a network analysis of these data would be a reflection of the co-occurrence structure of the community and not of potential compositional bias. Network analysis of OTU abundance at different taxonomic levels reveals the connectivity of unknown microbes Having demonstrated that the Unknown taxa comprise a substantial proportion of unique OTUs and have comparable abundances to Known taxa within a community, network metrics were used to effectively compare the ecological relevance of both Known and Unknown taxa in subsequent networks. Microbial association networks were constructed that featured only significant co-occurrence correlation relationships for OTUs with a notable prevalence in each environment, meaning that any OTUs that were not detected in a sufficient number of samples were removed. To select a suitable prevalence threshold, the proportion of Known and Unknown taxa across a range of sample percentages was evaluated (Supplementary Fig. S5). Across all taxonomic levels, the Known and Unknown taxa of hot springs and polar habitats were more prevalent than those of hypersaline and deep sea communities; therefore, a slightly more stringent prevalence threshold (40%) was chosen for the former, and a lower threshold value (30%) chosen for the latter environments. These thresholds resulted in the retention of a similar fraction of data from the initial OTU count (Table 1), with 102–297 nodes present per environment, making the networks both suitably large and comparable. Table 1 Breakdown of node and edge attributes across extreme environmental networks. Full size table The SpiecEasi Meinshausen–Buhlmann (MB) neighborhood algorithm was then used to construct networks (see “Methods”) that contained at least 100 nodes (i.e., OTUs) per environment and had edge-to-node ratios that varied from 1.9 and 2.9 (Table 1). Although most OTUs that passed the prevalence criteria became elements of the networks, no relationship between the initial data size (i.e., number of samples and taxa) and the interconnectedness (i.e., nodes/edges ratio) of the resulting network was observed (Table 1). The two environments with the highest number of initial OTUs (i.e., hypersaline and deep sea) had the lowest number of prevalent members, 193 and 102, respectively, and very different edge/node ratios, 2.7 and 1.9, respectively, indicating that high prevalence does not necessarily correlate with high co-occurrence. Similarly, the polar and hot springs networks retained a high number of prevalent OTUs but differed in edge number, yet again, indicating that the observed network structures were the result of the intrinsic properties of each environment and were not dependent on the sampling procedure. Interestingly, when evaluating the proportion of Unknown edges and Unknown-Unknown connections at the genus level, similar patterns were observed across environments. Between 45 and 62% of all connected nodes were Unknown OTUs and a higher proportion of Unknown-Unknown versus Known-Known links was present at the genus level (Table 1), for all environments but hot springs, where a higher proportion of Known-Unknown and Known-Known links was observed. The results of this global analysis of network construction and composition demonstrate that although the general community connectivity might be environment-specific, the relative contribution of Known and Unknown taxa to these networks is similar. Once again, network properties were not a direct outcome of sampling biases, but more likely, reflect the biology of their respective ecosystems. Unknown taxa play important roles in interconnectedness and connectivity of extreme environmental microbial networks Next, the position and neighborhood of Unknown nodes were examined. At the phylum level, Unknown taxa were present in the hot spring, hypersaline, and polar environments, but were not found in the deep sea network (Supplementary Fig. S6). The class level was the first taxonomic classification rank in which Unknown taxa were present in all environmental networks. To accurately assess the role of Unknowns, we evaluated class-level networks and observed that the hypersaline and polar Unknown OTUs created distinctive clusters, whereas hot springs and deep sea Unknown OTUs were intermixed with Known taxa (Fig. 3a). Hypersaline and polar Unknowns consistently appeared to be isolated and peripherally located compared to the centrally positioned hot springs and deep sea Unknown nodes across almost all classification ranks (Supplementary Fig. S6). Consequently, these results suggest that the clustering pattern is unrelated to a higher abundance of Unknowns and is more environment dependent. For example, the targeted hot spring environments had the highest number of Unknown OTUs yet showed the most dispersed connections between the these taxa (Fig. 3a). Thus, the inclusion of the Unknown taxa in our environmental networks models was, as anticipated, not solely the consequence of their level of prevalence, but rather a reflection of a particular abundance co-variation pattern. Fig. 3: Analysis of environmental network taxa interconnectedness. a Microbial networks at class taxonomic classification level. Nodes are colored by class assignment, with gray nodes representing Unknown taxa at the class level. b Bar graphs of the co-occurrence relationships (i.e., edges) of Unknown OTUs with other taxa at the class level within each environmental network. Y axis labels and colors signify the different classes with which Unknowns were found to co-occur. Unknown-Unknown relationships are represented in gray. Full size image For all four environments, Unknown OTUs had more frequently shared edges among them than with classified taxa (Fig. 3b). Unknown OTUs were found to frequently co-occur with each other within each environmental network, although the frequency of within-class interactions for Unknowns at the class level was found to be statistically no greater than all other within-class interactions for each environment (Supplementary Fig. S7). In fact, the pattern of a high frequency of shared edges among members of the same class held true for known classes as well (Supplementary Fig. S8). In accordance with other studies, these results demonstrate that OTUs of the same taxonomic classification most frequently co-occur with each other [40, 41]. Furthermore, the high frequency of shared edges between Unknown classes suggests that Unknown OTUs might be taxonomically related. To ensure that the observations found were reproducible, robust, and not biased by earlier steps of the analysis, the diversity and position of Unknown taxa in the networks were examined for several parameters. Although the number of Unknown nodes changed at each level (Supplementary Fig. S6), the environment-specific network patterns observed at the class level (Fig. 3a) were retained at other taxonomic levels. In addition, to determine whether the topology of the network was a direct consequence of our correlation metric of choice or the prevalence threshold, three other network construction approaches were used: SparCC [11], CClasso [42], and Pearson correlation. Network analyses were performed across a range of prevalence thresholds (15–35%, at 5% increments). Again, regardless of the network construction approach or sample percentage applied, network shape and Unknown OTU position remained consistent and each environment exhibited a distinctive pattern of Unknown taxa inclusion. For example, Unknown nodes continued to occupy peripheral positions for hypersaline and polar networks, whereas nodes in hot springs and deep sea environments were more centrally located when applying different correlation metrics (Supplementary Fig. S9). Although networks appeared “noisy” at more lenient prevalence thresholds (15–20%; Supplementary Figs. S10–13), the networks and positioning of Unknowns at higher percent thresholds were similar in appearance to the “Original” networks for all environments. Based on these results, we found that our general analysis strategy was robust across parameter choices and, therefore, these networks captured critical features of the relationships among taxa within each distinctive environment. Microbial dark matter acts as unifiers and frequent hubs within extreme environmental networks Although these results suggest that Unknown taxa were highly interconnected, these observations did not reveal how the presence of Unknowns affected the overall community structure. To more fully understand this role, we analyzed how network properties changed in the presence and absence of Unknown OTU nodes. We evaluated changes in degree, betweenness, and closeness, as different network metrics reveal different aspects of the relevance of nodes within their networks. This approach has the potential to ascertain whether certain Unknown OTUs were more centrally positioned (e.g., due to high closeness scores), more essential for joining other taxa (e.g., high betweenness), or simply more prevalent and likely to co-occur with others (e.g., high degree). To control for the effect of node removal and distinguish effects of Unknown taxa from network size, networks were generated that excluded several randomly picked Known nodes equal to the number of Unknown OTUs. This process was repeated 100 times to create a distribution of “Null” or “Bootstrap” networks for statistical comparisons. Differences in network parameters between networks without Unknown OTUs and the “Original” or “Bootstrap” networks were determined by the Wilcoxon test and p values were adjusted using the Holm method [43]. Strikingly, removal of the Unknown taxa caused a statistically significant impact on all measured network metrics in all four studied environments (Table 2, Fig. 4, and Supplementary Figs. S14–24). In the polar environments, for example, removal of Unknown OTUs caused a significant decrease in degree (p value  More

  • in

    Population genetic structure of the great star coral, Montastraea cavernosa, across the Cuban archipelago with comparisons between microsatellite and SNP markers

    1.
    Carson, H. S., Cook, G. S., López-Duarte, P. C. & Levin, L. A. Evaluating the importance of demographic connectivity in a marine metapopulation. Ecology 92, 1972–1984. https://doi.org/10.1890/11-0488.1 (2011).
    PubMed  Article  Google Scholar 
    2.
    Jackson, J. B. C., Donovan, M. K., Cramer, K. L., Lam, V. & Lam, W. Status and trends of Caribbean coral reefs: 1970–2012. Glob. Coral Reef Monit. Network, IUCN, Gland. Switz. 306 (2014).

    3.
    Palumbi, S. R. Population genetics, demographic connectivity, and the design of marine reserves. 13, 146–158, https://doi.org/10.1890/1051-0761(2003)013[0146:PGDCAT]2.0.CO;2 (2003).

    4.
    Botsford, L. W. et al. Connectivity and resilience of coral reef metapopulations in marine protected areas: matching empirical efforts to predictive needs. Coral Reefs 28, 327–337. https://doi.org/10.1007/s00338-009-0466-z (2009).
    ADS  PubMed  PubMed Central  CAS  Article  Google Scholar 

    5.
    Galindo, H. M., Olson, D. B. & Palumbi, S. R. Seascape genetics: a coupled oceanographic-genetic model predicts population structure of Caribbean corals. Curr. Biol. 16, 1622–1626. https://doi.org/10.1016/j.cub.2006.06.052 (2006).
    PubMed  CAS  Article  Google Scholar 

    6.
    Rippe, J. P. et al. Population structure and connectivity of the mountainous star coral, Orbicella faveolata, throughout the wider Caribbean region. Ecol. Evol. 7, 9234–9246. https://doi.org/10.1002/ece3.3448 (2017).
    PubMed  PubMed Central  Article  Google Scholar 

    7.
    Studivan, M. S. & Voss, J. D. Population connectivity among shallow and mesophotic Montastraea cavernosa corals in the Gulf of Mexico identifies potential for refugia. Coral Reefs 37, 1183–1196. https://doi.org/10.1007/s00338-018-1733-7 (2018).
    ADS  Article  Google Scholar 

    8.
    Baums, I. B., Johnson, M. E., Devlin-Durante, M. K. & Miller, M. W. Host population genetic structure and zooxanthellae diversity of two reef-building coral species along the Florida Reef Tract and wider Caribbean. Coral Reefs 29, 835–842. https://doi.org/10.1007/s00338-010-0645-y (2010).
    ADS  Article  Google Scholar 

    9.
    Serrano, X. M. et al. Long distance dispersal and vertical gene flow in the Caribbean brooding coral Porites astreoides. Sci. Rep. 6, 21619. https://doi.org/10.1038/srep21619 (2016).
    ADS  PubMed  PubMed Central  CAS  Article  Google Scholar 

    10.
    Serrano, X. et al. Geographic differences in vertical connectivity in the Caribbean coral Montastraea cavernosa despite high levels of horizontal connectivity at shallow depths. Mol. Ecol. 23, 4226–4240. https://doi.org/10.1111/mec.12861 (2014).
    PubMed  CAS  Article  Google Scholar 

    11.
    Bongaerts, P. et al. Deep reefs are not universal refuges: reseeding potential varies among coral species. Sci. Adv. 3, e1602373. https://doi.org/10.1126/sciadv.1602373 (2017).
    ADS  PubMed  PubMed Central  Article  Google Scholar 

    12.
    Eckert, R. J., Studivan, M. S. & Voss, J. D. Populations of the coral species Montastraea cavernosa on the Belize Barrier Reef lack vertical connectivity. Sci. Rep. 9, 7200. https://doi.org/10.1038/s41598-019-43479-x (2019).
    ADS  PubMed  PubMed Central  CAS  Article  Google Scholar 

    13.
    Goodbody-Gringley, G., Vollmer, S. V., Woollacott, R. M. & Giribet, G. Limited gene flow in the brooding coral Favia fragum (Esper, 1797). Mar. Biol. 157, 2591–2602. https://doi.org/10.1007/s00227-010-1521-6 (2010).
    Article  Google Scholar 

    14.
    Goodbody-Gringley, G., Woollacott, R. M. & Giribet, G. Population structure and connectivity in the Atlantic scleractinian coral Montastraea cavernosa (Linnaeus, 1767). Mar. Ecol. 33, 32–48. https://doi.org/10.1111/j.1439-0485.2011.00452.x (2012).
    ADS  CAS  Article  Google Scholar 

    15.
    Nunes, F. L. D., Norris, R. D. & Knowlton, N. Long distance dispersal and connectivity in Amphi-Atlantic corals at regional and basin scales. PLoS ONE 6, e22298. https://doi.org/10.1371/journal.pone.0022298 (2011).
    ADS  PubMed  PubMed Central  CAS  Article  Google Scholar 

    16.
    Brazeau, D. A., Lesser, M. P. & Slattery, M. Genetic structure in the coral, Montastraea cavernosa: assessing genetic differentiation among and within mesophotic reefs. PLoS ONE 8, e65845. https://doi.org/10.1371/journal.pone.0065845 (2013).
    ADS  PubMed  PubMed Central  CAS  Article  Google Scholar 

    17.
    Foster, N. L. et al. Connectivity of Caribbean coral populations: complementary insights from empirical and modelled gene flow. Mol. Ecol. 21, 1143–1157. https://doi.org/10.1111/j.1365-294X.2012.05455.x (2012).
    PubMed  Article  Google Scholar 

    18.
    García-Machado, E., Ulmo-Díaz, G., Castellanos-Gell, J. & Casane, D. Patterns of population connectivity in marine organisms of Cuba. Bull. Mar. Sci. 94, 193–211. https://doi.org/10.5343/bms.2016.1117 (2018).
    Article  Google Scholar 

    19.
    Ulmo-Díaz, G. et al. Genetic differentiation in the mountainous star coral Orbicella faveolata around Cuba. Coral Reefs 37, 1217–1227. https://doi.org/10.1007/s00338-018-1722-x (2018).
    ADS  Article  Google Scholar 

    20.
    Creary, M. et al. Status of coral reefs in the northern Caribbean and western Atlantic GCRMN node in 2008. in Status of Coral Reefs of the World (ed. Wilkinson, C.) 239–252 (2008).

    21.
    Claro, R., Reshetnikov, Y. S. & Alcolado, P. M. Physical attributes of coastal Cuba. Ecol. Mar. fishes Cuba 1–20 (2001).

    22.
    Whittle, D. & Rey Santos, O. Protecting Cuba’s environment: efforts to design and implement effective environmental laws and policies in Cuba. Cuban Stud. 37, 73–103. https://doi.org/10.1353/cub.2007.0018 (2006).
    Article  Google Scholar 

    23.
    Caballero, H., Alcolado, P. M. & Semidey, A. Condición de los arrecifes de coral frente a costas con asentamientos humanos y aportes terrígenos: el caso del litoral habanero. Cuba. Rev. Ciencias Mar. y Costeras 1, 49. https://doi.org/10.15359/revmar.1.3 (2009).
    Article  Google Scholar 

    24.
    González-Díaz, P. et al. Status of Cuban coral reefs. Bull. Mar. Sci. 94, 229–247. https://doi.org/10.5343/bms.2017.1035 (2018).
    Article  Google Scholar 

    25.
    Alcolado, P. M., Caballero, H. & Perera, S. Tendencia del cambio en el cubrimiento vivo por corales pétreos en los arrecifes coralinos de Cuba. Ser. Ocean. 5, 1–14 (2009).
    Google Scholar 

    26.
    Toth, L. T. et al. Do no-take reserves benefit Florida’s corals? 14 years of change and stasis in the Florida Keys National Marine Sanctuary. Coral Reefs 33, 565–577. https://doi.org/10.1007/s00338-014-1158-x (2014).
    ADS  Article  Google Scholar 

    27.
    Zlatarski, V. N. & Estalella, N. M. Los esclaractinios de Cuba. (2017).

    28.
    Zlatarski, V. N. Investigations on mesophotic coral ecosystems in Cuba (1970–1973) and Mexico (1983–1984). CICIMAR Oceánides 33, 27–43 (2018).
    Google Scholar 

    29.
    Reed, J. et al. Cuba’s mesophotic coral reefs and associated fish communities. Rev. Investig. Mar. 38, 56–125 (2018).
    Google Scholar 

    30.
    Baisre, J. A. An overview of Cuban commercial marine fisheries: the last 80 years. Bull. Mar. Sci. 94, 359–375. https://doi.org/10.5343/bms.2017.1015 (2018).
    Article  Google Scholar 

    31.
    Gil-Agudelo, D. L. et al. Coral reefs in the Gulf of Mexico large marine ecosystem: conservation status, challenges, and opportunities. Front. Mar. Sci. 6, 807. https://doi.org/10.3389/fmars.2019.00807 (2020).
    Article  Google Scholar 

    32.
    Perera Valderrama, S. et al. Marine protected areas in Cuba. Bull. Mar. Sci. 94, 423–442. https://doi.org/10.5343/bms.2016.1129 (2018).
    Article  Google Scholar 

    33.
    NOAA. Sister Sanctuary: Memorandum of Understanding. (2015).

    34.
    Budd, A. F., Nunes, F. L. D., Weil, E. & Pandolfi, J. M. Polymorphism in a common Atlantic reef coral (Montastraea cavernosa) and its long-term evolutionary implications. Evol. Ecol. 26, 265–290. https://doi.org/10.1007/s10682-010-9460-8 (2012).
    Article  Google Scholar 

    35.
    Bongaerts, P., Ridgway, T., Sampayo, E. M. & Hoegh-Guldberg, O. Assessing the `deep reef refugia’ hypothesis: focus on Caribbean reefs. Coral Reefs 29, 309–327. https://doi.org/10.1007/s00338-009-0581-x (2010).
    Article  Google Scholar 

    36.
    Reed, J. K. Deepest distribution of Atlantic hermaptypic corals discovered in the Bahamas. Proc. Fifth Int. Coral Reef Congr. 6, 249–254 (1985).
    Google Scholar 

    37.
    Szmant, A. M. Sexual reproduction by the Caribbean reef corals Montastrea annularis and M. cavernosa. Mar. Ecol. Prog. Ser. 74, 13–25. https://doi.org/10.3354/meps074013 (1991).
    ADS  Article  Google Scholar 

    38.
    Highsmith, R. C., Lueptow, R. L. & Schonberg, S. C. Growth and bioerosion of three massive corals on the Belize barrier reef. Mar. Ecol. Prog. Ser. 13, 261–271 (1983).
    ADS  Article  Google Scholar 

    39.
    Kitchen, S. A., Crowder, C. M., Poole, A. Z., Weis, V. M. & Meyer, E. De novo assembly and characterization of four anthozoan (Phylum Cnidaria) transcriptomes. G3 (Genes|Genomes|Genetics). 5, 2441–2452, https://doi.org/10.1534/g3.115.020164 (2015).
    PubMed  PubMed Central  CAS  Article  Google Scholar 

    40.
    Matz Lab. Montastraea cavernosa annotated genome. (2018).

    41.
    Drury, C., Pérez Portela, R., Serrano, X. M., Oleksiak, M. & Baker, A. C. Fine-scale structure among mesophotic populations of the great star coral Montastraea cavernosa revealed by SNP genotyping. Ecol. Evol. 1–11, https://doi.org/10.1002/ece3.6340 (2020).

    42.
    Ellegren, H. Microsatellites: simple sequences with complex evolution. Nat. Rev. Genet. 5, 435–445. https://doi.org/10.1038/nrg1348 (2004).
    PubMed  CAS  Article  Google Scholar 

    43.
    Joshi, D., Ram, R. N. & Lohani, P. Microsatellite markers and their application in fisheries. Int. J. Adv. Agric. Sci. Technol. 4, 67–104 (2017).
    Google Scholar 

    44.
    Jarne, P. & Lagoda, P. J. L. Microsatellites, from molecules to populations and back. Trends Ecol. Evol. 11, 424–429. https://doi.org/10.1016/0169-5347(96)10049-5 (1996).
    PubMed  CAS  Article  Google Scholar 

    45.
    Flores-Rentería, L. & Krohn, A. Scoring microsatellite loci. in Methods in molecular biology (ed. Kantartzi, S. K.) 319–336, https://doi.org/10.1007/978-1-62703-389-3_21 (Elsevier, 2013).

    46.
    Andrews, K. R., Good, J. M., Miller, M. R., Luikart, G. & Hohenlohe, P. A. Harnessing the power of RADseq for ecological and evolutionary genomics. Nat. Rev. Genet. 17, 81–92. https://doi.org/10.1038/nrg.2015.28 (2016).
    PubMed  PubMed Central  CAS  Article  Google Scholar 

    47.
    Davey, J. L. & Blaxter, M. W. RADseq: next-generation population genetics. Brief. Funct. Genomics 9, 416–423. https://doi.org/10.1093/bfgp/elq031 (2010).
    PubMed  CAS  Article  Google Scholar 

    48.
    Wang, S., Meyer, E., McKay, J. K. & Matz, M. V. 2b-RAD: a simple and flexible method for genome-wide genotyping. Nat. Methods 9, 808–810. https://doi.org/10.1038/nmeth.2023 (2012).
    PubMed  CAS  Article  Google Scholar 

    49.
    Bradbury, I. R. et al. Transatlantic secondary contact in Atlantic Salmon, comparing microsatellites, a single nucleotide polymorphism array and restriction-site associated DNA sequencing for the resolution of complex spatial structure. Mol. Ecol. 24, 5130–5144. https://doi.org/10.1111/mec.13395 (2015).
    PubMed  CAS  Article  Google Scholar 

    50.
    Jeffries, D. L. et al. Comparing RADseq and microsatellites to infer complex phylogeographic patterns, an empirical perspective in the Crucian carp, Carassius carassius. L. Mol. Ecol. 25, 2997–3018. https://doi.org/10.1111/mec.13613 (2016).
    PubMed  Article  Google Scholar 

    51.
    Bohling, J., Small, M., Von Bargen, J., Louden, A. & DeHaan, P. Comparing inferences derived from microsatellite and RADseq datasets: a case study involving threatened bull trout. Conserv. Genet. 20, 329–342. https://doi.org/10.1007/s10592-018-1134-z (2019).
    CAS  Article  Google Scholar 

    52.
    Thornhill, D. J., Xiang, Y., Fitt, W. K. & Santos, S. R. Reef endemism, host specificity and temporal stability in populations of symbiotic dinoflagellates from two ecologically dominant Caribbean corals. PLoS ONE https://doi.org/10.1371/journal.pone.0006262 (2009).
    PubMed  PubMed Central  Article  Google Scholar 

    53.
    Eckert, R. J., Reaume, A. M., Sturm, A. B., Studivan, M. S. & Voss, J. D. Depth influences Symbiodiniaceae associations among Montastraea cavernosa corals on the Belize Barrier Reef. Front. Microbiol. 11, 1–13. https://doi.org/10.3389/fmicb.2020.00518 (2020).
    Article  Google Scholar 

    54.
    Hume, B. C. C. et al. SymPortal: a novel analytical framework and platform for coral algal symbiont next-generation sequencing ITS2 profiling. Mol. Ecol. Resour. 19, 1063–1080. https://doi.org/10.1111/1755-0998.13004 (2019).
    PubMed  PubMed Central  CAS  Article  Google Scholar 

    55.
    Pochon, X., Putnam, H. M., Burki, F. & Gates, R. D. Identifying and characterizing alternative molecular markers for the symbiotic and free-living dinoflagellate genus Symbiodinium. PLoS ONE https://doi.org/10.1371/journal.pone.0029816 (2012).
    PubMed  PubMed Central  Article  Google Scholar 

    56.
    LaJeunesse, T. C. & Thornhill, D. J. Improved resolution of reef-coral endosymbiont (Symbiodinium) species diversity, ecology, and evolution through psbA non-coding region genotyping. PLoS ONE https://doi.org/10.1371/journal.pone.0029013 (2011).
    PubMed  PubMed Central  Article  Google Scholar 

    57.
    Manzello, D. P. et al. Role of host genetics and heat-tolerant algal symbionts in sustaining populations of the endangered coral Orbicella faveolata in the Florida Keys with ocean warming. Glob. Change Biol. 25, 1016–1031. https://doi.org/10.1111/gcb.14545 (2018).
    ADS  Article  Google Scholar 

    58.
    Warner, M. E., LaJeunesse, T. C., Robison, J. D. & Thur, R. M. The ecological distribution and comparative photobiology of symbiotic dinoflagellates from reef corals in Belize: potential implications for coral bleaching. Limnol Ocean. https://doi.org/10.4319/lo.2006.51.4.1887 (2006).
    Article  Google Scholar 

    59.
    Finney, J. C. et al. The relative significance of host–habitat, depth, and geography on the ecology, endemism, and speciation of coral endosymbionts in the genus Symbiodinium. Microb. Ecol. 60, 250–263. https://doi.org/10.1007/s00248-010-9681-y (2010).
    PubMed  Article  Google Scholar 

    60.
    Bongaerts, P. et al. Deep down on a Caribbean reef: lower mesophotic depths harbor a specialized coral-endosymbiont community. Sci. Rep. https://doi.org/10.1038/srep07652 (2015).
    PubMed  PubMed Central  Article  Google Scholar 

    61.
    Reed, J. et al. Cruise report Cuba’s twilight zone reefs: Remotely Operated Vehicle surveys of deep/mesophotic coral reefs and associated fish communities of Cuba. (2017).

    62.
    Arriaza, L. et al. Modelación numérica de corrientes marinas alrededor del occidente de Cuba. Serie Oceanológica. 10, 11–22. (2012).
    Google Scholar 

    63.
    Gordon, A. & Hannon, G. FASTX-Toolkit. FASTQ/A short-reads pre-processing tools. (2010). Available at: https://hannonlab.cshl.edu/fastx_toolkit/.

    64.
    Bayer, T. et al. Symbiodinium transcriptomes: genome insights into the dinoflagellate symbionts of reef-building corals. PLoS ONE https://doi.org/10.1371/journal.pone.0035269 (2012).
    PubMed  PubMed Central  Article  Google Scholar 

    65.
    Davies, S. W., Ries, J. B., Marchetti, A. & Castillo, K. D. Symbiodinium functional diversity in the coral Siderastrea siderea is influenced by thermal stress and reef environment, but not ocean acidification. Front. Mar. Sci. 5, 1–14. https://doi.org/10.3389/fmars.2018.00150 (2018).
    Article  Google Scholar 

    66.
    Ladner, J. T., Barshis, D. J. & Palumbi, S. R. Protein evolution in two co-occurring types of Symbiodinium: an exploration into the genetic basis of thermal tolerance in Symbiodinium clade D. BMC Evol. Biol. 12, 217. https://doi.org/10.1186/1471-2148-12-217 (2012).
    PubMed  PubMed Central  CAS  Article  Google Scholar 

    67.
    Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359. https://doi.org/10.1038/nmeth.1923 (2012).
    PubMed  PubMed Central  CAS  Article  Google Scholar 

    68.
    Korneliussen, T. S., Albrechtsen, A. & Nielsen, R. ANGSD: Analysis of Next Generation Sequencing Data. BMC Bioinformatics 15, 356. https://doi.org/10.1186/s12859-014-0356-4 (2014).
    PubMed  PubMed Central  Article  Google Scholar 

    69.
    Peakall, R. & Smouse, P. E. GenALEx 65: Genetic analysis in Excel. Population genetic software for teaching and research-an update. Bioinformatics 28, 2537–2539. https://doi.org/10.1093/bioinformatics/bts460 (2012).
    PubMed  PubMed Central  CAS  Article  Google Scholar 

    70.
    Excoffier, L., Smouse, P. E. & Quattro, J. M. Analysis of Molecular Variance inferred from metric distances among DNA haplotypes: application to human mitochondrial DNA restriction data. Genetics 131, 479–491 (1992).
    PubMed  PubMed Central  CAS  Google Scholar 

    71.
    Benjamini, Y. & Hochberg, Y. Controlling the False Discovery Rate: a practical and powerful approach to multiple testing. J. R. Stat. Soc. Ser. B 57, 289–300. https://doi.org/10.1111/j.2517-6161.1995.tb02031.x (1995).
    MathSciNet  MATH  Article  Google Scholar 

    72.
    Prevosti, A., Ocaña, J. & Alonso, G. Distances between populations of Drosophila subobscura, based on chromosome arrangement frequencies. Theor. Appl. Genet. 45, 231–241. https://doi.org/10.1007/BF00831894 (1975).
    PubMed  CAS  Article  Google Scholar 

    73.
    Kamvar, Z. N., Tabima, J. F. & Gr̈unwald, N. J. ,. Poppr: An R package for genetic analysis of populations with clonal, partially clonal, and/or sexual reproduction. PeerJ 2014, 1–14. https://doi.org/10.7717/peerj.281 (2014).
    Article  Google Scholar 

    74.
    R Core Team. R: A language and environment for statistical computing. (2019).

    75.
    Jombart, T. & Ahmed, I. adegenet 13–1: new tools for the analysis of genome-wide SNP data. Bioinformatics 27, 3070–3071. https://doi.org/10.1093/bioinformatics/btr521 (2011).
    PubMed  PubMed Central  CAS  Article  Google Scholar 

    76.
    Mantel, N. The detection of disease clustering and a generalized regression approach. Cancer Res. 27, 209–220 (1967).
    PubMed  CAS  Google Scholar 

    77.
    Pritchard, J. K., Stephens, M. & Donnelly, P. Inference of population structure using multilocus genotype data. Genetics 155, 945–959 (2000).
    PubMed  PubMed Central  CAS  Google Scholar 

    78.
    Besnier, F. & Glover, K. A. ParallelStructure: A R package to distribute parallel runs of the population genetics program STRUCTURE on multi-core computers. PLoS ONE 8, 1–5. https://doi.org/10.1371/journal.pone.0070651 (2013).
    CAS  Article  Google Scholar 

    79.
    Earl, D. A. & vonHoldt, B. M. STRUCTURE HARVESTER: A website and program for visualizing STRUCTURE output and implementing the Evanno method. Conserv. Genet. Resour. 4, 359–361. https://doi.org/10.1007/s12686-011-9548-7 (2012).
    Article  Google Scholar 

    80.
    Catchen, J., Hohenlohe, P. A., Bassham, S., Amores, A. & Cresko, W. A. Stacks: an analysis tool set for population genomics. Mol. Ecol. 22, 3124–3140. https://doi.org/10.1111/mec.12354 (2013).
    PubMed  PubMed Central  Article  Google Scholar 

    81.
    Pembleton, L. W., Cogan, N. O. I. & Forster, J. W. StAMPP: An R package for calculation of genetic differentiation and structure of mixed-ploidy level populations. Mol. Ecol. Resour. 13, 946–952. https://doi.org/10.1111/1755-0998.12129 (2013).
    PubMed  CAS  Article  Google Scholar 

    82.
    Oksanen, J. et al. Vegan: community ecology package. (2019).

    83.
    Alexander, D. H. D., Novembre, J. & Lange, K. Fast model-based estimation of ancestry in unrelated individuals. Genome Res. 19, 1655–1664. https://doi.org/10.1101/gr.094052.109 (2009).
    PubMed  PubMed Central  CAS  Article  Google Scholar 

    84.
    Skotte, L., Korneliussen, T. S. & Albrechtsen, A. Estimating individual admixture proportions from next generation sequencing data. Genetics 195, 693–702. https://doi.org/10.1534/genetics.113.154138 (2013).
    PubMed  PubMed Central  CAS  Article  Google Scholar 

    85.
    Foll, M. & Gaggiotti, O. A genome-scan method to identify selected loci appropriate for both dominant and codominant markers: a Bayesian perspective. Genetics 180, 977–993. https://doi.org/10.1534/genetics.108.092221 (2008).
    PubMed  PubMed Central  Article  Google Scholar 

    86.
    Evanno, G., Regnaut, S. & Goudet, J. Detecting the number of clusters of individuals using the software STRUCTURE: a simulation study. Mol. Ecol. 14, 2611–2620. https://doi.org/10.1111/j.1365-294X.2005.02553.x (2005).
    PubMed  CAS  Article  Google Scholar 

    87.
    Centurioni, L. R. & Niiler, P. P. On the surface currents of the Caribbean Sea. Geophys. Res. Lett. 30, 10–13. https://doi.org/10.1029/2002GL016231 (2003).
    Article  Google Scholar 

    88.
    Candela, J. et al. The flow through the gulf of Mexico. J. Phys. Oceanogr. 49, 1381–1401. https://doi.org/10.1175/JPO-D-18-0189.1 (2019).
    ADS  Article  Google Scholar 

    89.
    Kourafalou, V., Androulidakis, Y., Le Hénaff, M. & Kang, H. S. The Dynamics of Cuba Anticyclones (CubANs) and interaction with the Loop Current/Florida Current system. J. Geophys. Res. Ocean. 122, 7897–7923. https://doi.org/10.1002/2017JC012928 (2017).
    ADS  Article  Google Scholar 

    90.
    Arriaza, L. et al. Marine current estimations in southeast Cuban shelf. Ser. Ocean. 4, 1–10 (2008).
    Google Scholar 

    91.
    Carracedo-Hidalgo, D., Reyes-Perdomo, D., Calzada-estrada, A., Chang-Domínguez, D. & Rodríguez-Pupo, A. Characterization of sea currents in sea adjacent to Cuba . Main trends in the last years. Rev. Cuba. Meteorol. 25, (2019).

    92.
    Frys, C. et al. Fine-scale coral connectivity pathways in the Florida Reef Tract: implications for conservation and restoration. Front. Mar. Sci. 7, 1–42. https://doi.org/10.3389/fmars.2020.00312 (2020).
    Article  Google Scholar 

    93.
    Kuba, A. Transgenerational effects of thermal stress: impacts on and beyond coral reproduction. (Nova Southeastern University, 2016).

    94.
    Claro, R., Lindeman, K. C., Kough, A. S. & Paris, C. B. Biophysical connectivity of snapper spawning aggregations and marine protected area management alternatives in Cuba. Fish. Oceanogr. 28, 33–42. https://doi.org/10.1111/fog.12384 (2019).
    Article  Google Scholar 

    95.
    Holstein, D. M., Paris, C. B. & Mumby, P. J. Consistency and inconsistency in multispecies population network dynamics of coral reef ecosystems. Mar. Ecol. Prog. Ser. 499, 1–18. https://doi.org/10.3354/meps10647 (2014).
    ADS  Article  Google Scholar 

    96.
    Szmant, A. M. Reproductive ecology of Caribbean reef corals. Coral Reefs 5, 43–53. https://doi.org/10.1007/BF00302170 (1986).
    ADS  Article  Google Scholar 

    97.
    Drury, C. et al. Genomic variation among populations of threatened coral: Acropora cervicornis. BMC Genomics 17, 286. https://doi.org/10.1186/s12864-016-2583-8 (2016).
    PubMed  PubMed Central  CAS  Article  Google Scholar 

    98.
    Devlin-Durante, M. K. & Baums, I. B. Genome-wide survey of single-nucleotide polymorphisms reveals fine-scale population structure and signs of selection in the threatened Caribbean elkhorn coral. Acropora palmata. PeerJ 5, e4077. https://doi.org/10.7717/peerj.4077 (2017).
    PubMed  CAS  Article  Google Scholar 

    99.
    Wang, J., Feng, C., Jiao, T., Von Wettberg, E. B. & Kang, M. Genomic signature of adaptive divergence despite strong nonadaptive forces on Edaphic Islands: a case study of Primulina juliae. Genome Biol. Evol. 9, 3495–3508. https://doi.org/10.1093/gbe/evx263 (2017).
    PubMed  PubMed Central  Article  Google Scholar 

    100.
    Ramos-Silva, P. et al. The skeletal proteome of the coral Acropora millepora: the evolution of calcification by co-option and domain shuffling. Mol. Biol. Evol. 30, 2099–2112. https://doi.org/10.1093/molbev/mst109 (2013).
    PubMed  PubMed Central  CAS  Article  Google Scholar 

    101.
    Takeuchi, T., Yamada, L., Shinzato, C., Sawada, H. & Satoh, N. Stepwise evolution of coral biomineralization revealed with genome-wide proteomics and transcriptomics. PLoS ONE https://doi.org/10.1371/journal.pone.0156424 (2016).
    PubMed  PubMed Central  Article  Google Scholar 

    102.
    Aranda, M. et al. Differential sensitivity of coral larvae to natural levels of ultraviolet radiation during the onset of larval competence. Mol. Ecol. 20, 2955–2972. https://doi.org/10.1111/j.1365-294X.2011.05153.x (2011).
    PubMed  CAS  Article  Google Scholar 

    103.
    Reynolds, W. S., Schwarz, J. A. & Weis, V. M. Symbiosis-enhanced gene expression in cnidarian-algal associations: cloning and characterization of a cDNA, sym32, encoding a possible cell adhesion protein. Comp. Biochem. Physiol. Part A Mol. Integr. Physiol. 126, 33–44. https://doi.org/10.1016/S0742-8413(00)00099-2 (2000).
    CAS  Article  Google Scholar 

    104.
    Iguchi, A. et al. Apparent involvement of a β1 type integrin in coral fertilization. Mar. Biotechnol. 9, 760–765. https://doi.org/10.1007/s10126-007-9026-0 (2007).
    PubMed  CAS  Article  Google Scholar 

    105.
    Lesser, M. P. et al. Photoacclimatization by the coral Montastraea cavernosa in the mesophotic zone: light, food, and genetics. Ecology 91, 990–1003. https://doi.org/10.1890/09-0313.1 (2010).
    PubMed  Article  Google Scholar 

    106.
    Klepac, C. et al. Seasonal stability of coral-Symbiodinium associations in the subtropical coral habitat of St. Lucie Reef, Florida. Mar. Ecol. Prog. Ser. 532, 137–151. https://doi.org/10.3354/meps11369 (2015).
    ADS  Article  Google Scholar 

    107.
    Polinski, J. M. & Voss, J. D. Evidence of photoacclimatization at mesophotic depths in the coral-Symbiodinium symbiosis at Flower Garden Banks National Marine Sanctuary and McGrail Bank. Coral Reefs 37, 779–789. https://doi.org/10.1007/s00338-018-1701-2 (2018).
    ADS  Article  Google Scholar 

    108.
    LaJeunesse, T. C. et al. Systematic revision of Symbiodiniaceae highlights the antiquity and diversity of coral endosymbionts. Curr. Biol. 28, 1–11. https://doi.org/10.1016/j.cub.2018.07.008 (2018).
    CAS  Article  Google Scholar 

    109.
    Swain, T. D., Chandler, J., Backman, V. & Marcelino, L. Consensus thermotolerance ranking for 110 Symbiodinium phylotypes: an exemplar utilization of a novel iterative partial-rank aggregation tool with broad application potential. Funct. Ecol. 31, 172–183. https://doi.org/10.1111/1365-2435.12694 (2017).
    Article  Google Scholar 

    110.
    Hodel, R. G. J. et al. Adding loci improves phylogeographic resolution in red mangroves despite increased missing data: comparing microsatellites and RAD-Seq and investigating loci filtering. Sci. Rep. 7, 1–14. https://doi.org/10.1038/s41598-017-16810-7 (2017).
    CAS  Article  Google Scholar 

    111.
    Lemopoulos, A. et al. Comparing RADseq and microsatellites for estimating genetic diversity and relatedness — implications for brown trout conservation. Ecol. Evol. 9, 2106–2120. https://doi.org/10.1002/ece3.4905 (2019).
    PubMed  PubMed Central  Article  Google Scholar 

    112.
    Puckett, E. E. Variability in total project and per sample genotyping costs under varying study designs including with microsatellites or SNPs to answer conservation genetic questions. Conserv. Genet. Resour. 9, 289–304. https://doi.org/10.1007/s12686-016-0643-7 (2017).
    Article  Google Scholar 

    113.
    Hale, M. L., Burg, T. M. & Steeves, T. E. Sampling for microsatellite-based population genetic studies: 25 to 30 individuals per population is enough to accurately estimate allele frequencies. PLoS ONE 7, e45170. https://doi.org/10.1371/journal.pone.0045170 (2012).
    ADS  PubMed  PubMed Central  CAS  Article  Google Scholar 

    114.
    Luikart, G., Sherwin, W. B., Steele, B. M. & Allendorf, F. W. Usefulness of molecular markers for detecting population bottlenecks via monitoring genetic change. Mol. Ecol. 7, 963–974. https://doi.org/10.1046/j.1365-294x.1998.00414.x (1998).
    PubMed  CAS  Article  Google Scholar 

    115.
    Willing, E.-M., Dreyer, C. & van Oosterhout, C. Estimates of genetic differentiation measured by FST do not necessarily require large sample sizes when using many SNP markers. PLoS ONE 7, e42649. https://doi.org/10.1371/journal.pone.0042649 (2012).
    ADS  PubMed  PubMed Central  CAS  Article  Google Scholar 

    116.
    Nazareno, A. G., Bemmels, J. B., Dick, C. W. & Lohmann, L. G. Minimum sample sizes for population genomics: an empirical study from an Amazonian plant species. Mol. Ecol. Resour. 17, 1136–1147. https://doi.org/10.1111/1755-0998.12654 (2017).
    PubMed  CAS  Article  Google Scholar  More

  • in

    China’s researchers have valuable experiences that the world needs to hear about

    EDITORIAL
    22 September 2020

    As China prepares to take on a crucial role in the governance of global biodiversity, its researchers must be at the table.

    China’s researchers have shown that the country’s largest body of fresh water, Poyang Lake (pictured) in Jianxi province, has been drying out, in part because of the Three Gorges Dam.Credit: Fu Jianbin/Xinhua/ZUMA Wire

    Last week, the United Nations confirmed that the world has failed, again, to achieve its goals to protect nature. This grim conclusion was delivered in the fifth edition of the United Nations Global Biodiversity Outlook report.
    The report from the UN Convention on Biological Diversity reviewed progress towards 20 biodiversity targets that the convention’s participating countries set for themselves in Aichi, Japan, a decade ago (www.cbd.int/gbo).
    None of the targets, which include making progress towards the sustainable harvesting of fish, controlling the spread of invasive species and preventing the extinction of threatened wildlife, will have been achieved by the deadline at the end of this year.
    This is no time for regret or apology, but for urgency to act. Last year, an analysis by the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services revealed that some one million plant and animal species are at risk of extinction. And the wildlife charity WWF’s latest Living Planet Index, published earlier this month (see go.nature.com/32wzvdz), was similarly sobering, stating that vertebrate populations monitored between 1970 and 2017 have declined by an average of 68%.
    All nations must do more, but some of the greatest responsibility now rests on the shoulders of China: the nation, along with the leaders of the UN biodiversity convention, will jointly host the next Conference of the Parties (COP) in Kunming next year. That summit, originally scheduled for this year, is where biodiversity targets for the next decade must be set.
    As we have written before, the previous targets were destined to fail, in part because their format made progress hard to measure, and because countries did not need to report on what they were doing. This must now change. The targets, furthermore, need to be more closely aligned with the UN System of Environmental Economic Accounting, which is becoming the global standard for environmental reporting. Without these changes, the next set of biodiversity targets will almost certainly fail again.
    At the same time, China’s biodiversity scientists and policy researchers should be at the table, too, as plans for Kunming start to take shape. The country has decades of experience of studying how to — and how not to — balance economic development with controlling species and ecosystems loss. The world needs to hear these stories, in all their complexity.
    Learn from China
    The Global Biodiversity Outlook report confirms that known species are on an accelerated path to extinction, with cycad and coral species among the groups most at risk. The report shows that, although deforestation has slowed in the past decade, forests are still being splintered by agriculture, tree-felling and urban growth. Such fragmentation will further harm biodiversity and increase carbon emissions.
    Demand for food and agricultural production continue to be the main drivers of biodiversity loss. And governments are not helping. On average, they invest some US$500 billion per year in initiatives that harm the environment — eclipsing financing for biodiversity projects by a factor of 6, the report says.
    China has a set of experiences that could help the world learn valuable lessons. Its rapid economic growth lifted a generation out of poverty; however, this created a cascade of environmental problems, not least elevated pollution in the air and on land. People in China rightly questioned their leaders for underestimating — if not downplaying — the environmental and social impacts of its industrialization. Partly in response, China’s authorities have been working with researchers from China and around the world to chart a greener way forward.
    For example, national and local administrations have been devising and experimenting with environmental targets, and creating mechanisms for monitoring and reporting progress towards them — albeit with mixed success.
    China’s national biodiversity strategy includes creating what it calls ‘redlines’ — areas where human activities are restricted to protect biodiversity — across the country.
    Then there’s China’s US$6-trillion Belt and Road Initiative — a massive programme to build roads, ports and infrastructure, which will run through natural habitats across Asia, Europe and Africa. Much of this investment did not initially come with safeguards to mitigate environmental risks — but these are now being actively studied.
    And last but not least, China has a large community of researchers working to quantify, in monetary terms, the value of natural capital and ecosystem services, so that people and policymakers can more clearly understand that nature’s services to people do not come for free.
    On 30 September, heads of governments will meet at the UN for a day of talks on biodiversity, ahead of next year’s Kunming COP. Nature spoke to a number of representatives of national delegations who plan to attend this meeting, including researchers and non-governmental observers. All want the Kunming COP to succeed in bringing nations together and reaching an agreement on targets that are measurable and meaningful. But they expressed concern over the limited public engagement from China’s government about its goals or strategy for Kunming — and the relatively limited involvement of its researchers in the process so far.
    Scientists in China have been central to their country’s conservation and economic-development journey. Their collective experience on what works, and what doesn’t, can provide important learning opportunities for countries as they look to slow down and eventually reverse bio-diversity and ecosystem loss. These researchers are in the academy of sciences; in universities; in the academy of environmental planning; and in the community of Chinese and international non-governmental organizations.
    Many are also active in the China Council for Inter-national Cooperation on Environment and Development, an organization located in both Canada and China, which last week concluded a two-day conference presenting its latest research outputs. This important but little-known advisory body, now nearly three decades old, has been instrumental in connecting China’s environmental-science and environmental-policy communities with international counterparts.
    Next year will be the first time that China has hosted an international environmental meeting — similar to the 2015 Paris climate accords — where the stakes are too high to fail. It must draw on its rich diversity of talent and experience. Other nations’ researchers must be equally forthcoming with their knowledge. All sides must put aside political differences to agree on ambitious targets, ways to achieve them and methods to measure that progress.
    The best way to preserve and revive biodiversity is to acknowledge where we’ve all failed it before, to learn from that and to try again, together.

    Nature 585, 481-482 (2020)
    doi: 10.1038/d41586-020-02697-4

    Latest on:

    Biodiversity

    Government

    An essential round-up of science news, opinion and analysis, delivered to your inbox every weekday.

    Related Articles More

  • in

    Potential of aquatic weeds to improve water quality in natural waterways of the Zambezi catchment

    Study area description
    The Zambezi is Africa’s fourth largest river and the most important draining into the Indian Ocean. We focused our analysis on four sub-catchments of the Zambezi River in southern Zambia where invasions of Water Hyacinth have been reported in literature and/or which we observed from satellite imagery. These rivers are: Kafue, Chongwe, Maramba and Little Chongwe (Fig. 1). The catchments differ in size and dominant human land-use.
    Figure 1

    Location of the four catchments: Chongwe, Kafue (subcatchment between Itezhi-Tezhi Dam and Kafue Gorge), Little Chongwe and Maramba, covered in this study with urban and agricultural land cover15, wetlands and major dams in southern Zambia (www.openstreetmap.org). The four primary surface water sampling sites (pink) as well as reference sampling sites (yellow) are located through crosses. Map created using QGIS 3.4.11 (https://qgis.org).

    Full size image

    The Kafue River is a major tributary of the Zambezi and is characterized by two major hydropower dams bracketing the large Kafue Flats floodplain: Itezhi-Tezhi Dam upstream and Kafue Gorge Upper Power Station (which we subsequently refer to as simply “Kafue Gorge Dam”) downstream. Operation of these dams has greatly altered the hydrology of the lower reaches of the Kafue River leading to changes in the seasonal biochemical functioning of the Kafue Flats16,17 compared to the Barotse Plains, a reference floodplain unmodified by dams18. The lower end of the Kafue Flats has expansive coverage of sugar cane cultivation and the downstream river reach above the reservoir formed by the Kafue Gorge Dam (“Kafue Gorge Reservoir”) receives urban and industrial wastewaters from Kafue. Local reports implicate these nutrient sources as drivers of the recurrence of Water Hyacinth blooms in this area including the Kafue Gorge Reservoir. The Kafue River was subject to nutrient loading studies and intensive weed control campaigns from 1998 to 200019. Confusingly, control efforts are described as “ineffective” in the literature19, and yet the weed problem abated for a decade from 2001 to 2011. Water Hyacinth is back in recent years, following a seasonal pattern of coverage on the Kafue Gorge Reservoir surface. Mechanical control efforts have also been resumed (personal observations of the authors, February 2019).
    The Chongwe River drains one of the more densely populated catchments in Zambia, including parts of the capital, Lusaka, and several nearby townships. After tumbling down the same geographic escarpment as the Kafue, it meets the Zambezi in the Lower Zambezi National Park, approximately 55 km downstream from the confluence of the Kafue with the Zambezi River.
    The Maramba River meets the Zambezi just upstream of Victoria Falls and drains a small, but highly urbanized catchment containing most of Livingstone, a popular tourist destination for its convenient access to the falls. According to reports, overflow discharge from Livingstone’s wastewater treatment pools enters the Maramba just upstream of the confluence with the Zambezi19. During the latter stages of the dry season, low flows in the Maramba allow the river to build up a dense coverage of Water Hyacinth. These mats are seasonally flushed into the Zambezi when the rains at the end of the dry season restore the flow of the Maramba20,21.
    The Little Chongwe River meets the Kafue just 8 km upstream of the Kafue’s confluence with the Zambezi. Its small catchment drains a small portion of a large array of pivot irrigation agriculture.
    Assessment of floating vegetation cover
    To assess levels of floating vegetation, we used a combination of field surveys and analysis of satellite imagery. Floating vegetation on the Kafue Gorge Reservoir can clearly be detected with Landsat imagery. We used Google Earth Engine to group all available images in the Landsat archive in two-month intervals and extracted floating vegetation cover from 1990 to 201922.
    To assess floating vegetation cover on the Maramba, we visually inspected all available historical high-resolution imagery in Google Earth Pro (35 images from 2005 to 2019). We hand-digitized floating vegetation cover over the lower reaches of the river for which channel morphology was clearly and consistently visible and resolvable (2.3 km of linear stream reach). Since there were multiple cases of complete coverage on the Maramba River surface we interpreted complete coverage as a conservative estimate for annual Water Hyacinth biomass export in the catchment.
    Relatively few high-resolution satellite images are available in Google Earth Pro for the lower reaches of the Chongwe River and the Little Chongwe (8 each) and they do not provide complete seasonal coverage, making it impossible to apply the same process as for the Maramba. Instead we simply hand-digitized the area of visible floating vegetation from the image with the most cover (16 June, 2016) and added to this an estimate of the area of fringing floating vegetation based on four ground-based inspections of the coverage in 2018 and 2019. We estimated that fringing vegetation has a coverage of 0.5 m wide running for 1.2 km upstream of the confluence with the Zambezi. Hippo presence prevented us from surveying further upstream and we assumed no floating vegetation presence beyond this point, making our estimates conservative.
    To convert areal coverage to biomass and nutrient content, we used a synthetic mean of biomass per area from 15 studies and synthetic mean nitrogen and phosphorus content from 14 and 17 studies respectively to estimate the total pool macronutrients bound to Water Hyacinth (Supplemental Tables S1,S2). The nutrient content of Water Hyacinth varies substantially between and within studies, presumably in response to differences in nutrient availability and limitation is experiences as it grows and ages23. The Water Hyacinth in Zambian waters spends different parts of its life cycle in very different nutrient settings, ranging from urban wastewater effluent to hyperoligotrophic natural rivers, and we should therefore expect that its nutrient content should also vary substantially in space and time. The synthetic mean nutrient content from varied nutrient settings probably represent a reasonable estimate the nutrient content of Water Hyacinth in Zambia, which grows under a similarly wide range of nutrient conditions. We report uncertainty surrounding mean nutrient content using the standard error of this mean.
    Nutrient sampling
    In order to assess the potential for floating vegetation to sequester nutrients from river systems we modelled river nutrient loading for four Zambezi tributaries infested with invasive floating vegetation and estimated the amount of nutrients bound within floating vegetation biomass. We collected surface water samples from these rivers: the Maramba, the Chongwe, the Little Chongwe and the Kafue near the town of Kafue, once every three months for a year starting in March 2018. To evaluate the nutrient environment in the backwaters where we expect Water Hyacinth to be originating, we also sampled two additional sites in the larger catchments in November 2019. In the Kafue Catchment, we sampled two drainage canals conveying industrial wastewater from Kafue. In the Chongwe catchment, we sampled two points on the Gwerere River, an urban stream draining densely populated portions of Lusaka. We sampled additional points on the Kafue and Zambezi Rivers to provide reference nutrient conditions for the region’s major rivers. These sites include the Kafue River: near Hook Bridge, below the Itezhi-Tezhi Dam, near Mazabuka, near Chirundu; and the Zambezi: near Livingstone, below Kariba Dam, near Chirundu, just above Lower Zambezi National Park (for coordinates and exact sampling dates, see Supplemental Table S3).
    We passed nitrate and phosphate samples through pre-combusted, pre-weighed glass fiber filters and collected unfiltered samples for analysis of bulk nitrogen and phosphorus content. We collected all samples in triplicate and kept them cool during handling and until analysis in the laboratories at Eawag, Switzerland. We digested unfiltered water samples via autoclave with an alkaline potassium peroxidisulfate solution. We analyzed filtered water samples and digests colorimetrically using a Skalar (Breda, Netherlands) SAN++ automated flow injection analyzer following24 and standard procedure ISO 13395:1996. We note that acid digestion of unfiltered surface water is susceptible to underestimation of total phosphorus because of settling of clay particles during sample storage and sub-sampling leading to a bias against potentially phosphorus-rich particulate matter25. We assume that such underestimation in our samples is relatively minor because of low stream velocities we observed at our sampling sites, dominance of sandy rather than clay-rich soils in the region, and because we observed low C:N ratios in particulate matter (Supplemental Table S4), indicating a low-density microbial (rather than mineral) composition. Nevertheless, we refer to our P data as “digestible” rather than “total,” to allow for this potential underestimation.
    Discharge calculations, nutrient load estimation and relative importance
    We calculated average monthly discharges based on hydrographs collected by the Zambia Electricity Supply Corporation (ZESCO) between 1977 and 2017 for the Kafue Gorge Dam. For the Maramba, Chongwe and Little Chongwe Rivers, we estimated discharge by generating a catchment area: discharge curve using nearby stations from the Global Runoff Data Centre. We estimated rainfall for the Maramba catchment through monthly means from the Climate Hazards Group InfraRed Precipitation with Stations (CHIRPS) dataset26.
    In order to estimate annual river nutrient loading, we multiplied mean annual discharge by mean digestible phosphorus and mean digestible nitrogen concentrations from our surface water sampling. Our reliance on using the mean of four seasonal concentration values may not fully capture the seasonality of load and is susceptible to effects from outliers. In order to check for evidence of bias from outliers, we also calculated loading based on median concentrations and found that differences translated into a maximum decreased relative importance of plant-bound nutrients of approximately 2%. This is a minor difference compared to other sources of error, such as those stemming from uncertainty in plant biomass per area and nutrient content of plant biomass.
    We estimated the relative contributions of plant-bound versus bulk surface water nutrients to riverine nutrient export, by simply calculating the percentage of each relative to their sum. We used the product of standard errors of mean plant biomass and nutrient content values from literature to generate uncertainty envelops around our estimations for the importance of plant-bound nutrients. Here we assumed that the peak plant biomass we detected represents an annual export of intact plant material from each watershed. This assumption provides a conservative underestimate of the importance of plants because the choke points where the plants accumulate seasonally are unlikely to be 100% efficient as traps. Some plant biomass may be exported without being accounted for, but at least in the case of the Maramba and Kafue where hydrologic conditions favor clear trapping and flushing seasons, this is the most reliable approach available to estimate annual plant-bound nutrient export.
    Landcover analysis
    The river catchments were extracted from the HydroSheds global database27. For the Kafue River, we considered only the catchment below the Itezhi-Tezhi Dam, as we expected the dam to interfere with the nutrient transport from further upstream28.
    To calculate the area of urban and agricultural land within each catchment we used fractional land cover maps for 2015 from the Copernicus Global Land Service29. Using the raster package30 in R, we extracted pixel values within each catchment polygon and then calculated the area for all pixels with a fraction of more than 50% for the respective class.
    All analyses and figures were completed using R version 3.6.131. More

  • in

    Forensic tracers of exposure to produced water in freshwater mussels: a preliminary assessment of Ba, Sr, and cyclic hydrocarbons

    1.
    Clark, C. & Veil, J. Produced water volumes and management practices in the United States. Argonne Natl. Lab. Rep. https://doi.org/10.2172/1007397 (2009).
    Article  Google Scholar 
    2.
    Dolan, F. C., Cath, T. Y. & Hogue, T. S. Assessing the feasibility of using produced water for irrigation in Colorado. Sci. Total Environ. 640–641, 619–628 (2018).
    ADS  PubMed  Google Scholar 

    3.
    McDevitt, B. et al. Isotopic and element ratios fingerprint salinization impact from beneficial use of oil and gas produced water in the Western U.S. Sci. Total Environ. 716, 137006 (2020).
    ADS  CAS  PubMed  Google Scholar 

    4.
    McLaughlin, M. C., Borch, T., McDevitt, B., Warner, N. R. & Blotevogel, J. Water quality assessment downstream of oil and gas produced water discharges intended for beneficial reuse in arid regions. Sci. Total Environ. 713, 136607 (2020).
    ADS  CAS  PubMed  Google Scholar 

    5.
    Kahrilas, G. A., Blotevogel, J., Stewart, P. S. & Borch, T. Biocides in hydraulic fracturing fluids: a critical review of their usage, mobility, degradation, and toxicity. Environ. Sci. Technol. 49, 16–32 (2015).
    ADS  CAS  PubMed  Google Scholar 

    6.
    Haluszczak, L. O., Rose, A. W. & Kump, L. R. Geochemical evaluation of flowback brine from Marcellus gas wells in Pennsylvania, USA. Appl. Geochem. 28, 55–61 (2013).
    CAS  Google Scholar 

    7.
    Lester, Y. et al. Characterization of hydraulic fracturing flowback water in Colorado: Implications for water treatment. Sci. Total Environ. 512–513, 637–644 (2015).
    ADS  PubMed  Google Scholar 

    8.
    Michael Thurman, E., Ferrer, I., Blotevogel, J. & Borch, T. Analysis of hydraulic fracturing flowback and produced waters using accurate mass: identification of ethoxylated surfactants. Anal. Chem. 86, 9653–9661 (2014).
    PubMed  Google Scholar 

    9.
    Hoelzer, K. et al. Indications of transformation products from hydraulic fracturing additives in shale-gas wastewater. Environ. Sci. Technol. 50, 8036–8048 (2016).
    ADS  CAS  PubMed  Google Scholar 

    10.
    Piotrowski, P. K. et al. Elucidating environmental fingerprinting mechanisms of unconventional gas development through hydrocarbon analysis. Anal. Chem. 90, 5466–5473 (2018).
    CAS  PubMed  Google Scholar 

    11.
    Piotrowski, P. K. et al. Non-Targeted chemical characterization of a Marcellus shale gas well through GC × GC with scripting algorithms and high-resolution time-of-flight mass spectrometry. Fuel 215, 363–369 (2018).
    CAS  Google Scholar 

    12.
    He, Y. et al. Chemical and toxicological characterizations of hydraulic fracturing flowback and produced water. Water Res. 114, 78–87 (2017).
    CAS  PubMed  Google Scholar 

    13.
    Llewellyn, G. T. et al. Evaluating a groundwater supply contamination incident attributed to Marcellus Shale gas development. Proc. Natl. Acad. Sci. 112, 6325 (2015).
    ADS  CAS  PubMed  Google Scholar 

    14.
    Drollette, B. D. et al. Elevated levels of diesel range organic compounds in groundwater near Marcellus gas operations are derived from surface activities. Proc. Natl. Acad. Sci. USA 112, 13184–13189 (2015).
    ADS  CAS  PubMed  Google Scholar 

    15.
    Cozzarelli, I. M. et al. Environmental signatures and effects of an oil and gas wastewater spill in the Williston Basin, North Dakota. s 579, 1781–1793 (2016).
    Google Scholar 

    16.
    Digiulio, D. C. & Jackson, R. B. Impact to underground sources of drinking water and domestic wells from production well stimulation and completion practices in the Pavillion, Wyoming, Field. Environ. Sci. Technol. 50, 4524–4536 (2016).
    ADS  CAS  PubMed  Google Scholar 

    17.
    Gross, S. A. et al. Analysis of BTEX groundwater concentrations from surface spills associated with hydraulic fracturing operations. J. Air Waste Manag. Assoc. 63, 424–432 (2013).
    CAS  PubMed  Google Scholar 

    18.
    Hildenbrand, Z. L. et al. Temporal variation in groundwater quality in the Permian Basin of Texas, a region of increasing unconventional oil and gas development. Sci. Total Environ. 562, 906–913 (2016).
    ADS  CAS  PubMed  Google Scholar 

    19.
    Kassotis, C. D. et al. Endocrine disrupting activities of surface water associated with a West Virginia oil and gas industry wastewater disposal site. Sci. Total Environ. 557–558, 901–910 (2016).
    ADS  PubMed  Google Scholar 

    20.
    Burgos, W. D. et al. Watershed-scale impacts from surface water disposal of oil and gas wastewater in Western Pennsylvania. Environ. Sci. Technol. 51, 8851–8860 (2017).
    ADS  CAS  PubMed  Google Scholar 

    21.
    Getzinger, G. J. et al. Natural gas residual fluids: sources, endpoints, and organic chemical composition after centralized waste treatment in Pennsylvania. Environ. Sci. Technol. 49, 8347–8355 (2015).
    ADS  CAS  PubMed  Google Scholar 

    22.
    Blewett, T. A. et al. Sublethal and reproductive effects of acute and chronic exposure to flowback and produced water from hydraulic fracturing on the water Flea Daphnia magna. Environ. Sci. Technol. 51, 3032–3039 (2017).
    ADS  CAS  PubMed  Google Scholar 

    23.
    Blewett, T. A., Delompré, P. L. M., Glover, C. N. & Goss, G. G. Physical immobility as a sensitive indicator of hydraulic fracturing fluid toxicity towards Daphnia magna. Sci. Total Environ. 635, 639–643 (2018).
    ADS  CAS  PubMed  Google Scholar 

    24.
    He, Y. et al. Developmental toxicity of the organic fraction from hydraulic fracturing Flowback and produced waters to early life stages of Zebrafish (Danio rerio). Environ. Sci. Technol. 52, 3820–3830 (2018).
    ADS  CAS  PubMed  Google Scholar 

    25.
    He, Y. et al. Effects on biotransformation, oxidative stress, and endocrine disruption in rainbow trout (Oncorhynchus mykiss) exposed to hydraulic fracturing flowback and produced water. Environ. Sci. Technol. 51, 940–947 (2017).
    ADS  CAS  PubMed  Google Scholar 

    26.
    Blewett, T. A., Weinrauch, A. M., Delompré, P. L. M. & Goss, G. G. The effect of hydraulic flowback and produced water on gill morphology, oxidative stress and antioxidant response in rainbow trout (Oncorhynchus mykiss). Sci. Rep. 7, 46582 (2017).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    27.
    Tasker, T. L. et al. Environmental and human health impacts of spreading oil and gas wastewater on roads. Environ. Sci. Technol. 52, 7081–7091 (2018).
    ADS  CAS  PubMed  Google Scholar 

    28.
    McDevitt, B. et al. Emerging investigator series: radium accumulation in carbonate river sediments at oil and gas produced water discharges: implications for beneficial use as disposal management. Environ. Sci. Process. Impacts 21, 324–338 (2019).
    MathSciNet  CAS  PubMed  Google Scholar 

    29.
    McLaughlin, M. C. et al. Mutagenicity assessment downstream of oil and gas produced water discharges intended for agricultural beneficial reuse. Sci. Total Environ. 715, 136944 (2020).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    30.
    Folkerts, E. J., Blewett, T. A., He, Y. & Goss, G. G. Alterations to Juvenile Zebrafish (Danio rerio) swim performance after acute embryonic exposure to sub-lethal exposures of hydraulic fracturing flowback and produced water. Aquat. Toxicol. 193, 50–59 (2017).
    CAS  PubMed  Google Scholar 

    31.
    Folkerts, E. J., Blewett, T. A., He, Y. & Goss, G. G. Cardio-respirometry disruption in zebrafish (Danio rerio) embryos exposed to hydraulic fracturing flowback and produced water. Environ. Pollut. 231, 1477–1487 (2017).
    CAS  PubMed  Google Scholar 

    32.
    Delompré, P. L. M. et al. The osmotic effect of hyper-saline hydraulic fracturing fluid on rainbow trout, Oncorhynchus mykiss. s 211, 1–10 (2019).
    Google Scholar 

    33.
    Wang, N. et al. Acute toxicity of sodium chloride and potassium chloride to a unionid mussel (Lampsilis siliquoidea) in water exposures. Environ. Toxicol. Chem. 37, 3041–3049 (2018).
    CAS  PubMed  PubMed Central  Google Scholar 

    34.
    Wang, N., Kunz, J. L., Cleveland, D., Steevens, J. A. & Cozzarelli, I. M. Biological effects of elevated major ions in surface water contaminated by a produced water from oil production. Arch. Environ. Contam. Toxicol. 76, 670–677 (2019).
    CAS  PubMed  Google Scholar 

    35.
    Wang, N. et al. Evaluation of chronic toxicity of sodium chloride or potassium chloride to a unionid mussel (Lampsilis siliquoidea) in water exposures using standard and refined toxicity testing methods. Environ. Toxicol. Chem. 37, 3050–3062 (2018).
    CAS  PubMed  Google Scholar 

    36.
    Patnode, K. A., Hittle, E., Anderson, R. M., Zimmerman, L. & Fulton, J. W. Effects of high salinity wastewater discharges on unionid mussels in the allegheny river Pennsylvania. J. Fish Wildl. Manag. 6, 55–70 (2015).
    Google Scholar 

    37.
    Geeza, T. J., Gillikin, D. P., McDevitt, B., Van Sice, K. & Warner, N. R. Accumulation of marcellus formation oil and gas wastewater metals in freshwater mussel shells. Environ. Sci. Technol. 52, 10883–10892 (2018).
    ADS  CAS  PubMed  Google Scholar 

    38.
    De los Ríos, A. et al. Assessment of the effects of discontinuous sources of contamination through biomarker analyses on caged mussels. Sci. Total Environ. https://doi.org/10.1016/j.scitotenv.2018.03.297 (2018).
    Article  PubMed  Google Scholar 

    39.
    Pilote, M., André, C., Turcotte, P., Gagné, F. & Gagnon, C. Metal bioaccumulation and biomarkers of effects in caged mussels exposed in the Athabasca oil sands area. Sci. Total Environ. https://doi.org/10.1016/j.scitotenv.2017.08.023 (2018).
    Article  PubMed  Google Scholar 

    40.
    Bonnefille, B., Arpin-Pont, L., Gomez, E., Fenet, H. & Courant, F. Metabolic profiling identification of metabolites formed in Mediterranean mussels (Mytilus galloprovincialis) after diclofenac exposure. Sci. Total Environ. 583, 257–268 (2017).
    ADS  CAS  PubMed  Google Scholar 

    41.
    Spring, S. NOAA Technical Memorandum NOS ORCA 71 National Status and Trends Program for Marine Environmental Quality Sampling and Analytical Methods of the National Status and Trends Program National Benthic Surveillance and Mussel Watch Projects 1984–1992 Volume I Overview and Summary of Methods noaa NATIONAL OCEANIC AND ATMOSPHERIC ADMINISTRATION Coastal Monitoring and Bioeffects Assessment Division Office of Ocean Resources Conservation and Assessment National Ocean Service. (1993).

    42.
    Lydeard, C. et al. The global decline of Nonmarine Mollusks. Bioscience 54, 321–330 (2004).
    Google Scholar 

    43.
    Strayer, D. L. et al. Changing perspectives on pearly mussels, North America’s most imperiled animals. Bioscience 54, 429–439 (2004).
    Google Scholar 

    44.
    Lopes-Lima, M. et al. Conservation status of freshwater mussels in Europe: state of the art and future challenges. Biol. Rev. 92, 572–607 (2017).
    PubMed  Google Scholar 

    45.
    Shulkin, V. M., Presley, B. J. & Kavun, V. I. Metal concentrations in mussel Crenomytilus grayanus and oyster Crassostrea gigas in relation to contamination of ambient sediments. Environ. Int. https://doi.org/10.1016/S0160-4120(03)00004-7 (2003).
    Article  PubMed  Google Scholar 

    46.
    Ishikawa, Y., Kagaya, H. & Saga, K. Biomagnification of 7Be, 234Th, and 228Ra in marine organisms near the northern Pacific coast of Japan. J. Environ. Radioact. https://doi.org/10.1016/j.jenvrad.2004.03.021 (2004).
    Article  PubMed  Google Scholar 

    47.
    Brenner, M., Smoak, J. M., Leeper, D. A., Streubert, M. & Baker, S. M. Radium-226 accumulation in Florida freshwater mussels. Limnol. Oceanogr. 52, 1614–1623 (2007).
    ADS  CAS  Google Scholar 

    48.
    Jeffree, R. A., Markich, S. J. & Brown, P. L. Comparative accumulation of alkaline-earth metals by two freshwater mussel species from the Nepean River, Australia: consistencies and a resolved paradox. Aust. J. Mar. Freshw. Res. 44, 609–634 (1993).
    CAS  Google Scholar 

    49.
    Markich, S. J., Brown, P. L. & Jeffree, R. A. Divalent metal accumulation in freshwater bivalves: an inverse relationship with metal phosphate solubility. Sci. Total Environ. https://doi.org/10.1016/S0048-9697(00)00721-X (2001).
    Article  PubMed  Google Scholar 

    50.
    Markich, S. J. & Jeffree, R. A. Absorption of divalent trace metals as analogues of calcium by Australian freshwater bivalves: an explanation of how water hardness reduces metal toxicity. Aquat. Toxicol. https://doi.org/10.1016/0166-445X(94)90072-8 (1994).
    Article  Google Scholar 

    51.
    Fritz, L., Ragone, L., Lutz, R. & Swapp, S. Biomineralization of barite in the shell of the freshwater asiatic clam Corbicula fluminea (Mollusca: Bivalvia). Limnol. Oceanogr. 35, 756–762 (1990).
    ADS  CAS  Google Scholar 

    52.
    Ricciardi, A., Neves, R. J. & Rasmussen, J. B. Impending extinctions of North American freshwater mussels (Unionoida) following the zebra mussel (Dreissena polymorpha) invasion. J. Anim. Ecol. 67, 613–619 (1998).
    Google Scholar 

    53.
    Ricciardi, A. & Rasmussen, J. B. Extinction rates of North American freshwater fauna. Conserv. Biol. 13, 1220–1222 (1999).
    Google Scholar 

    54.
    Vaughn, C. C. & Taylor, C. M. Impoundments and the decline of freshwater mussels: a case study of an extinction gradient. Conserv. Biol. 13, 912–920 (1999).
    Google Scholar 

    55.
    Weggler, B. A. et al. Untargeted identification of wood type-specific markers in particulate matter from wood combustion. Environ. Sci. Technol. 50, 10073–10081 (2016).
    ADS  CAS  PubMed  Google Scholar 

    56
    Colborn, T. & Smolen, M. J. Epidemiological analysis of persistent organochlorine contaminants in cetaceans. Rev. Environ. Contam. Toxicol. https://doi.org/10.1007/978-1-4613-8478-6_4 (1996).
    Article  PubMed  Google Scholar 

    57.
    Halbrook, R. S., Kirkpatrick, R. L., Scanlon, P. F., Vaughan, M. R. & Veit, H. P. Environmental Contamination a n d Toxicology Muskrat Populations in Virginia’s Elizabeth River: Physiological Condition and Accumulation of Environmental Contaminants. Arch. Environ. Contain. Toxicol 25, (ARCHIVES OF, 1993).

    58.
    Birdsall, K., Kukor, J. J. & Cheney, M. A. Uptake of polycyclic aromatic hydrocarbon compounds by the gills of the bivalve mollusk Elliptio complanata. Environ. Toxicol. Chem. 20, 309–316 (2001).
    CAS  PubMed  Google Scholar 

    59.
    Thorsen, W. A. et al. Elimination rate constants of 46 polycyclic aromatic hydrocarbons in the unionid mussel, Elliptio complanata. Arch. Environ. Contam. Toxicol. https://doi.org/10.1007/s00244-004-3186-y (2004).
    Article  PubMed  Google Scholar 

    60.
    Gewurtz, S. B., Drouillard, K. G., Lazar, R. & Haffner, G. D. Quantitative biomonitoring of PAHs using the barnes mussel (Elliptio complanata). Arch. Environ. Contam. Toxicol. https://doi.org/10.1007/s00244-002-1153-z (2002).
    Article  PubMed  Google Scholar 

    61.
    Gewurtz, S. G., Lazar, R. & Haffner, G. D. Biomonitoring of bioavailable PAH and PCB water concentrations in the Detroit River using the freshwater mussel Elliptio complanata. J. Great Lakes Res. 29, 242–255 (2003).
    CAS  Google Scholar 

    62.
    Metcalfe, J. L. & Charlton, M. N. Freshwater mussels as biomonitors for organic industrial contaminants and pesticides in the St Lawrence River. Sci. Total Environ. 98, 595–615 (1990).
    ADS  Google Scholar 

    63
    Drouillard, K. G. et al. Quantitative biomonitoring in the Detroit River using Elliptio complanata: verification of steady state correction factors and temporal trends of PCBs in water between 1998 and 2015. Bull. Environ. Contam. Toxicol. 97, 757–762 (2016).
    CAS  PubMed  Google Scholar 

    64.
    Drouillard, K. G., Chan, S., O’Rourke, S., Douglas Haffner, G. & Letcher, R. J. Elimination of 10 polybrominated diphenyl ether (PBDE) congeners and selected polychlorinated biphenyls (PCBs) from the freshwater mussel, Elliptio complanata. Chemosphere 69, 362–370 (2007).
    ADS  CAS  PubMed  Google Scholar 

    65.
    O’Rourke, S., Drouillard, K. G. & Haffner, G. D. Determination of laboratory and field elimination rates of polychlorinated biphenyls (PCBs) in the freshwater mussel, Elliptio complanata. Environ. Contam. Toxicol. 47, 74–83 (2004).
    Google Scholar 

    66.
    Xu, M. et al. Quantitative structure-activity relationship for the depuration rate constants of polychlorinated biphenyls in the freshwater mussel, Elliptio complanata. J. Environ. Sci. Heal. Part B 44, 278–283 (2009).
    CAS  Google Scholar 

    67.
    Sabik, H., Gagné, F., Blaise, C., Marcogliese, D. J. & Jeannot, R. Occurrence of alkylphenol polyethoxylates in the St. Lawrence River and their bioconcentration by mussels (Elliptio complanata). Chemosphere 51, 349–356 (2003).
    ADS  CAS  PubMed  Google Scholar 

    68.
    Donkin, P., Smith, E. L. & Rowland, S. J. Toxic effects of unresolved complex mixtures of aromatic hydrocarhons accumulated by mussels, Mytilus edulis, from contaminated field sites. Environ. Sci. Technol. 37, 4825–4830 (2003).
    ADS  CAS  PubMed  Google Scholar 

    69.
    Booth, A. M. et al. Unresolved complex mixtures of aromatic hydrocarbons: thousands of overlooked persistent, bioaccumulative, and toxic contaminants in mussels. Environ. Sci. Technol. 41, 457–464 (2007).
    ADS  CAS  PubMed  Google Scholar 

    70.
    Gillikin, D. P. et al. Barium uptake into the shells of the common mussel (Mytilus edulis) and the potential for estuarine paleo-chemistry reconstruction. Geochim. Cosmochim. Acta https://doi.org/10.1016/j.gca.2005.09.015 (2006).
    Article  Google Scholar 

    71.
    Geeza, T. J. et al. Controls on magnesium, manganese, strontium, and barium concentrations recorded in freshwater mussel shells from Ohio. Chem. Geol. https://doi.org/10.1016/j.chemgeo.2018.01.001 (2019).
    Article  Google Scholar 

    72.
    Abualfaraj, N., Gurian, P. L. & Olson, M. S. Characterization of marcellus shale flowback water. Environ. Eng. Sci. 31, 514–524 (2014).
    CAS  Google Scholar 

    73.
    Chapman, E. C. et al. Geochemical and strontium isotope characterization of produced waters from marcellus shale natural gas extraction. Environ. Sci. Technol. 46, 3545–3553 (2012).
    ADS  CAS  PubMed  Google Scholar 

    74.
    Poulain, C. et al. An evaluation of Mg/Ca, Sr/Ca, and Ba/Ca ratios as environmental proxies in aragonite bivalve shells. Chem. Geol. 396, 42–50 (2015).
    ADS  CAS  Google Scholar 

    75.
    Lorens, R. B. & Bender, M. L. The impact of solution chemistry on Mytilus edulis calcite and aragonite. Geochim. Cosmochim. Acta 44, 1265–1278 (1980).
    ADS  CAS  Google Scholar 

    76.
    Wheeler, A. Mechanisms of molluscan shell formation. (1992).

    77.
    Wilbur, K. M. The Mollusca (Academic Press, Cambridge, 1983).
    Google Scholar 

    78.
    Crenshaw, M. A. The inorganic composition of molluscan extrapallial fluid. Biol. Bull. 143, 6–512 (1972).
    Google Scholar 

    79.
    Mount, A. S., Wheeler, A. P., Paradkar, R. P. & Snider, D. Hemocyte-mediated shell mineralization in the Eastern Oyster. Science 304, 297–300 (2004).
    ADS  CAS  PubMed  Google Scholar 

    80.
    Nair, P. S. & Robinson, W. E. Calcium speciation and exchange between blood and extrapallial fluid of the quahog Mercenaria mercenaria (L.). Biol. Bull. 195, 43–51 (1998).
    CAS  PubMed  Google Scholar 

    81.
    Kelemen, Z., Gillikin, D. P. & Bouillon, S. Relationship between river water chemistry and shell chemistry of two tropical African freshwater bivalve species. Chem. Geol. 526, 130–141 (2019).
    ADS  CAS  Google Scholar 

    82.
    Canadian Council of Ministers of the Environment. Canadian water quality guidelines for the protection of aquatic life. Water Quality Index. (2001).

    83.
    Todd, A. K. & Kaltenecker, M. G. Warm season chloride concentrations in stream habitats of freshwater mussel species at risk. Environ. Pollut. 171, 199–206 (2012).
    CAS  PubMed  Google Scholar 

    84.
    O’Neil, D. D. & Gillikin, D. P. Do freshwater mussel shells record road-salt pollution?. Sci. Rep. 4, 1–6 (2014).
    Google Scholar 

    85.
    Ganoe, L. S., Brown, J. D., Yabsley, M. J., Lovallo, M. J. & Walter, W. D. A review of pathogens, diseases, and contaminants of muskrats (Ondatra zibethicus) in North America. Frontiers Veterin. Sci. 7, 233 (2020).
    Google Scholar 

    86.
    Warner, N. R., Christie, C. A., Jackson, R. B. & Vengosh, A. Impacts of shale gas wastewater disposal on water quality in Western Pennsylvania. Environ. Sci. Technol. 47, 11849–11857 (2013).
    ADS  CAS  PubMed  Google Scholar 

    87.
    United States Environmental Protection Agency. Method 3510C (SW-846): Separatory Funnel Liquid-Liquid Extraction, Revision 3. (1996).

    88.
    United States Environmental Protection Agency. Method 3005A (SW-846): Acid Digestion of Waters for Total Recoverable or Dissolved Metals for Analysis by FLAA or ICP Spectroscopy, Revision 3. (1992).

    89.
    Chong, J. et al. MetaboAnalyst 4.0: towards more transparent and integrative metabolomics analysis. Web Serv. 46, W486–W494 (2018).
    CAS  Google Scholar  More

  • in

    A database of chlorophyll and water chemistry in freshwater lakes

    1.
    Beeton, A. M. Large freshwater lakes: present state, trends, and future. Environ Conserv. 29, 21–38 (2002).
    CAS  Google Scholar 
    2.
    Shiklomanov, I. A. Water in Crisis: A Guide to the World’s Fresh Water Resources (Oxford Univ. Press, 1993).

    3.
    McMichael, A. J., Woodruff, R. E. & Hales, S. Climate change and human health: present and future risks. The Lancet 367, 859–869 (2006).
    Google Scholar 

    4.
    Meyer, M. F., Labou, S. G., Cramer, A. N., Brousil, M. R. & Luff, B. T. The global lake area, climate, and population dataset. Sci. Data 7, 1–12 (2020).
    Google Scholar 

    5.
    Wrona, F. J. et al. Climate change effects on aquatic biota, ecosystem structure and function. Ambio 35, 359–369 (2006).
    CAS  PubMed  Google Scholar 

    6.
    Adrian, R. et al. Lakes as sentinels of climate change. Limnol. Oceanogr. 54, 2283–2297 (2009).
    ADS  PubMed  PubMed Central  Google Scholar 

    7.
    Nürnberg, G. K. & Shaw, M. Productivity of clear and humic lakes: nutrients, phytoplankton, bacteria. Hydrobiologia 382, 97–112 (1998).
    Google Scholar 

    8.
    Makri, S., Lami, A., Lods-Crozet, B. & Loizeau, J. L. Reconstruction of trophic state shifts over the past 90 years in a eutrophicated lake in western Switzerland, inferred from the sedimentary record of photosynthetic pigments. J. Paleolimnol. 61, 129–145 (2019).
    ADS  Google Scholar 

    9.
    Håkanson, L. & Boulion, V. V. Regularities in primary production, Secchi depth and fish yield and a new system to define trophic and humic state indices for lake ecosystems. Int. Rev. Hydrobiol. 86, 23–62 (2001).
    Google Scholar 

    10.
    Carlson, R. E. A trophic state index for lakes. Limnol. Oceanogr. 22, 361–369 (1977).
    ADS  CAS  Google Scholar 

    11.
    Sterner, R. W. In situ-measured primary production in Lake Superior. J. Great Lakes Res. 36, 139–149 (2010).
    Google Scholar 

    12.
    Li, X., Sha, J. & Wang, Z. L. Chlorophyll-a prediction of lakes with different water quality patterns in China based on hybrid neural networks. Water 9, 524 (2017).
    ADS  Google Scholar 

    13.
    Vollenweider, R. & Kerekes, J. Eutrophication of Waters: Monitoring, Assessment and Control. OECD (1982).

    14.
    Bennion, D. H., Warner, D. M., Esselman, P. C., Hobson, B. & Kieft, B. A comparison of chlorophyll a values obtained from an autonomous underwater vehicle to satellite-based measures for Lake Michigan. J. Great Lakes Res. 45, 726–734 (2019).
    CAS  Google Scholar 

    15.
    Elser, J. J. et al. Global analysis of nitrogen and phosphorus limitation of primary producers in freshwater, marine and terrestrial ecosystems. Ecol. Lett. 10, 1135–1142 (2007).
    PubMed  Google Scholar 

    16.
    Hall, R. I., Leavitt, P. R., Quinlan, R., Dixit, A. S. & Smol, J. P. Effects of agriculture, urbanization, and climate on water quality in the northern Great Plains. Limnol. Oceanogr. 44, 739–756 (1999).
    ADS  CAS  Google Scholar 

    17.
    Bennett, E. M., Carpenter, S. R. & Caraco, N. F. Human impact on erodable phosphorus and eutrophication: a global perspective: increasing accumulation of phosphorus in soil threatens rivers, lakes, and coastal oceans with eutrophication. AIBS Bulletin 51, 227–234 (2001).
    Google Scholar 

    18.
    Williamson, C. E., Dodds, W., Kratz, T. K. & Palmer, M. A. Lakes and streams as sentinels of environmental change in terrestrial and atmospheric processes. Front. Ecol. Environ. 6, 247–254 (2008).
    Google Scholar 

    19.
    Carpenter, S. R. et al. Nonpoint pollution of surface waters with phosphorus and nitrogen. Ecol. Appl. 8, 559–568 (1998).
    Google Scholar 

    20.
    Williamson, C. E., Saros, J. E., Vincent, W. F. & Smol, J. P. Lakes and reservoirs as sentinels, integrators, and regulators of climate change. Limnol. Oceanogr. 54, 2273–2282 (2009).
    ADS  Google Scholar 

    21.
    Li, L., Li, L., Shi, K., Li, Z. & Song, K. A semi-analytical algorithm for remote estimation of phycocyanin in inland waters. Sci. Total Environ. 435, 141–150 (2012).
    ADS  PubMed  Google Scholar 

    22.
    Odermatt, D., Danne, O., Philipson, P. & Brockmann, C. Diversity II water quality parameters from ENVISAT (2002-2012): a new global information source for lakes. Earth Syst. Sci. Data 10, 1527–1549 (2018).
    ADS  Google Scholar 

    23.
    Palmer, S. C., Kutser, T. & Hunter, P. D. Remote sensing of inland waters: Challenges, progress and future directions. Remote Sens. Environ. 157, 1–8 (2015).
    ADS  Google Scholar 

    24.
    Salama, M. S. & Verhoef, W. Two-stream remote sensing model for water quality mapping: 2SeaColor. Remote Sens. Environ. 157, 111–122 (2015).
    ADS  Google Scholar 

    25.
    Soranno, P. A. et al. LAGOS-NE: A multi-scaled geospatial and temporal database of lake ecological context and water quality for thousands of U.S. lakes. Gigascience 6, 1–22 (2017).
    PubMed  PubMed Central  Google Scholar 

    26.
    Zeng, L. H. & Li, D. L. Development of in situ sensors for chlorophyll concentration measurement. J. Sens. 2015, 1–16 (2015).
    Google Scholar 

    27.
    Shimaraeva, S. V., Pislegina, E. V., Krashchuk, L. S., Shchapov, K. S. & Silow, E. A. Dynamics of chlorophyll a concentration in the South Baikal pelagic during the direct temperature stratification period. Inland Water Biol. 10, 59–63 (2017).
    Google Scholar 

    28.
    Eaton, A. D., & Franson, M. A. H. Standard Methods for the Examination of Water and Wastewater. American Public Health Association, American Water Works Association, Water Environment Federation, Washington, Denver, Alexandria (2005).

    29.
    Torremorell, A., del Carmen Diéguez, M., Queimaliños, C., Izaguirre, I. & Zagarese, H. E. Phytoplankton limitation in Patagonian and Pampean shallow lakes: effect of phosphorus and light. Hydrobiologia 816, 91–105 (2018).
    CAS  Google Scholar 

    30.
    Messager, M. L., Lehner, B., Grill, G., Nedeva, I. & Schmitt, O. Estimating the volume and age of water stored in global lakes using a geo-statistical approach. Nat. Commun. 7, 13603 (2016).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    31.
    Filazzola, A. et al. A global database of chlorophyll and water chemistry in freshwater lakes. KNB Data Repository https://doi.org/10.5063/F1RV0M1S (2020).

    32.
    Marselina, M. & Burhanudin, M. Trophic status assessment of Saguling Reservoir, Upper Citarum Basin, Indonesia. Air, Soil and Water Res. 10, 1–8 (2017).
    Google Scholar 

    33.
    R Development Core Team. R: A language and environment for statistical computing (2019).

    34.
    Wickham, H. & Henry, L. tidyr: Tidy Messy Data. R package version 1.0.0. https://CRAN.R-project.org/package=tidyr (2019).

    35.
    Wickham, H., François, R., Henry, L., & Müller, K. dplyr: A Grammar of Data Manipulation. R package version 0.8.3. https://CRAN.R-project.org/package=dplyr (2019).

    36.
    Wickham, H. ggplot2: Elegant Graphics for Data Analysis. (Springer-Verlag, New York, 2016).
    Google Scholar 

    37.
    Verpoorter, C., Kutser, T., Seekell, D. A. & Tranvik, L. J. A global inventory of lakes based on high‐resolution satellite imagery. Geophys. Res. Lett. 41, 6396–6402 (2014).
    ADS  Google Scholar 

    38.
    Filazzola, A. afilazzola/ChlorophyllDataPaper: Initial-Release. Zenodo https://doi.org/10.5281/zenodo.3968735 (2020).

    39.
    Hampton, S. E. et al. Ecology under lake ice. Ecol. Lett. 20, 98–111 (2017).
    PubMed  Google Scholar 

    40.
    Karatayev, Vadim, A. et al. Eutrophication and Dreissena invasion as drivers of biodiversity: A century of change in the mollusc community of Oneida Lake. PloS One 9 (2014).

    41.
    Richardson, D. C. et al. Transparency, geomorphology and mixing regime explain variability in trends in lake temperature and stratification across northeastern North America (1975–2014). Water 9, 442 (2017).
    Google Scholar 

    42.
    Mantzouki, E. et al. The European Multi Lake Survey (EMLS) dataset of physical, chemical, algal pigments and cyanotoxin parameters 2015. Environmental Data Initiative (2018).

    43.
    Pollard, AminaI., Hampton, StephanieE. & Leech, DinaM. The Promise and Potential of Continental‐Scale Limnology Using the US Environmental Protection Agency’s National Lakes. Assessment. Limnol.Oceanogr. Bull. 27, 36–41 (2018).
    Google Scholar 

    44.
    Burnett, L., Moorhead, D., Hawes, I. & Howard-Williams, C. Environmental factors associated with deep chlorophyll maxima in Dry Valley lakes, South Victoria Land, Antarctica. Arct. Antarct. Alp. Res. 38, 179–189 (2006).
    Google Scholar 

    45.
    Takamura, N. & Nakagawa, M. The densities of bacteria, picophytoplankton, heterotrophic nanoflagellates and ciliates in Lake Kasumigaura (Japan) monitored monthly since 1996. Ecol. Res. 27, 839 (2012).
    Google Scholar 

    46.
    Gries, C., Gahler, M. R., Hanson, P. C., Kratz, T. K. & Stanley, E. H. Information management at the North Temperate Lakes Long-term Ecological Research site—Successful support of research in a large, diverse, and long running project. Ecol. Inform. 36, 201–208 (2016).
    Google Scholar  More