More stories

  • in

    Multifaceted characteristics of dryland aridity changes in a warming world

    1.
    Reynolds, J. F. et al. Global desertification: building a science for dryland development. Science 316, 847–851 (2007).
    Article  Google Scholar 
    2.
    Huang, J. et al. Dryland climate change: recent progress and challenges. Rev. Geophys. 55, 719–778 (2017).
    Article  Google Scholar 

    3.
    Intergovernmental Panel on Climate Change (IPCC). Climate Change and Land: An IPCC Special Report on Climate Change, Desertification, Land Degradation, Sustainable Land Management, Food Security, and Greenhouse Gas Fluxes in Terrestrial Ecosystems (eds Akhtar-Schuster, M., Driouech, F. & Sankaran, M.) Ch. 3 (IPCC, Cambridge Univ. Press, 2019).

    4.
    Prăvălie, R. Drylands extent and environmental issues. A global approach. Earth Sci. Rev. 161, 259–278 (2016).
    Article  Google Scholar 

    5.
    D’Odorico, P., Bhattachan, A., Davis, K. F., Ravi, S. & Runyan, C. W. Global desertification: drivers and feedbacks. Adv. Water Resour. 51, 326–344 (2013).
    Article  Google Scholar 

    6.
    Ahlström, A. et al. The dominant role of semi-arid ecosystems in the trend and variability of the land CO2 sink. Science 348, 895–899 (2015). Highlights the critical role of drylands in the global carbon budget by demonstrating that semi-arid ecosystems dominate the inter-annual variability and the increasing trend of global terrestrial carbon sink.
    Article  Google Scholar 

    7.
    Maestre, F. T. et al. Plant species richness and ecosystem multifunctionality in global drylands. Science 335, 214–218 (2012).
    Article  Google Scholar 

    8.
    Millennium Ecosystem Assessment. Ecosystems and Human Well-Being: Desertification Synthesis (World Resources Institute, 2005).

    9.
    El-Beltagy, A. & Madkour, M. Impact of climate change on arid lands agriculture. Agric. Food Secur. 1, 3 (2012).
    Article  Google Scholar 

    10.
    Huang, J., Yu, H., Guan, X., Wang, G. & Guo, R. Accelerated dryland expansion under climate change. Nat. Clim. Change 6, 166–171 (2016).
    Article  Google Scholar 

    11.
    Cook, K. H. & Vizy, E. K. Detection and analysis of an amplified warming of the Sahara Desert. J. Clim. 28, 6560–6580 (2015).
    Article  Google Scholar 

    12.
    Zhou, L., Chen, H. & Dai, Y. Stronger warming amplification over drier ecoregions observed since 1979. Environ. Res. Lett. 10, 064012 (2015).
    Article  Google Scholar 

    13.
    Fu, B. et al. The Global-DEP conceptual framework — research on dryland ecosystems to promote sustainability. Curr. Opin. Environ. Sustain. 48, 17–28 (2021). Proposes a conceptual framework that aims to facilitate actionable pathways towards sustainable development of global dryland socio-ecological systems.
    Article  Google Scholar 

    14.
    Larigauderie, A. & Mooney, H. A. The Intergovernmental science-policy Platform on Biodiversity and Ecosystem Services: moving a step closer to an IPCC-like mechanism for biodiversity. Curr. Opin. Environ. Sustain. 2, 9–14 (2010).
    Article  Google Scholar 

    15.
    Convention on Biological Diversity. Aichi Biodiversity Targets http://www.cbd.int/sp/targets/default.shtml (2010).

    16.
    United Nations. Transforming Our World: The 2030 Agenda for Sustainable Development (United Nations General Assembly, 2015).

    17.
    Park, C.-E. et al. Keeping global warming within 1.5 °C constrains emergence of aridification. Nat. Clim. Change 8, 70–74 (2018).
    Article  Google Scholar 

    18.
    Sherwood, S. & Fu, Q. A drier future? Science 343, 737–739 (2014).
    Article  Google Scholar 

    19.
    Ukkola, A. M. et al. Reduced streamflow in water-stressed climates consistent with CO2 effects on vegetation. Nat. Clim. Change 6, 75–78 (2015).
    Article  Google Scholar 

    20.
    Wang, J. et al. Recent global decline in endorheic basin water storages. Nat. Geosci. 11, 926–932 (2018). Provides observational evidence for widespread loss of terrestrial water storage over global endorheic basins during 2002–2016 from climate variability and human water extractions.
    Article  Google Scholar 

    21.
    Scheff, J. & Frierson, D. M. W. Terrestrial aridity and its response to greenhouse warming across CMIP5 climate models. J. Clim. 28, 5583–5600 (2015).
    Article  Google Scholar 

    22.
    Koutroulis, A. G. Dryland changes under different levels of global warming. Sci. Total Environ. 655, 482–511 (2019).
    Article  Google Scholar 

    23.
    Schewe, J. et al. Multimodel assessment of water scarcity under climate change. Proc. Natl Acad. Sci. USA 111, 3245–3250 (2014).
    Article  Google Scholar 

    24.
    Feng, S. & Fu, Q. Expansion of global drylands under a warming climate. Atmos. Chem. Phys. Discuss. 13, 14637–14665 (2013).
    Google Scholar 

    25.
    Cook, B. I., Smerdon, J. E., Seager, R. & Coats, S. Global warming and 21st century drying. Clim. Dyn. 43, 2607–2627 (2014).
    Article  Google Scholar 

    26.
    Zhang, P. et al. Abrupt shift to hotter and drier climate over inner East Asia beyond the tipping point. Science 370, 1095–1099 (2020).
    Article  Google Scholar 

    27.
    He, B., Wang, S., Guo, L. & Wu, X. Aridity change and its correlation with greening over drylands. Agric. For. Meteorol. 278, 107663 (2019).
    Article  Google Scholar 

    28.
    Donohue, R. J., Roderick, M. L., McVicar, T. R. & Farquhar, G. D. Impact of CO2 fertilization on maximum foliage cover across the globe’s warm, arid environments. Geophys. Res. Lett. 40, 3031–3035 (2013). Reveals widespread greening in global arid regions despite warming, and provides quantitative theoretical evidence linking this greening pattern with elevated CO2.
    Article  Google Scholar 

    29.
    Fensholt, R. et al. Greenness in semi-arid areas across the globe 1981–2007 — an Earth Observing Satellite based analysis of trends and drivers. Remote Sens. Environ. 121, 144–158 (2012).
    Article  Google Scholar 

    30.
    Andela, N., Liu, Y. Y., van Dijk, A. I. J. M., de Jeu, R. A. M. & McVicar, T. R. Global changes in dryland vegetation dynamics (1988–2008) assessed by satellite remote sensing: comparing a new passive microwave vegetation density record with reflective greenness data. Biogeosciences 10, 6657–6676 (2013).
    Article  Google Scholar 

    31.
    Zhu, Z. et al. Greening of the Earth and its drivers. Nat. Clim. Change 6, 791–795 (2016).
    Article  Google Scholar 

    32.
    Piao, S. et al. Characteristics, drivers and feedbacks of global greening. Nat. Rev. Earth Environ. 1, 14–27 (2020).
    Article  Google Scholar 

    33.
    Beck, H. E. et al. Global evaluation of four AVHRR–NDVI data sets: Intercomparison and assessment against Landsat imagery. Remote Sens. Environ. 115, 2547–2563 (2011).
    Article  Google Scholar 

    34.
    United Nations World Water Assessment Programme. The United Nations World Water Development Report 2015: Water for a Sustainable World (UNESCO, 2015).

    35.
    Wang, L. et al. Dryland ecohydrology and climate change: critical issues and technical advances. Hydrol. Earth Syst. Sci. 16, 2585–2603 (2012).
    Article  Google Scholar 

    36.
    Roderick, M. L., Greve, P. & Farquhar, G. D. On the assessment of aridity with changes in atmospheric CO2. Water Resour. Res. 51, 5450–5463 (2015). A comprehensive summary of contradictory viewpoints of ‘warmer is more arid’ versus ‘warmer is less arid’ that arise from different interpretations of aridity changes, and provides a road map for reconciling such disparities.
    Article  Google Scholar 

    37.
    Swann, A. L. S. Plants and drought in a changing climate. Curr. Clim. Change Rep. 4, 192–201 (2018).
    Article  Google Scholar 

    38.
    Greve, P. et al. Global assessment of trends in wetting and drying over land. Nat. Geosci. 7, 716–721 (2014).
    Article  Google Scholar 

    39.
    Smith, W. K. et al. Remote sensing of dryland ecosystem structure and function: progress, challenges, and opportunities. Remote Sens. Environ. 233, 111401 (2019).
    Article  Google Scholar 

    40.
    Greve, P., Roderick, M. L., Ukkola, A. M. & Wada, Y. The aridity index under global warming. Environ. Res. Lett. 14, 124006 (2019).
    Article  Google Scholar 

    41.
    Huang, J., Yu, H., Dai, A., Wei, Y. & Kang, L. Drylands face potential threat under 2°C global warming target. Nat. Clim. Change 7, 417–422 (2017).
    Article  Google Scholar 

    42.
    Middleton, N. & Thomas, D. World Atlas of Desertification (Arnold, 1997).

    43.
    Fu, Q. & Feng, S. Responses of terrestrial aridity to global warming. J. Geophys. Res. Atmos. 119, 7863–7875 (2014).
    Article  Google Scholar 

    44.
    Yang, Y. et al. Disconnection between trends of atmospheric drying and continental runoff. Water Resour. Res. 54, 4700–4713 (2018).
    Article  Google Scholar 

    45.
    Yang, Y., Roderick, M. L., Zhang, S., McVicar, T. R. & Donohue, R. J. Hydrologic implications of vegetation response to elevated CO2 in climate projections. Nat. Clim. Change 9, 44–48 (2019).
    Article  Google Scholar 

    46.
    Greve, P., Roderick, M. L. & Seneviratne, S. I. Simulated changes in aridity from the last glacial maximum to 4xCO2. Environ. Res. Lett. 12, 114021 (2017).
    Article  Google Scholar 

    47.
    Yuan, W. et al. Increased atmospheric vapor pressure deficit reduces global vegetation growth. Sci. Adv. 5, eaax1396 (2019).
    Article  Google Scholar 

    48.
    Zhou, S. et al. Land–atmosphere feedbacks exacerbate concurrent soil drought and atmospheric aridity. Proc. Natl Acad. Sci. USA 116, 18848–18853 (2019).
    Article  Google Scholar 

    49.
    Grossiord, C. et al. Plant responses to rising vapor pressure deficit. New Phytol. 226, 1550–1566 (2020).
    Article  Google Scholar 

    50.
    Berdugo, M. et al. Global ecosystem thresholds driven by aridity. Science 367, 787–790 (2020).
    Article  Google Scholar 

    51.
    Keenan, T. F., Luo, X., Zhang, Y. & Zhou, S. Ecosystem aridity and atmospheric CO2. Science 368, 251–252 (2020).
    Google Scholar 

    52.
    Burrell, A. L., Evans, J. P. & De Kauwe, M. G. Anthropogenic climate change has driven over 5 million km2 of drylands towards desertification. Nat. Commun. 11, 3853 (2020).
    Article  Google Scholar 

    53.
    Trenberth, K. E. et al. Global warming and changes in drought. Nat. Clim. Change 4, 17–22 (2013).
    Article  Google Scholar 

    54.
    Milly, P. C. D. & Dunne, K. A. Potential evapotranspiration and continental drying. Nat. Clim. Change 6, 946–949 (2016).
    Article  Google Scholar 

    55.
    Milly, P. & Dunne, K. A. A hydrologic drying bias in water-resource impact analyses of anthropogenic climate change. J. Am. Water Resour. Assoc. 53, 822–838 (2017).
    Article  Google Scholar 

    56.
    De Jeu, R. A. et al. Global soil moisture patterns observed by space borne microwave radiometers and scatterometers. Surv. Geophys. 29, 399–420 (2008).
    Article  Google Scholar 

    57.
    Berg, A. & Sheffield, J. Climate change and drought: the soil moisture perspective. Curr. Clim. Change Rep. 4, 180–191 (2018).
    Article  Google Scholar 

    58.
    Feng, H. & Zhang, M. Global land moisture trends: drier in dry and wetter in wet over land. Sci. Rep. 5, 18018 (2015).
    Article  Google Scholar 

    59.
    Berg, A., Sheffield, J. & Milly, P. C. D. Divergent surface and total soil moisture projections under global warming. Geophys. Res. Lett. 44, 236–244 (2017).
    Article  Google Scholar 

    60.
    Cook, B. I. et al. Twenty-first century drought projections in the CMIP6 forcing scenarios. Earths Future 8, e2019EF001461 (2020).
    Article  Google Scholar 

    61.
    Li, M., Wu, P., Ma, Z., Lv, M. & Yang, Q. Changes in soil moisture persistence in China over the past 40 years under a warming climate. J. Clim. 33, 9531–9550 (2020).
    Article  Google Scholar 

    62.
    Rodell, M. et al. Emerging trends in global freshwater availability. Nature 557, 651–659 (2018).
    Article  Google Scholar 

    63.
    Dai, A., Zhao, T. & Chen, J. Climate change and drought: a precipitation and evaporation perspective. Curr. Clim. Change Rep. 4, 301–312 (2018). Highlights the dominant role of CO2radiative forcing in shaping global land surface drying patterns for the twenty-first century.
    Article  Google Scholar 

    64.
    Schlaepfer, D. R. et al. Climate change reduces extent of temperate drylands and intensifies drought in deep soils. Nat. Commun. 8, 14196 (2017).
    Article  Google Scholar 

    65.
    Li, L. et al. Global trends in water and sediment fluxes of the world’s large rivers. Sci. Bull. 65, 62–69 (2019).
    Article  Google Scholar 

    66.
    Yang, H. et al. Regional patterns of future runoff changes from Earth system models constrained by observation. Geophys. Res. Lett. 44, 5540–5549 (2017).
    Article  Google Scholar 

    67.
    Wang, S. et al. Reduced sediment transport in the Yellow River due to anthropogenic changes. Nat. Geosci. 9, 38–41 (2015).
    Article  Google Scholar 

    68.
    Milly, P. C. D. & Dunne, K. A. Colorado River flow dwindles as warming-driven loss of reflective snow energizes evaporation. Science 367, 1252–1255 (2020).
    Article  Google Scholar 

    69.
    Trancoso, R., Larsen, J. R., McVicar, T. R., Phinn, S. R. & McAlpine, C. A. CO2-vegetation feedbacks and other climate changes implicated in reducing base flow. Geophys. Res. Lett. 44, 2310–2318 (2017).
    Article  Google Scholar 

    70.
    Dai, A., Qian, T., Trenberth, K. E. & Milliman, J. D. Changes in continental freshwater discharge from 1948 to 2004. J. Clim. 22, 2773–2792 (2009).
    Article  Google Scholar 

    71.
    Hobeichi, S., Abramowitz, G., Evans, J. & Beck, H. E. Linear Optimal Runoff Aggregate (LORA): A global gridded synthesis runoff product. Hydrol. Earth Syst. Sci. 23, 851–870 (2019).
    Article  Google Scholar 

    72.
    Novick, K. A. et al. The increasing importance of atmospheric demand for ecosystem water and carbon fluxes. Nat. Clim. Change 6, 1023–1027 (2016).
    Article  Google Scholar 

    73.
    Chen, C. et al. China and India lead in greening of the world through land-use management. Nat. Sustain. 2, 122–129 (2019).
    Article  Google Scholar 

    74.
    Eyring, V. et al. Overview of the Coupled Model Intercomparison Project Phase 6 (CMIP6) experimental design and organization. Geosci. Model Dev. 9, 1937–1958 (2016).
    Article  Google Scholar 

    75.
    Eyring, V. et al. Taking climate model evaluation to the next level. Nat. Clim. Change 9, 102–110 (2019).
    Article  Google Scholar 

    76.
    Berg, A. et al. Land–atmosphere feedbacks amplify aridity increase over land under global warming. Nat. Clim. Change 6, 869–874 (2016). Reveals the important mechanism that declining soil moisture and altered vegetation physiology under climate change and rising CO2could make the near-surface air even warmer and drier.
    Article  Google Scholar 

    77.
    Stephens, C. M., McVicar, T. R., Johnson, F. M. & Marshall, L. A. Revisiting pan evaporation trends in Australia a decade on. Geophys. Res. Lett. 45, 11164–11172 (2018).
    Google Scholar 

    78.
    Byrne, M. P. & O’Gorman, P. A. The response of precipitation minus evapotranspiration to climate warming: why the “wet-get-wetter, dry-get-drier” scaling does not hold over land. J. Clim. 28, 8078–8092 (2015).
    Article  Google Scholar 

    79.
    Zhou, S. et al. Soil moisture–atmosphere feedbacks mitigate declining water availability in drylands. Nat. Clim. Change 11, 38–44 (2020).
    Article  Google Scholar 

    80.
    Lau, W. K. & Kim, K. M. Robust Hadley circulation changes and increasing global dryness due to CO2 warming from CMIP5 model projections. Proc. Natl Acad. Sci. USA 112, 3630–3635 (2015).
    Article  Google Scholar 

    81.
    Lau, W. K. M. & Tao, W. Precipitation–radiation–circulation feedback processes associated with structural changes of the ITCZ in a warming climate during 1980–2014: an observational portrayal. J. Clim. 33, 8737–8749 (2020).
    Article  Google Scholar 

    82.
    Burls, N. J. & Fedorov, A. V. Wetter subtropics in a warmer world: contrasting past and future hydrological cycles. Proc. Natl Acad. Sci. USA 114, 12888–12893 (2017).
    Article  Google Scholar 

    83.
    Condon, L. E., Atchley, A. L. & Maxwell, R. M. Evapotranspiration depletes groundwater under warming over the contiguous United States. Nat. Commun. 11, 873 (2020).
    Article  Google Scholar 

    84.
    Jung, M. et al. Recent decline in the global land evapotranspiration trend due to limited moisture supply. Nature 467, 951–954 (2010).
    Article  Google Scholar 

    85.
    García, M. et al. Actual evapotranspiration in drylands derived from in-situ and satellite data: Assessing biophysical constraints. Remote Sens. Environ. 131, 103–118 (2013).
    Article  Google Scholar 

    86.
    Betts, R. A. et al. Projected increase in continental runoff due to plant responses to increasing carbon dioxide. Nature 448, 1037–1041 (2007).
    Article  Google Scholar 

    87.
    Swann, A. L., Hoffman, F. M., Koven, C. D. & Randerson, J. T. Plant responses to increasing CO2 reduce estimates of climate impacts on drought severity. Proc. Natl Acad. Sci. USA 113, 10019–10024 (2016).
    Article  Google Scholar 

    88.
    Fowler, M. D., Kooperman, G. J., Randerson, J. T. & Pritchard, M. S. The effect of plant physiological responses to rising CO2 on global streamflow. Nat. Clim. Change 9, 873–879 (2019).
    Article  Google Scholar 

    89.
    Lemordant, L., Gentine, P., Swann, A. S., Cook, B. I. & Scheff, J. Critical impact of vegetation physiology on the continental hydrologic cycle in response to increasing CO2. Proc. Natl Acad. Sci. USA 115, 4093–4098 (2018).
    Article  Google Scholar 

    90.
    Haverd, V. et al. Higher than expected CO2 fertilization inferred from leaf to global observations. Glob. Change Biol. 26, 2390–2402 (2020).
    Article  Google Scholar 

    91.
    Nie, M., Lu, M., Bell, J., Raut, S. & Pendall, E. Altered root traits due to elevated CO2: a meta-analysis. Glob. Ecol. Biogeogr. 22, 1095–1105 (2013).
    Article  Google Scholar 

    92.
    Zhang, Y. et al. Multi-decadal trends in global terrestrial evapotranspiration and its components. Sci. Rep. 6, 19124 (2016).
    Article  Google Scholar 

    93.
    Sun, X., Wilcox, B. P. & Zou, C. B. Evapotranspiration partitioning in dryland ecosystems: A global meta-analysis of in situ studies. J. Hydrol. 576, 123–136 (2019).
    Article  Google Scholar 

    94.
    Lian, X. et al. Partitioning global land evapotranspiration using CMIP5 models constrained by observations. Nat. Clim. Change 8, 640–646 (2018).
    Article  Google Scholar 

    95.
    Yang, H., Huntingford, C., Wiltshire, A., Sitch, S. & Mercado, L. Compensatory climate effects link trends in global runoff to rising atmospheric CO2 concentration. Environ. Res. Lett. 14, 124075 (2019).
    Article  Google Scholar 

    96.
    Mankin, J. S. et al. Blue water trade-offs with vegetation in a CO2-enriched climate. Geophys. Res. Lett. 45, 3115–3125 (2018).
    Article  Google Scholar 

    97.
    Stocker, B. D. et al. Quantifying soil moisture impacts on light use efficiency across biomes. New Phytol. 218, 1430–1449 (2018).
    Article  Google Scholar 

    98.
    Liu, L. et al. Soil moisture dominates dryness stress on ecosystem production globally. Nat. Commun. 11, 4892 (2020).
    Article  Google Scholar 

    99.
    Morgan, J. A. et al. C4 grasses prosper as carbon dioxide eliminates desiccation in warmed semi-arid grassland. Nature 476, 202–205 (2011).
    Article  Google Scholar 

    100.
    Farrior, C. E., Rodriguez-Iturbe, I., Dybzinski, R., Levin, S. A. & Pacala, S. W. Decreased water limitation under elevated CO2 amplifies potential for forest carbon sinks. Proc. Natl Acad. Sci. USA 112, 7213–7218 (2015).
    Article  Google Scholar 

    101.
    Lu, X., Wang, L. & McCabe, M. F. Elevated CO2 as a driver of global dryland greening. Sci. Rep. 6, 20716 (2016).
    Article  Google Scholar 

    102.
    Ukkola, A. M., Keenan, T. F., Kelley, D. I. & Prentice, I. C. Vegetation plays an important role in mediating future water resources. Environ. Res. Lett. 11, 094022 (2016).
    Article  Google Scholar 

    103.
    Mankin, J. S., Smerdon, J. E., Cook, B. I., Williams, A. P. & Seager, R. The curious case of projected twenty-first-century drying but greening in the American West. J. Clim. 30, 8689–8710 (2017).
    Article  Google Scholar 

    104.
    Zarakas, C. M., Swann, A. L. S., Laguë, M. M., Armour, K. C. & Randerson, J. T. Plant physiology increases the magnitude and spread of the transient climate response to CO2 in CMIP6 Earth system models. J. Clim. 33, 8561–8578 (2020).
    Article  Google Scholar 

    105.
    Mankin, J. S., Seager, R., Smerdon, J. E., Cook, B. I. & Williams, A. P. Mid-latitude freshwater availability reduced by projected vegetation responses to climate change. Nat. Geosci. 12, 983–988 (2019).
    Article  Google Scholar 

    106.
    Song, J. et al. Elevated CO2 does not stimulate carbon sink in a semi-arid grassland. Ecol. Lett. 22, 458–468 (2019).
    Article  Google Scholar 

    107.
    Obermeier, W. A. et al. Reduced CO2 fertilization effect in temperate C3 grasslands under more extreme weather conditions. Nat. Clim. Change 7, 137–141 (2016).
    Article  Google Scholar 

    108.
    Craine, J. M. et al. Isotopic evidence for oligotrophication of terrestrial ecosystems. Nat. Ecol. Evol. 2, 1735–1744 (2018).
    Article  Google Scholar 

    109.
    Medlyn, B. E. et al. How do leaf and ecosystem measures of water-use efficiency compare? New Phytol. 216, 758–770 (2017).
    Article  Google Scholar 

    110.
    Keenan, T. F. et al. Increase in forest water-use efficiency as atmospheric carbon dioxide concentrations rise. Nature 499, 324–327 (2013).
    Article  Google Scholar 

    111.
    Peters, W. et al. Increased water-use efficiency and reduced CO2 uptake by plants during droughts at a continental-scale. Nat. Geosci. 11, 744–748 (2018).
    Article  Google Scholar 

    112.
    Skinner, C. B., Poulsen, C. J. & Mankin, J. S. Amplification of heat extremes by plant CO2 physiological forcing. Nat. Commun. 9, 1094 (2018).
    Article  Google Scholar 

    113.
    Lemordant, L. & Gentine, P. Vegetation response to rising CO2 impacts extreme temperatures. Geophys. Res. Lett. 46, 1383–1392 (2019).
    Article  Google Scholar 

    114.
    Sellers, P. J. et al. Comparison of radiative and physiological effects of doubled atmospheric CO2 on climate. Science 271, 1402–1406 (1996).
    Article  Google Scholar 

    115.
    Warren, J. M., Norby, R. J. & Wullschleger, S. D. Elevated CO2 enhances leaf senescence during extreme drought in a temperate forest. Tree Physiol. 31, 117–130 (2011).
    Article  Google Scholar 

    116.
    De Kauwe, M. G. et al. Examining the evidence for decoupling between photosynthesis and transpiration during heat extremes. Biogeosciences 16, 903–916 (2019).
    Article  Google Scholar 

    117.
    Huang, M. et al. Air temperature optima of vegetation productivity across global biomes. Nat. Ecol. Evol. 3, 772–779 (2019).
    Article  Google Scholar 

    118.
    Reich, P. B., Hobbie, S. E., Lee, T. D. & Pastore, M. A. Unexpected reversal of C3 versus C4 grass response to elevated CO2 during a 20-year field experiment. Science 360, 317–320 (2018).
    Article  Google Scholar 

    119.
    Norby, R. J. et al. Model-data synthesis for the next generation of forest free-air CO2 enrichment (FACE) experiments. New Phytol. 209, 17–28 (2016).
    Article  Google Scholar 

    120.
    Steffen, W. et al. The emergence and evolution of Earth System Science. Nat. Rev. Earth Environ. 1, 54–63 (2020).
    Article  Google Scholar 

    121.
    Hoekstra, A. Y. & Mekonnen, M. M. The water footprint of humanity. Proc. Natl Acad. Sci. USA 109, 3232–3237 (2012).
    Article  Google Scholar 

    122.
    Marvel, K. et al. Twentieth-century hydroclimate changes consistent with human influence. Nature 569, 59–65 (2019).
    Article  Google Scholar 

    123.
    Di Baldassarre, G. et al. Sociohydrology: scientific challenges in addressing the sustainable development goals. Water Resour. Res. 55, 6327–6355 (2019).
    Article  Google Scholar 

    124.
    van der Esch, S. et al. Exploring Future Changes in Land Use and Land Condition and the Impacts on Food, Water, Climate Change and Biodiversity: Scenarios for the UNCCD Global Land Outlook (PBL Netherlands Environmental Assessment Agency, 2017).

    125.
    Gleick, P. H. Transitions to freshwater sustainability. Proc. Natl Acad. Sci. USA 115, 8863–8871 (2018).
    Article  Google Scholar 

    126.
    Wada, Y., de Graaf, I. E. M. & van Beek, L. P. H. High-resolution modeling of human and climate impacts on global water resources. J. Adv. Model Earth Syst. 8, 735–763 (2016).
    Article  Google Scholar 

    127.
    Wada, Y. et al. Modeling global water use for the 21st century: the Water Futures and Solutions (WFaS) initiative and its approaches. Geosci. Model Dev. 9, 175–222 (2016). Provides an ensemble model projection of significant increases in the twenty-first century’s water demand by major water-use sectors under envisaged population growth and socio-economic developments.
    Article  Google Scholar 

    128.
    Wada, Y., van Beek, L. P. H. & Bierkens, M. F. P. Nonsustainable groundwater sustaining irrigation: a global assessment. Water Resour. Res. 48, W00L06 (2012).
    Article  Google Scholar 

    129.
    Chen, Y. et al. Recent global cropland water consumption constrained by observations. Water Resour. Res. 55, 3708–3738 (2019).
    Article  Google Scholar 

    130.
    Allen, L. H. Jr., Kakani, V. G., Vu, J. C. & Boote, K. J. Elevated CO2 increases water use efficiency by sustaining photosynthesis of water-limited maize and sorghum. J. Plant Physiol. 168, 1909–1918 (2011).
    Article  Google Scholar 

    131.
    Elliott, J. et al. Constraints and potentials of future irrigation water availability on agricultural production under climate change. Proc. Natl Acad. Sci. USA 111, 3239–3244 (2014).
    Article  Google Scholar 

    132.
    Urban, D. W., Sheffield, J. & Lobell, D. B. Historical effects of CO2 and climate trends on global crop water demand. Nat. Clim. Change 7, 901–905 (2017).
    Article  Google Scholar 

    133.
    Gleeson, T., Wada, Y., Bierkens, M. F. & van Beek, L. P. Water balance of global aquifers revealed by groundwater footprint. Nature 488, 197–200 (2012).
    Article  Google Scholar 

    134.
    Bierkens, M. F. P. & Wada, Y. Non-renewable groundwater use and groundwater depletion: a review. Environ. Res. Lett. 14, 063002 (2019).
    Article  Google Scholar 

    135.
    Rodell, M., Velicogna, I. & Famiglietti, J. S. Satellite-based estimates of groundwater depletion in India. Nature 460, 999–1002 (2009).
    Article  Google Scholar 

    136.
    Feng, W. et al. Evaluation of groundwater depletion in North China using the Gravity Recovery and Climate Experiment (GRACE) data and ground-based measurements. Water Resour. Res. 49, 2110–2118 (2013).
    Article  Google Scholar 

    137.
    Eamus, D. & Froend, R. Groundwater-dependent ecosystems: the where, what and why of GDEs. Aust. J. Bot. 54, 91–96 (2006).
    Article  Google Scholar 

    138.
    Griebler, C. & Avramov, M. Groundwater ecosystem services: a review. Freshw. Sci. 34, 355–367 (2015).
    Article  Google Scholar 

    139.
    Devitt, T. J., Wright, A. M., Cannatella, D. C. & Hillis, D. M. Species delimitation in endangered groundwater salamanders: implications for aquifer management and biodiversity conservation. Proc. Natl Acad. Sci. USA 116, 2624–2633 (2019).
    Article  Google Scholar 

    140.
    Feng, X. et al. Revegetation in China’s Loess Plateau is approaching sustainable water resource limits. Nat. Clim. Change 6, 1019–1022 (2016).
    Article  Google Scholar 

    141.
    Hong, S. et al. Divergent responses of soil organic carbon to afforestation. Nat. Sustain. 3, 694–700 (2020).
    Article  Google Scholar 

    142.
    McVicar, T. R. et al. Developing a decision support tool for China’s re-vegetation program: Simulating regional impacts of afforestation on average annual streamflow in the Loess Plateau. For. Ecol. Manag. 251, 65–81 (2007).
    Article  Google Scholar 

    143.
    Zhao, M. et al. Ecological restoration impact on total terrestrial water storage. Nat. Sustain. 4, 56–62 (2020).
    Article  Google Scholar 

    144.
    Kwon, H.-Y. et al. in Economics of Land Degradation and Improvement – A Global Assessment for Sustainable Development Ch. 8 (eds Nkonya E., Mirzabaev A. & von Braun J.) 197-214 (Springer, 2016).

    145.
    Asner, G. P., Elmore, A. J., Olander, L. P., Martin, R. E. & Harris, A. T. Grazing systems, ecosystem responses, and global change. Annu. Rev. Environ. Resour. 29, 261–299 (2004).
    Article  Google Scholar 

    146.
    Dunne, T., Western, D. & Dietrich, W. E. Effects of cattle trampling on vegetation, infiltration, and erosion in a tropical rangeland. J. Arid. Environ. 75, 58–69 (2011).
    Article  Google Scholar 

    147.
    Griscom, B. W. et al. Natural climate solutions. Proc. Natl Acad. Sci. USA 114, 11645–11650 (2017).
    Article  Google Scholar 

    148.
    Lewis, S. L., Wheeler, C. E., Mitchard, E. T. & Koch, A. Restoring natural forests is the best way to remove atmospheric carbon. Nature 568, 25–28 (2019).
    Article  Google Scholar 

    149.
    Reisman-Berman, O., Keasar, T. & Tel-Zur, N. Native and non-native species for dryland afforestation: bridging ecosystem integrity and livelihood support. Ann. For. Sci. 76, 114 (2019).
    Article  Google Scholar 

    150.
    Zhang, J. et al. Carrying capacity for vegetation across northern China drylands. Sci. Total Environ. 710, 136391 (2020).
    Article  Google Scholar 

    151.
    Liu, Y., Kumar, M., Katul, G. G. & Porporato, A. Reduced resilience as an early warning signal of forest mortality. Nat. Clim. Change 9, 880–885 (2019).
    Article  Google Scholar 

    152.
    Fita, A., Rodríguez-Burruezo, A., Boscaiu, M., Prohens, J. & Vicente, O. Breeding and domesticating crops adapted to drought and salinity: a new paradigm for increasing food production. Front. Plant Sci. 6, 978 (2015).
    Article  Google Scholar 

    153.
    Graham, N. T. et al. Water sector assumptions for the Shared Socioeconomic Pathways in an integrated modeling framework. Water Resour. Res. 54, 6423–6440 (2018).
    Article  Google Scholar 

    154.
    Muhs, D. R. The geologic records of dust in the Quaternary. Aeolian Res. 9, 3–48 (2013).
    Article  Google Scholar 

    155.
    Scheff, J., Seager, R., Liu, H. & Coats, S. Are glacials dry? Consequences for paleoclimatology and for greenhouse warming. J. Clim. 30, 6593–6609 (2017).
    Article  Google Scholar 

    156.
    Lambert, F. et al. Dust-climate couplings over the past 800,000 years from the EPICA Dome C ice core. Nature 452, 616–619 (2008).
    Article  Google Scholar 

    157.
    Salzmann, U. et al. Climate and environment of a Pliocene warm world. Palaeogeogr. Palaeoclimatol. Palaeoecol. 309, 1–8 (2011).
    Article  Google Scholar 

    158.
    Fu, Q., Lin, L., Huang, J., Feng, S. & Gettelman, A. Changes in terrestrial aridity for the period 850–2080 from the Community Earth System Model. J. Geophys. Res. Atmos. 121, 2857–2873 (2016).
    Article  Google Scholar 

    159.
    Prudhomme, C. et al. Hydrological droughts in the 21st century, hotspots and uncertainties from a global multimodel ensemble experiment. Proc. Natl Acad. Sci. USA 111, 3262–3267 (2014).
    Article  Google Scholar 

    160.
    Cook, B. I., Ault, T. R. & Smerdon, J. E. Unprecedented 21st century drought risk in the American Southwest and Central Plains. Sci. Adv. 1, e1400082 (2015).
    Article  Google Scholar 

    161.
    Ault, T. R. On the essentials of drought in a changing climate. Science 368, 256–260 (2020).
    Article  Google Scholar 

    162.
    Reichstein, M. et al. Climate extremes and the carbon cycle. Nature 500, 287–295 (2013).
    Article  Google Scholar 

    163.
    Anderegg, W. R. L., Kane, J. M. & Anderegg, L. D. L. Consequences of widespread tree mortality triggered by drought and temperature stress. Nat. Clim. Change 3, 30–36 (2012).
    Article  Google Scholar 

    164.
    Williams, A. P. et al. Forest responses to increasing aridity and warmth in the southwestern United States. Proc. Natl Acad. Sci. USA 107, 21289–21294 (2010).
    Article  Google Scholar 

    165.
    Pellegrini, A. F. A. et al. Fire frequency drives decadal changes in soil carbon and nitrogen and ecosystem productivity. Nature 553, 194–198 (2018).
    Article  Google Scholar 

    166.
    Bowman, D. M. J. S. et al. Vegetation fires in the Anthropocene. Nat. Rev. Earth Environ. 1, 500–515 (2020).
    Article  Google Scholar 

    167.
    Andela, N. et al. A human-driven decline in global burned area. Science 356, 1356–1362 (2017).
    Article  Google Scholar 

    168.
    Pechony, O. & Shindell, D. T. Driving forces of global wildfires over the past millennium and the forthcoming century. Proc. Natl Acad. Sci. USA 107, 19167–19170 (2010).
    Article  Google Scholar 

    169.
    Hoover, D. L., Knapp, A. K. & Smith, M. D. Resistance and resilience of a grassland ecosystem to climate extremes. Ecology 95, 2646–2656 (2014).
    Article  Google Scholar 

    170.
    Greve, P. et al. Global assessment of water challenges under uncertainty in water scarcity projections. Nat. Sustain. 1, 486–494 (2018).
    Article  Google Scholar 

    171.
    Scanlon, B. R. et al. Global models underestimate large decadal declining and rising water storage trends relative to GRACE satellite data. Proc. Natl Acad. Sci. USA 115, E1080–E1089 (2018).
    Article  Google Scholar 

    172.
    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).
    Article  Google Scholar 

    173.
    Roderick, M. L., Sun, F., Lim, W. H. & Farquhar, G. D. A general framework for understanding the response of the water cycle to global warming over land and ocean. Hydrol. Earth Syst. Sci. 18, 1575–1589 (2014).
    Article  Google Scholar 

    174.
    Gudmundsson, L., Greve, P. & Seneviratne, S. I. The sensitivity of water availability to changes in the aridity index and other factors—A probabilistic analysis in the Budyko space. Geophys. Res. Lett. 43, 6985–6994 (2016).
    Article  Google Scholar 

    175.
    American Meteorological Society. Glossary of Meteorology http://glossary.ametsoc.org/wiki/Aridity (2000).

    176.
    Allen, R. G., Pereira, L. S., Raes, D. & Smith, M. Crop Evapotranspiration — Guidelines for Computing Crop Water Requirements — FAO Irrigation and Drainage Paper 56 (Food and Agriculture Organization of the United Nations, 1998).

    177.
    Dai, A. Drought under global warming: a review. Wiley Interdiscip. Rev. Clim. Change 2, 45–65 (2011).
    Article  Google Scholar 

    178.
    Dai A. in Terrestrial Water Cycle and Climate Change: Natural and Human-Induced Impacts 1st edn, Ch. 2 (eds Tang, Q. & Oki, T.) 17-37 (Wiley, 2016).

    179.
    Sitch, S. et al. Trends and drivers of regional sources and sinks of carbon dioxide over the past two decades. Biogeosci. Discuss. 10, 20113–20177 (2013).
    Google Scholar 

    180.
    Donohue, R. J., Roderick, M. L., McVicar, T. R. & Yang, Y. A simple hypothesis of how leaf and canopy-level transpiration and assimilation respond to elevated CO2 reveals distinct response patterns between disturbed and undisturbed vegetation. J. Geophys. Res. Biogeosci. 122, 168–184 (2017).
    Article  Google Scholar 

    181.
    Barton, C. V. M. et al. Effects of elevated atmospheric [CO2] on instantaneous transpiration efficiency at leaf and canopy scales in Eucalyptus saligna. Glob. Change Biol. 18, 585–595 (2012).
    Article  Google Scholar 

    182.
    Savvides, A. M. & Fotopoulos, V. Two inexpensive and non-destructive techniques to correct for smaller-than-gasket leaf area in gas exchange measurements. Front. Plant Sci. 9, 548 (2018).
    Article  Google Scholar  More

  • in

    Hidden diversity of the most basal tapeworms (Cestoda, Gyrocotylidea), the enigmatic parasites of holocephalans (Chimaeriformes)

    Almost 50 years ago, Simmons26 called gyrocotylideans a “century-old enigma” and this status still persists despite the advent of more advanced identification methods3. The poor understanding of the group (e.g., the complete life cycle of none of the species is known) is linked with the scarcity of available data and the biological peculiarities of these tapeworms and their holocephalan hosts. In particular, most of the host species are rarely available deep-sea dwellers, which often could not be examined fresh or were frozen with their parasites prior to examination. If isolated alive, gyrocotylideans exhibit an unusual morphological variability due to the contraction of their large bodies and as a result of different fixative procedures which were tested to ensure their relaxation (e.g.27). Despite these issues, several comprehensive studies have been conducted, e.g.15,16,21,28, which provided deep insight into the biology, ecology and taxonomy of these enigmatic tapeworms. Nevertheless, the poor quality of the specimens studied and the use of different, not always appropriate, methods of parasite fixation, unintentionally affected the quality of morphological descriptions of most gyrocotylidean species, which prevented the establishing of clear morphological borders to delimit individual species. As a result, the informative value of morphological traits used for species delimitation should be re-assessed, based on the simultaneous use of molecular data, i.e., the use of hologenophores to match morphology and molecular data. Existing problems with species delimitation and morphological variability even led to complete omission of morphological characterisation of two new species described just recently6.
    Herein, the genotyping of the Gyrocotyle spp. specimens acquired in Taiwan revealed four distinct genotypes, each one more related to the North Atlantic isolates identified as “Gyrocotyle urna” off Ireland (the isolate is genetically diverse from G. urna off Norway), “G. rugosa” off Alaska (probably misidentified, see below), G. discoveryi off Ireland and G. confusa off Norway, respectively, than to each other.
    In addition to casting doubts on the restriction of gyrocotylideans to individual oceans, our data also question the proclaimed strict host specificity3,7, because specimens of Gyrocotyle sp. genotype 3 were found in two hosts species, which are not the closest relatives to one another—C. phantasma and C. cf. argiloba (Fig. 4). Broader host specificity was also reported for G. fimbriata, which was found in Hydrolagus colliei and Chimaera phantasma, and for G. rugosa, recorded in Callorhinchus callorynchus and C. milii14,15,24,29. Gyrocotyle urna was also found in several holocephalans, including Chimaera monstrosa, Callorhinchus callorynchus, Hydrolagus ogilbyi Waite and H. colliei24,29,30. In contrast, Bandoni & Brooks16 revised the host spectrum of this parasite, considering C. monstrosa as the only host of G. urna.
    The suitability of the molecular markers employed for this group also requires attention, because a considerable amount of phylogenetic information was also lost in the un-rooted dataset due to treatment of the numerous gaps in the 28S rRNA alignment. The involvement of partial COI gene sequences seemed to be informative for estimating gyrocotylidean phylogeny, because we obtained a no-gap COI alignment and improved support for some nodes in the three-gene network. The suitability of this marker requires assessment employing further taxa, because except for our isolates off Taiwan and Argentina, only a single sequence of the COI gene (i.e., that of G. urna off Norway; GenBank acc. no. JQ268546) is currently available.
    A single specimen of Gyrocotyle sp. genotype 4 was conspicuously different morphologically from the remaining ones by having few folds on the lateral margins, many acetabular spines, a narrow funnel and a small rosette. However, its formal description as a new species would be premature, because only a single specimen was found. Morphological differences among the specimens of the other genotypes were not so obvious, even though a careful examination of the hologenophores allowed us to find several morphological traits that were characteristic for particular genotypes (see “Results” section). Among them, the number of acetabular spines and the distribution of the body spines and their size may be potentially useful for species differentiation, especially because the body contraction can hardly affect them. Since body contraction cannot be absolutely excluded even when live specimens are properly fixed, its effect could be overcome to some degree by an evaluation of ratios related to the main body dimensions (e.g., length of uterine sac/total body length) rather than comparison of total measurements of internal structures.
    The specimens off Taiwan most probably represent several new species, but we decided not to describe them formally as new taxa, mainly because of the shortage of comparative data. In addition to these specimens, two hologenophores of Gyrocotyle rugosa off Argentina were examined, which made it possible to characterise the type species of the genus. The host of G. rugosa described by Diesing10 was questionable until Callorhynchus antarcticus (= C. callorynchus—see31) off New Zealand was finally established as its currently accepted type host3,32. Gyrocotyle rugosa was found in coastal waters of South America, South Africa and New Zealand as a parasite of C. callorynchus and C. milii, suggesting its broader host specificity16,24. Our specimens from C. callorynchus off Argentina were identified as G. rugosa based on crenulated (i.e., without any folds) lateral margins, a tiny uterine sac, a branched uterus and embryonated eggs in the uterine sac; the latter two traits are unique to this species21. Genetically, it clustered with an unspecified isolate of Gyrocotyle from C. milii off Australia, and these specimens seem to be conspecific.
    In contrast, an isolate from Hydrolagus colliei off Alaska identified as G. rugosa (GenBank acc. nos. AF286925 and AF124455) was apparently misidentified, because (i) it was found in an unrelated definitive host (H. colliei belongs to the family Chimaeridae, whereas the type host to the family Callorhinchidae), (ii) its distant geographic origin (the type locality of G. rugosa is unclear, but it is definitely in the Southern hemisphere), and (iii) its genetic divergence from our isolate of G. rugosa from the type host off Argentina. The isolate from H. colliei may represent Gyrocotyle fimbriata or G. parvispinosa, which have been reported from this host off the Pacific coast of North America, but its identification was not possible because morphological vouchers were not available to the present authors.
    Gyrocotylideans were generally considered to be oioxenous, i.e. strictly specific parasites sensu Euzet and Combes33, with each gyrocotylidean species parasitising a single holocephalan species. Although several species were reported from two or more hosts species16,24, these findings are usually considered as misidentifications due to the unclear taxonomy of the order. Moreover, some holocephalans, such as Ch. monstrosa, H. colliei, H. affinis, and Ca. callorynchus, were often found to harbour two or more gyrocotylidean species, one common and the other rare9,10,21,22,23. Our findings of Gyrocotyle sp. genotypes 1 and 3 in Ch. phantasma and Gyrocotyle sp. genotypes 2, 3 and 4 in Ch. cf. argiloba suggested stenoxenous host specificity (i.e., the occurrence in a few closely related hosts) of gyrocotylideans, because the specimens of genotype 3 were found in both species of Chimaera. The obvious genetic similarity of our G. rugosa specimen from Ca. callorynchus and the isolate of Gyrocotyle sp. from Ca. milii also questions the strict specificity of this group, but morphological vouchers of the latter, which are necessary for the confirmation of their conspecificity, are not available.
    Our genetic analyses provided insight into the interrelationships among the gyrocotylideans, even though the absence of a suitable outgroup did not enable us to broadly assess the possible evolutionary scenario of this earliest evolving group of tapeworms. Moreover, genetic data on only half of the nominal species of Gyrocotyle are available, not considering the possibility of misidentifications of previously sequenced specimens, for which hologenophores are not available. However, some clues of host-parasite coevolution can be inferred from the network. The mutual genetic distance of species/genotypes from the same host species suggests multiple colonisation events rather than co-speciation with their hosts within the order. It seems that G. phantasma might have been colonised by Gyrocotyle sp. genotype 1 or genotype 3, because these two genotypes are not the closest relatives in our analyses. The same pattern is obvious for C. cf. argiloba parasitised by Gyrocotyle sp. genotype 2, 3 and 4, and also for C. monstrosa, which harbours G. urna, G. confusa and G. nybelini. Indeed, Colin et al.27 considered these species from C. monstrosa to be conspecific, but our genetic data support the validity of three separate and genetically distant species. Moreover, G. nybelini formed by far the most distant lineage among all isolates, which may suggest the validity of the genus Gyrocotyloides Furhmann, 1931.
    Genetic divergence of congeneric tapeworms from the same host species was also observed in several elasmobranch/teleost-cestode assemblages, e.g., Acanthobothrium spp. (Onchoproteocephalidea) and the mumburarr whipray Urogymnus acanthobothrium Last, White & Kyne; Echeneibothrium spp. (Rhinebothriidea) and the yellownose skate Dipturus chilensis (Guichenot); and Pseudoendorchis spp. (Onchoproteocephalidea) and the catfish Pimelodus maculatus Lacepède34,35,36.
    The aim of this paper was to provide new insight into the phylogenetic relationships within the enigmatic order Gyrocotylidea, but, in particular, to demonstrate the lack of geographical patterns in the distribution of most its species and the limited suitability of current morphological characteristics for species circumscription. Herein, we have outlined a methodology (fixation of live specimens with hot fixative and the exclusive use of hologenophores) that should be used in future taxonomic, ecological and biogeographical studies of gyrocotylideans in order to reliably circumscribe their actual species diversity and to unravel associations with their hosts, a relict group of marine vertebrates. Gyrocotylideans represent one of the key groups of parasitic flatworms (Neodermata) in terms of a better understanding of their evolutionary history and the switch of free-living flatworms to parasitism. More

  • in

    Telomere lengths correlate with fitness but assortative mating by telomeres confers no benefit to fledgling recruitment

    1.
    Jiang, Y., Bolnick, D. I. & Kirkpatrick, M. Assortative mating in animals. Am. Nat. 181, E125–E138. https://doi.org/10.1086/670160 (2013).
    Article  PubMed  Google Scholar 
    2.
    Cooke, F., Finney, G. H. & Rockwell, R. F. Assortative mating in lesser snow geese (Anser caerulescens). Behav. Genet. 6, 127–140 (1976).
    CAS  Article  Google Scholar 

    3.
    Otronen, M. Size assortative mating in the yellow dung fly Scatophaga stercoraria. Behavior 126, 63–76 (1993).
    Article  Google Scholar 

    4.
    Wang, D. et al. Scrutinizing assortative mating in birds. PLoS Biol. 17, e3000156. https://doi.org/10.1371/journal.pbio.3000156 (2019).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    5.
    Sin, Y. W. et al. MHC class II-assortative mate choice in European badgers (Meles meles). Mol. Ecol. 24, 3138–3150. https://doi.org/10.1111/mec.13217 (2015).
    Article  PubMed  Google Scholar 

    6.
    Angelier, F., Weimerskirch, H., Barbraud, C. & Chastel, O. Is telomere length a molecular marker of individual quality? Insights from a long‐lived bird. Funct. Ecol. https://doi.org/10.1111/1365-2435.13307 (in press).

    7.
    Monaghan, P. Organismal stress, telomeres and life histories. J. Exp. Biol. 217, 57–66. https://doi.org/10.1242/jeb.090043 (2014).
    Article  PubMed  Google Scholar 

    8.
    Bronikowski, A. M. The evolution of aging phenotypes in snakes: a review and synthesis with new data. Age 30, 169–176. https://doi.org/10.1007/s11357-008-9060-5 (2008).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    9.
    Tricola, G. et al. The rate of telomere loss is related to maximum lifespan in birds. Philos. Trans. R. Soc. B 373, 20160445. https://doi.org/10.1098/rstb.2016.0445 (2018).
    Article  Google Scholar 

    10.
    Young, R. C. et al. Age, sex, and telomere dynamics in a long-lived seabird with male-biased parental care. PLoS ONE 8, e74931. https://doi.org/10.1371/journal.pone.0074931 (2013).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    11.
    Sudyka, J., Arct, A., Drobniak, S. M., Gustafsson, L. & Cichon, M. Birds with high lifetime reproductive success experience increased telomere loss. Biol. Lett. 15, 20180637. https://doi.org/10.1098/rsbl.2018.0637 (2019).
    Article  PubMed  PubMed Central  Google Scholar 

    12.
    Parolini, M. et al. Telomere length is reflected by plumage coloration and predicts seasonal reproductive success in the barn swallow. Mol. Ecol. 26, 6100–6109. https://doi.org/10.1111/mec.14340 (2017).
    CAS  Article  PubMed  Google Scholar 

    13.
    Le Vaillant, M. et al. Telomere length reflects individual quality in free-living adult king penguins. Polar Biol. 38, 2059–2067. https://doi.org/10.1007/s00300-015-1766-0 (2015).
    Article  Google Scholar 

    14.
    Salomons, H. M. et al. Telomere shortening and survival in free-living corvids. Proc. R. Soc. B 276, 3157–3165. https://doi.org/10.1098/rspb.2009.0517 (2009).
    CAS  Article  PubMed  Google Scholar 

    15.
    Salmón, P., Nilsson, J. F., Watson, H., Bensch, S. & Isaksson, C. Selective disappearance of great tits with short telomeres in urban areas. Proc. R. Soc. B 284, 20171349. https://doi.org/10.1098/rspb.2017.1349 (2017).
    CAS  Article  PubMed  Google Scholar 

    16.
    Cerchiara, J. A. et al. Magellanic penguin telomeres do not shorten with age with increased reproductive effort, investment, and basal corticosterone. Ecol. Evol. 7, 5682–5691. https://doi.org/10.1002/ece3.3128 (2017).
    Article  PubMed  PubMed Central  Google Scholar 

    17.
    Aabye, I. Telomere length does not correlate with individual quality in a population of blue tits (Cyanistes caeruleus) Masters thesis, University of Oslo, (2017).

    18.
    Caprioli, M. et al. Nestling telomere length does not predict longevity, but covaries with adult body size in wild barn swallows. Biol. Lett. 9, 20130340. https://doi.org/10.1098/rsbl.2013.0340 (2013).
    Article  PubMed  PubMed Central  Google Scholar 

    19.
    Bauch, C., Becker, P. H. & Verhulst, S. Telomere length reflects phenotypic quality and costs of reproduction in a long-lived seabird. Proc. R. Soc. B 280, 20122540. https://doi.org/10.1098/rspb.2012.2540 (2013).
    Article  PubMed  Google Scholar 

    20.
    Monaghan, P. Telomeres and life histories: the long and the short of it. Ann. N. Y. Acad. Sci. 1206, 130–142. https://doi.org/10.1111/j.1749-6632.2010.05705.x (2010).
    ADS  Article  PubMed  Google Scholar 

    21.
    Haussmann, M. F. & Heidinger, B. J. Telomere dynamics may link stress exposure and ageing across generations. Biol. Lett. 11, 20150396. https://doi.org/10.1098/rsbl.2015.0396 (2015).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    22.
    Giraudeau, M., Angelier, F. & Sepp, T. Do telomeres influence pace-of-life-strategies in response to environmental conditions over a lifetime and between generations?. BioEssays https://doi.org/10.1002/bies.201800162 (2019).
    Article  PubMed  Google Scholar 

    23.
    Young, A. J. The role of telomeres in the mechanisms and evolution of life-history trade-offs and ageing. Philos. Trans. R. Soc. B 373, 20160452. https://doi.org/10.1098/rstb.2016.0452 (2018).
    CAS  Article  Google Scholar 

    24.
    Olsson, M., Wapstra, E. & Friesen, C. R. Evolutionary ecology of telomeres: a review. Ann. N. Y. Acad. Sci. 1422, 5–28. https://doi.org/10.1111/nyas.13443 (2018).
    ADS  Article  PubMed  Google Scholar 

    25.
    Schull, Q. et al. Assortative pairing by telomere length in king penguins and relationships with breeding success. Can. J. Zool. 96, 639–647. https://doi.org/10.1139/cjz-2017-0094 (2018).
    CAS  Article  Google Scholar 

    26.
    Belmaker, A. The role of telomere length in the life history and behavior of Tree Swallows (Tachycineta bicolor) PhD thesis, Cornell University, (2016).

    27.
    Johnsen, A., Pauliny, A., Lifjeld, J. T. & Blomqvist, D. Is telomere length associated with mate choice in a songbird with a high rate of extrapair paternity?. PLoS ONE 12, e0182446. https://doi.org/10.1371/journal.pone.0182446 (2017).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    28.
    Khoriauli, L. et al. Assortative mating for telomere length and antioxidant capacity in barn swallows (Hirundo rustica). Behav. Ecol. Sociobiol. 71, 124. https://doi.org/10.1007/s00265-017-2352-y (2017).
    Article  Google Scholar 

    29.
    Broer, L. et al. Meta-analysis of telomere length in 19 713 subjects reveals high heritability, stronger maternal inheritance and a paternal age effect. Eur. J. Hum. Genet. 21, 1163–1168. https://doi.org/10.1038/ejhg.2012.303 (2013).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    30.
    Burley, N. Parental investment, mate choice, and mate quality. Proc. Natl. Acad. Sci. USA 74, 3476–3479 (1977).
    ADS  CAS  Article  Google Scholar 

    31.
    Guerra, M. & Drummond, H. Reversed sexual size dimorphism and parental care: minimal division of labour in the blue-footed booby. Behaviour 132, 479–496 (1995).
    Article  Google Scholar 

    32.
    Torres, R. & Velando, A. A dynamic trait affects continuous pair assessment in the blue-footed booby Sula nebouxii. Behav. Ecol. Sociobiol. 55, 65–72. https://doi.org/10.1007/s00265-003-0669-1 (2003).
    Article  Google Scholar 

    33.
    Torres, R. & Velando, A. Male preference for female foot colour in the socially monogamous blue-footed booby Sula nebouxii. Anim. Behav. 69, 59–65. https://doi.org/10.1016/j.anbehav.2004.03.008 (2005).
    Article  Google Scholar 

    34.
    Sánchez-Macouzet, O. Monogamia serial y fidelidad en el bobo de patas azules PhD thesis, Universidad Nacional Autónoma de México, (2017).

    35.
    Drummond, H. & Rodríguez, C. Viability of booby offspring is maximized by having one young parent and one old parent. PLoS ONE 10, e0133213. https://doi.org/10.1371/journal.pone.0133213 (2015).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    36.
    Nelson, J. B. Pelicans, Cormorants, and their Relatives: The Pelecaniformes Vol. 661 (Oxford University Press, 2006).

    37.
    Torres, R. & Velando, A. Male reproductive senescence: the price of immune-induced oxidative damage on sexual attractiveness in the blue-footed booby. J. Anim. Ecol. 76, 1161–1168. https://doi.org/10.1111/j.1365-2656.2007.01282.x (2007).
    Article  PubMed  Google Scholar 

    38.
    Velando, A., Beamonte-Barrientos, R. & Torres, R. Pigment-based skin colour in the blue-footed booby: an honest signal of current condition used by females to adjust reproductive investment. Oecologia 149, 535–542. https://doi.org/10.1007/s00442-006-0457-5 (2006).
    ADS  Article  PubMed  Google Scholar 

    39.
    Torres, R. & Velando, A. in Advances in the Study of Behavior Vol. 42: Behavioral ecology of tropical animals (ed Regina Macedo) 155–188 (2010).

    40.
    Kiere, L. M. & Drummond, H. Extrapair behaviour reveals flexible female choosiness and mixed support for classic good genes in blue-footed boobies. Anim. Behav. 95, 145–153. https://doi.org/10.1016/j.anbehav.2014.07.007 (2014).
    Article  Google Scholar 

    41.
    Noguera, J. C., Metcalfe, N. B., Boner, W. & Monaghan, P. Sex-dependent effects of nutrition on telomere dynamics in zebra finches (Taeniopygia guttata). Biol. Lett. 11, 20140938. https://doi.org/10.1098/rsbl.2014.0938 (2015).
    Article  PubMed  PubMed Central  Google Scholar 

    42.
    Tarry-Adkins, J. L. & Ozanne, S. E. The impact of early nutrition on the aging trajectory. Proc. Nutr. Soc. 73, 289–301. https://doi.org/10.1017/S002966511300387X (2014).
    CAS  Article  PubMed  Google Scholar 

    43.
    Taff, C. C. & Freeman-Gallant, C. R. Sexual signals reflect telomere dynamics in a wild bird. Ecol. Evol. 7, 3436–3442. https://doi.org/10.1002/ece3.2948 (2017).
    Article  PubMed  PubMed Central  Google Scholar 

    44.
    Holveck, M.-J. & Riebel, K. Low-quality females prefer low-quality males when choosing a mate. Proc. R. Soc. B 277, 153–160. https://doi.org/10.1098/rspb.2009.1222 (2010).
    Article  PubMed  Google Scholar 

    45.
    Ratikainen, I. I. & Kokko, H. Differential allocation and compensation: Who deserves the silver spoon?. Behav. Ecol. 21, 195–200. https://doi.org/10.1093/beheco/arp168 (2010).
    Article  Google Scholar 

    46.
    Sheldon, B. C. Differential allocation: tests, mechanisms and implications. Trends Ecol. Evol. 15, 397–402 (2000).
    CAS  Article  Google Scholar 

    47.
    Lessells, C. K. M. & McNamara, J. M. Sexual conflict over parental investment in repeated bouts: negotiation reduces overall care. Proc. R. Soc. B 279, 1506–1514 (2012).
    CAS  Article  Google Scholar 

    48.
    Slagsvold, T. & Lifjeld, J. T. Hatching asynchrony in birds: the hypothesis of sexual conflict over parental investment. Am. Nat. 134, 239–253 (1989).
    Article  Google Scholar 

    49.
    Kokko, H. Good genes, old age and life-history trade-offs. Evol. Ecol. 12, 739–750 (1998).
    Article  Google Scholar 

    50.
    Kokko, H. & Lindstrom, J. Evolution of female preference for old mates. Proc. R. Soc. B 263, 1533–1538. https://doi.org/10.1098/rspb.1996.0224 (1996).
    ADS  Article  Google Scholar 

    51.
    Ramos, A. G. et al. Interactive effects of male and female age on extra-pair paternity in a socially monogamous seabird. Behav. Ecol. Sociobiol. https://doi.org/10.1007/s00265-014-1769-9 (2014).
    Article  Google Scholar 

    52.
    Drummond, H., Torres, R. & Krishnan, V. V. Buffered development: resilience after aggressive subordination in infancy. Am. Nat. 161, 794–807. https://doi.org/10.1086/375170 (2003).
    Article  PubMed  Google Scholar 

    53.
    Beamonte-Barrientos, R., Velando, A., Drummond, H. & Torres, R. Senescence of maternal effects: aging influences egg quality and rearing capacities of a long-lived bird. Am. Nat. https://doi.org/10.1086/650726 (2010).
    Article  PubMed  Google Scholar 

    54.
    Kim, S.-Y., Torres, R., Domínguez, C. A. & Drummond, H. Lifetime philopatry in the blue-footed booby: a longitudinal study. Behav. Ecol. 18, 1132–1138. https://doi.org/10.1093/beheco/arm091 (2007).
    Article  Google Scholar 

    55.
    Osorio-Beristain, M. & Drummond, H. Natal dispersal and deferred breeding in the blue-footed booby. Auk 110, 234–239 (1993).
    Google Scholar 

    56.
    Ramos, A. G. et al. Habitat structure and colony structure constrain extrapair paternity in a colonial bird. Anim. Behav. 95, 121–127. https://doi.org/10.1016/j.anbehav.2014.07.003 (2014).
    Article  Google Scholar 

    57.
    Gaunt, A. S. et al. Guidelines to the use of wild birds in research (The Ornithological Counciil, 1999).

    58.
    Merkling, T. et al. Food availability and offspring sex in a monogamous seabird: insights from an experimental approach. Behav. Ecol. 23, 751–758. https://doi.org/10.1093/beheco/ars023 (2012).
    Article  Google Scholar 

    59.
    Cawthon, R. M. Telomere measurement by quantitative PCR. Nucleic Acids Res. https://doi.org/10.1093/nar/30.10.e47 (2002).
    Article  PubMed  PubMed Central  Google Scholar 

    60.
    Foote, C. G. Avian Telomere Dynamics Ph.D. thesis, University of Glasgow, (2008).

    61.
    Morinha, F., Magalhães, P. & Blanco, G. Standard guidelines for the publication of telomere qPCR results in evolutionary ecology. Mol. Ecol. Resour. https://doi.org/10.1111/1755-0998.13152 (in press).

    62.
    Pfaffl, M. W. A new mathematical model for relative quantification in real-time RT-PCR. Nucl. Acids Res. 29, e45 (2001).
    CAS  Article  Google Scholar 

    63.
    R: A language and environment for statistical computing v. 3.5.0 (Vienna, Austria, 2020).

    64.
    nlme: Linear and Nonlinear Mixed Effects Models (2011).

    65.
    Verhulst, S. Improving comparability between qPCR-based telomere studies. Mol. Ecol. Resour. 20, 11–13. https://doi.org/10.1111/1755-0998.13114 (2020).
    CAS  Article  PubMed  Google Scholar  More

  • in

    Non-responder phenotype reveals apparent microbiome-wide antibiotic tolerance in the murine gut

    Antibiotic duration experiment
    Twenty-eight-week-old female C57BL/6J mice from the same birth cohort were co-housed (5–6 mice per cage) prior to beginning the experiment, and then separated into individual cages 1 week prior to antibiotic treatment. Singly housed mice were exposed to 0.5 mg/mL33 cefoperazone in their drinking water for 0, 2, 4, 8, or 16 days (Fig. 1A). Based on the literature, we calculated the minimum dose of cefoperazone based on the mean and standard deviation of water consumption by C57BL/6J mice ((7.7, mp 0.3,{mathrm{mL}}) per 30 g of body weight)44. If the heaviest mouse in our study (~22 g) consistently consumed water at 2 SD below the mean (i.e. 5.5 of 0.5 mg/mL cefoperazone), they would still receive 125 mg/kg/day of cefoperazone, which is within the therapeutic dosing range for humans (100–150 mg/kg/day; although cefoperazone is administered to humans via intravenous injection)45.
    Fig. 1: Effect of antibiotic exposure duration on non-responder phenotype.

    The table in the center denotes the number of non-responder and responder mice in each treatment duration group. A Experimental design for the duration experiment. Circles denote sampled time points. Time points were considered sampled “during” antibiotic treatment between day 0 and day 2, 4, 8, and 16, respectively, as denoted by orange shades. B Relative abundance of phyla on the last day of antibiotics treatment. The control panel is an average over all untreated controls from all time points. Only phyla with a relative abundance of at least 0.1% are shown. Each barchart denotes means from at least two samples and white insets are the sample size used for each barchart. C Percentage of mitochondria and chloroplast sequences in 16S amplicon data relative to antibiotic treatment. Colors: red—controls not treated with antibiotics, green—non-responders, blue—responders. D Principal coordinate analysis (PCoA) of samples during and after antibiotic exposure (n = 143 samples with >10,000 reads per sample, day ≥ 0). Ellipses denote 95% confidence intervals from a Student t-distribution. Each point denotes a sample. ASV abundances were rarefied to 10,000 reads for each sample and percentages in brackets denote the explained variance. Samples with less than 10,000 reads per sample were not included in the analysis. E Dynamics of amplicon sequence variants (ASVs). Gained ASVs are variants that were not present before antibiotics treatment but are present after. Similarly, lost ASVs were present before treatment but not after, and persistent ASVs were present before and after. Stars denote significance under a Mann–Whitney U test: *p  More

  • in

    Amplification of potential thermogenetic mechanisms in cetacean brains compared to artiodactyl brains

    Specimens
    We used brains obtained from three cetacean species (harbour porpoise—Phocoena phocoena, minke whale—Balaenoptera acutorostrata, and humpback whale—Megaptera novaeangliae) and 11 artiodactyl species (sand gazelle—Gazella marica, domestic pig—Sus scrofa, Nubian ibex—Capra nubiana, springbok—Antidorcas marsupialis, blesbok—Damaliscus pygargus, greater kudu—Tragelaphus strepsiceros, blue wildebeest—Connochaetes taurinus, dromedary camel—Camelus dromedarius, nyala—Tragelaphus angasii, river hippopotamus—Hippopotamus amphibius, and African buffalo—Syncerus caffer) (Table 1). All artiodactyl brains were perfusion fixed with 4% paraformaldehyde in 0.1 M phosphate buffer through the carotid arteries following euthanasia45. The harbour porpoise specimens were perfusion fixed through the heart following euthanasia, while the minke whale and humpback whale brains were immersion fixed in 4% paraformaldehyde in 0.1 M phosphate buffer. All brains were then stored in an antifreeze solution at – 20 °C until use45. All specimens were taken under appropriate governmental permissions, with ethical clearance provided by the University of the Witwatersrand Animal Ethics Committee (Clearance number 2008/36/1), which uses guidelines similar to those of the National Institutes of Health regarding the use of animals in scientific research and is compliant with ARRIVE guidelines.
    Immunohistochemical staining
    Blocks of tissue from the anterior cingulate (dorsal to the rostrum of the corpus callosum, in all species apart from the humpback whale where we did not have this tissue block) and occipital cortex (presumably primary visual cortex, from all species) with underlying white matter were taken from each of the specimens. These were placed in a 30% sucrose in 0.1 M phosphate buffer solution at 4 °C until equilibrated. The blocks were frozen in crushed dry ice, mounted on an aluminium stage and sectioned at 50 µm orthogonal to the pial surface. Alternate sections were stained for Nissl (with 1% cresyl violet), UCP1, UCP2, UCP3, UCP4, UCP5, dopamine-ß-hydroxylase (DBH) and tyrosine hydroxylase (TH). To investigate the presence of neural structures immunolocalizing uncoupling proteins, DBH and TH, we used standard immunohistochemical procedures with antibodies directed against UCP1, UCP2, UCP3, UCP4, UCP5, DBH and TH. While immunolocalization for UCP1, UCP4, UCP5, DBH and TH were clear, only occasional cortical neurons were immunopositive for UCP2, and no immunolocalization could be detected for UCP3 in the species studied. It should be noted here that immunostaining for DBH and TH did not work in the humpback whale specimen, perhaps due to the fixation procedure or the conformation of the targeted proteins in this species preventing recognition of the DBH and TH proteins by the antibodies used. Sections used for the Nissl series were mounted on 0.5% gelatine-coated glass slides, cleared in a solution of 1:1 chloroform and absolute alcohol, then stained with 1% cresyl violet to reveal cell bodies. For the immunohistochemical staining, each section was treated with endogenous peroxidase inhibitor (49.2% methanol:49.2% 0.1 M PB:1.6% of 30% H2O2) for 30 min and subsequently subjected to three 10 min 0.1 M PB rinses. Sections were then incubated for 2 h, at room temperature, in blocking buffer (containing 3% normal rabbit serum, NRS, for the UCP1-5 sections/3% normal horse serum, NHS, for the DBH sections/3% normal goat serum, NGS, for the TH sections, plus 2% bovine serum albumin and 0.25% Triton-X in 0.1 M PB). This was followed by three 10 min rinses in 0.1 M PB. The sections were then placed in the primary antibody solution that contained the appropriately diluted primary antibody in blocking buffer for 48 h at 4°C under gentle shacking. The optimal dilutions for the UCP primary antibodies were determined with a series of stains in which the dilution of the primary antibodies ranged from 1:300 through to 1:9600, with any staining in all species being absent at a dilution of 1:4800 irrespective of fixation method. We used antibodies directed against UCP1 (Santa Cruz Biotechnology, C-17, sc-6528, Lot# D0411, goat polyclonal IgG, dilution 1:300, RRID:AB_2304265), UCP2 (Santa Cruz Biotechnology, C-20, sc-6525, Lot# E0211, goat polyclonal IgG, dilution 1:300, RRID:AB_2213585), UCP3 (Santa Cruz Biotechnology, C-20, sc-7756, Lot# A2511, goat polyclonal IgG, dilution 1:300, RRID:AB_2213922), UCP4 (Santa Cruz Biotechnology, N-16, sc-17582, Lot# E2004, goat polyclonal IgG, dilution 1:300, RRID:AB_793648), UCP5 (Santa Cruz Biotechnology, Q-16, sc-50540, Lot# B1207, goat polyclonal IgG, dilution 1:300, RRID:AB_2286101), DBH (Merck-Millipore, MAB308, mouse monoclonal IgG, dilution 1:4000, RRID:AB_2245740) and TH (Merck-Millipore, AB151, rabbit polyclonal IgG, dilution 1:3000, RRID:AB_10000323). This incubation was followed by three 10 min rinses in 0.1 M PB and the sections were then incubated in a secondary antibody solution (1:1000 dilution of biotinylated anti-goat IgG, BA-5000, Vector Labs, for UCP1-5 sections/1:1000 dilution of biotinylated anti-mouse IgG, BA 2001, Vector labs, for DBH sections/1:1000 dilution of biotinylated anti-rabbit IgG, BA-1000, Vector Labs, for TH sections, in a blocking buffer containing 3% NRS/NHS/NGS and 2% BSA in 0.1 M PB) for 2 h at room temperature. This was followed by three 10 min rinses in 0.1 M PB, after which sections were incubated for 1 h in avidin-biotin solution (at a dilution of 1:125, Vector Labs), followed by three 10 min rinses in 0.1 M PB. Sections were then placed in a solution of 0.05% 3,3′-diaminobenzidine (DAB) in 0.1 M PB for 5 min, followed by the addition of 3 ml of 3% hydrogen peroxide to each 1 ml of solution in which each section was immersed. Chromatic precipitation was visually monitored and verified under a low power stereomicroscope. Staining was allowed to continue until such time as the background stain was at a level that would assist architectural reconstruction and matching without obscuring the immunopositive neurons. Development was halted by placing the sections in 0.1 M PB, followed by two more rinses in 0.1M PB. To test for non-specific staining of the immunohistochemical protocol, in selected sections the primary antibody or the secondary antibody were omitted, which resulted in no staining of the tissue. The immunostained sections were then mounted on 0.5% gelatine coated glass slides, dried overnight, dehydrated in a graded series of alcohols, cleared in xylene and coverslipped with Depex. Digital photomicrographs were captured using Zeiss Axioshop and Axiovision software. No pixilation adjustments, or manipulation of the captured images were undertaken, except for the adjustment of contrast, brightness, and levels using Adobe Photoshop 7.
    Western immunoblotting
    Protein expression for UCP1 and UCP4 was assayed using standard qualitative Western immunoblotting techniques. To verify the specificity of the UCP1 antibody for the UCP1 protein, we tested this antibody with rat brown fat. For the UCP4 antibody protein samples were extracted from the paraformaldehyde fixed tissue using the Qproteome FFPE Tissue Kit (Qiagen, Germany). The tissue blocks analysed here were taken from the anterior cingulate and occipital cortex (as described above) and contained both gray and white matter. 30–40 mg of the sample were incubated in 100 µl of Extraction Buffer EXB Plus (Qiagen, Germany) containing 6% β-mercaptoethanol on ice for 5 min and mixed by vortexing. The samples were boiled for 20 min at 100°C and subsequently incubated at 50°C overnight with agitation at 300 rpm. The samples were then placed on ice for 1 min and centrifuged for 15 min at 14 000g at 4°C. The supernatant was transferred into clean tubes and the protein concentration was determined using the Bradford protein assay kit (Bio-Rad Laboratories, USA). The protein extracts (20 µg) were made soluble in sample buffer comprised of 0.0625 M Tris–HCl, pH 6.8, 10% glycerol, 2% SDS, 2.5% β-mercaptoethanol and 0.001% bromophenol blue, boiled at 95°C for 5 min and subjected to 12% SDS-polyacrylamide gel electrophoresis and transferred to polyvinylidene difluoride (PVDF) (Millipore) at 20 V/cm for 1h. Electrophoresis and protein transfer was achieved using Mini Trans-Blot Electrophoretic Transfer Cell (Bio-Rad Laboratories, Inc. USA). After the transfer the blots were blocked for 2 h in 1 × Animal-Free Blocker (SP-5030 Vector Labs, USA). The blots were incubated over night at 4°C under gentle agitation in the primary antibody solutions (1:300 goat anti-UCP1, Santa Cruz Biotechnology, sc-6528 or 1:300 goat anti-UCP4, Santa Cruz Biotechnology, sc-17582). The blots were washed for 3 × 10 min in 1 × Animal-Free Blocker and incubated for 1 h at room temperature in HRP-conjugated rabbit anti-goat secondary antibody (1:1000, Dako, USA) for 1 h. This was followed by 3 × 10 min washes with 50 mM Tris buffer, pH 7.2. The protein bands were detected using 3,3′-diaminobenzidine tetrahydrochloride hydrate (DAB) (Sigma, D5637). The blots were incubated in a solution containing 1mg/ml DAB in 50 mM Tris, pH 7.2 for 5 min at room temperature, followed by the addition of an equal amount of 0.02% hydrogen peroxide solution. Development was arrested by placing the blots in 50 mM Tris (pH 7.2) for 10 min, followed by two more 10 min rinses in distilled water.
    Stereological analysis
    Using a design-based stereological approach we analysed immunohistochemically stained sections in the grey matter of the anterior cingulate and occipital cortex, as well as the underlying white matter from these regions of 14 cetartiodactyl species. Regions of interest (ROI) were drawn from similar locations across species as supported by published anatomical descriptions of the cetacean and artiodactyl brain. Using a light microscope equipped with a motorized stage, digital camera, MicroBrightfield system (MBF Bioscience, USA) system and StereoInvestigator software (MBF Bioscience, version 2018.1.1; 64-bit), we quantified UCP1-immunoreactive neuron densities in the grey matter, UCP4-immunoreactive glia densities in the grey and white matter, and DBH- and TH-immunoreactive bouton densities in the grey and white matter of these cortical regions. Separate pilot studies for each immunohistochemical stain was conducted to optimise sampling parameters, such as the counting frame and sampling grid sizes, and achieve a coefficient of error (CE) below 0.127,46,47,48,49. In addition, we measured the tissue section thickness at every sampling site, and the vertical guard zone was determined according to tissue thickness to avoid errors/biases due to sectioning artefacts27,46,47,48,49. Supplementary Tables S1–S4 provide details of the parameters used for each neuroanatomical region and stain and between the species in the current study. To estimate the ROI total number, we used the ‘Optical Fractionator’ probe.
    UCP1- and UCP4-immunoreactive neuron and glia densities were obtained by sampling the cortical areas of interest and subjacent white matter with the aid of an optical disector. The cortex and white matter were outlined separately at low magnification (2X), and the optical disector was performed at 40X. UCP-immunoreactive neuron and glia density was calculated as the total number of UCP-immunoreactive neurons and glia divided by the product of surface area (x, y), the tissue sampling fraction, and the sectioned thickness (50 µm). The tissue sampling fraction was calculated as the ratio of the optical disector height to mean measured section thickness. Given that overall cell density per unit volume is known to vary with differences in brain size, we calculated the percentage of UCP-immunoreactive neurons or glia, expressed as the ratio of UCP-immunoreactive neurons or glia to total neuronal or glial density for each region of interest, to standardize the data for cross species comparison. Using Nissl-stained sections we obtained estimates of neuronal and glial densities within the cortex and glial density within the white matter using optical disector probes combined with a fractionator sampling scheme46. A pilot study determined the optimal sampling parameters and grid dimensions to place disector frames in a systematic-random manner. For DBH and TH bouton densities, ‘spot’ densities were calculated by multiplying the ROI area by the cut section thickness, and then using the generated volume as the denominator to the ROI estimated number. For all tissue sampled the optical fractionator was used while maintaining strict criteria, e.g. only complete boutons were counted, 63 X oil immersion, and obeying all commonly known stereological rules. The stereologic analyses presented here resulted in sampling an average of 118 counting frames per region of interest with a total of 13,053 counting frames investigated.
    Statistical analyses
    We hypothesized that the percentage of cortical neurons immunoreactive to UCP1 were significantly different between artiodactyls and cetaceans. To test this hypothesis, we compared the proportion of UCP1 expression in the anterior cingulate and occipital cortex of 16 cetartiodactyls. For the anterior cingulate cortex, we sampled a total of 1109 sampling sites (~ 100 sites per species) within the artiodactyl group and found that 36.83% of sampled cortical neurons were immunoreactive to UCP1. In comparison our cetacean sample consisted of 723 sampling sites (~ 145 sites per species), with 87.28% of the sampled cortical neurons immunoreactive to UCP1. For the occipital cortex, we sampled a total of 1 038 sites (~ 94 sites per species) within the artiodactyl group and found that 34% of sampled cortical neurons within the occipital cortex were immunoreactive to UCP1. The cetacean sample consisted of 723 sampling sites (~ 145 sites per species), and we found that 92.36% of the sampled cortical neurons were immunoreactive to UCP1.
    To test if the respective underlying proportions were different between the sample groups, we conducted statistical hypothesis testing using the Two-Proportions Z-test as implemented in the R Programming language. Our Null hypothesis (Ho) stated that there is no significant difference between the proportions of artiodactyl immunoreactive UCP1 sampled cortical neurons (π1) and the proportions of cetacean UCP1 sampled cortical neurons (π2)—that is, π1 − π2 = 0. The alternate hypothesis (H1) stated that there is a significant difference in these proportions such that π1 − π2 ≠ 0, with one of the proportions being either less than or greater than the other. We thus conducted a two-sided hypothesis test, with the significance level (α) set at 0.05 (i.e., P-values less than, or equal to, α, would reject the null hypothesis in favour of the alternate hypothesis). Based on these analyses the proportion of immunolabelled UCP1 cortical neurons were found to be significantly different between the groups, with cetaceans having a significantly higher proportion of UCP1-immunoreactive neurons in the anterior cingulate cortex (χ2 = 51.69; df =1, P = 6.49 × 10−13, 95% confidence interval = − 0.122; − 0.067) and occipital cortex (χ2 = 56.30; P = 6.21 × 10−14, 95% confidence interval = − 0.114; − 0.060).
    We used a two sample T-test (as implemented in R) to test for significant differences in noradrenergic bouton density between cetaceans and artiodactyls. Cetaceans were found to have significantly higher mean DBH-immunoreactive bouton densities in the anterior cingulate cortex as compared to artiodactyls (t = − 3.595; df =15, P = 0.011). Cetaceans were also found to have significantly higher mean DBH-immunoreactive bouton densities in the occipital cortex as compared to artiodactyls (t = − 4.546; df =15, P = 0.002). Similarly, we tested for significant differences in mean DBH bouton density in the underlying cortical white matter of cetaceans and artiodactyls. We did not find any significant differences in DBH-immunoreactive bouton density for the anterior cingulate (t =− 0.597; df =15, P = 0.585) or occipital cortex (t = − 0.08; df =15, P = 0.941).
    To test for the effect of confounding variables on the significant differences observed in DBH bouton density in the cortex, we used an analysis of covariance controlling sequentially for the effect of cortical neuron density, cortical glia density and brain mass. Our analyses revealed that after adjusting for the density of cortical neurons cetaceans still had significantly higher DBH-immunoreactive bouton density in the anterior cingulate cortex (adjusted mean = 10.176) in comparison to artiodactyls (adjusted mean = 8.176) (F = 5.222; df =13, P = 0.041). Adjusting for the covariate cortical neuron density, resulted in a similar result for the occipital cortex (adjusted mean = 14.678) in comparison to artiodactyls (adjusted mean = 10.395) (F = 14.05; df =13, P = 0.00278). When controlling for the density of cortical glia, cetaceans also had significantly higher DBH-immunoreactive bouton densities in the anterior cingulate cortex (adjusted mean = 10.62) in comparison to artiodactyls (adjusted mean = 8.01) (F = 9.72; df =13, P = 0.00889). Similar results were found for the occipital cortex, with cetaceans having significantly higher DBH-immunoreactive bouton density (adjusted mean = 14.471) compared to artiodactyls (adjusted mean = 10.395) (F = 11.2; df =13, P = 0.00581). When controlling for brain mass, cetaceans were also found to have a significantly higher DBH-immunoreactive bouton densities in the anterior cingulate (adjusted mean = 11.36) in comparison to artiodactyls (adjusted mean = 7.75) (F = 11.06; df = 13, P = 0.00604) as well as in the occipital cortex (cetacean adjusted mean = 15.406, artiodactyls adjusted mean = 10.055) (F = 11.85; df = 13, P = 0.00488). More

  • in

    Climate and seasonality drive the richness and composition of tropical fungal endophytes at a landscape scale

    1.
    Chesson, P. Mechanisms of maintenance of species diversity. Annu. Rev. Ecol. Syst. 31, 343–366 (2000).
    Article  Google Scholar 
    2.
    Carson, W. & Schnitzer, S. Tropical Forest Community Ecology (John Wiley & Sons, 2011).

    3.
    Givnish, T. J. On the causes of gradients in tropical tree diversity. J. Ecol. 87, 193–210 (1999).
    Article  Google Scholar 

    4.
    Condit, R., Engelbrecht, B. M. J., Pino, D., Pérez, R. & Turner, B. L. Species distributions in response to individual soil nutrients and seasonal drought across a community of tropical trees. Proc. Natl Acad. Sci. USA 110, 5064–5068 (2013).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    5.
    Bagchi, R. et al. Pathogens and insect herbivores drive rainforest plant diversity and composition. Nature 506, 85–88 (2014).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    6.
    Zalamea, P.-C. et al. Seedling growth responses to phosphorus reflect adult distribution patterns of tropical trees. New Phytol. 212, 400–408 (2016).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    7.
    Sarmiento, C. et al. Soilborne fungi have host affinity and host-specific effects on seed germination and survival in a lowland tropical forest. Proc. Natl Acad. Sci. USA 114, 11458–11463 (2017).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    8.
    Ter Steege, H., Pitman, N. & Sabatier, D. A spatial model of tree alpha-diversity and tree density for the Amazon. Biodivers. Conserv. 12, 2255–2277 (2003).
    Article  Google Scholar 

    9.
    Leigh, E. G. Jr. et al. Why do some tropical forests have so many species of trees? Biotropica 36, 447–473 (2004).
    Google Scholar 

    10.
    Rahbek, C. & Graves, G. R. Detection of macro-ecological patterns in South American hummingbirds is affected by spatial scale. Proc. R. Soc. Lond. B. 267, 2259–2265 (2000).
    CAS  Article  Google Scholar 

    11.
    Abrahamczyk, S., Kluge, J., Gareca, Y., Reichle, S. & Kessler, M. The influence of climatic seasonality on the diversity of different tropical pollinator groups. PLoS ONE 6, e27115 (2011).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    12.
    Tedersoo, L. et al. Global diversity and geography of soil fungi. Science 346, 1256688 (2014).
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    13.
    Tonkin, J. D., Bogan, M. T., Bonada, N., Rios-Touma, B. & Lytle, D. A. Seasonality and predictability shape temporal species diversity. Ecology 98, 1201–1216 (2017).
    PubMed  Article  PubMed Central  Google Scholar 

    14.
    Doležal, J., Lanta, V., Mudrák, O. & Lepš, J. Seasonality promotes grassland diversity: interactions with mowing, fertilization and removal of dominant species. J. Ecol. 107, 203–215 (2019).

    15.
    Arnold, A. E. et al. Fungal endophytes limit pathogen damage in a tropical tree. Proc. Natl Acad. Sci. USA 100, 15649–15654 (2003).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    16.
    Arnold, A. E. & Engelbrecht, B. M. J. Fungal endophytes nearly double minimum leaf conductance in seedlings of a neotropical tree species. J. Trop. Ecol. 23, 369–372 (2007).
    Article  Google Scholar 

    17.
    Costa Pinto, L. S., Azevedo, J. L., Pereira, J. O., Carneiro Vieira, M. L. & Labate, C. A. Symptomless infection of banana and maize by endophytic fungi impairs photosynthetic efficiency. New Phytol. 147, 609–615 (2000).
    Article  Google Scholar 

    18.
    U’Ren, J. M. et al. Diversity and evolutionary origins of fungi associated with seeds of a neotropical pioneer tree: a case study for analysing fungal environmental samples. Mycol. Res. 113, 432–449 (2009).
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    19.
    Sanchez-Azofeifa, A., Oki, Y., Wilson Fernandes, G., Ball, R. A. & Gamon, J. Relationships between endophyte diversity and leaf optical properties. Trees 26, 291–299 (2012).
    Article  Google Scholar 

    20.
    Vincent, J. B., Weiblen, G. D. & May, G. Host associations and beta diversity of fungal endophyte communities in New Guinea rainforest trees. Mol. Ecol. 25, 825–841 (2016).
    CAS  PubMed  Article  Google Scholar 

    21.
    Suryanarayanan, T. S., Murali, T. S. & Venkatesan, G. Occurrence and distribution of fungal endophytes in tropical forests across a rainfall gradient. Can. J. Bot. 80, 818–826 (2002).
    Article  Google Scholar 

    22.
    Zimmerman, N. B. & Vitousek, P. M. Fungal endophyte communities reflect environmental structuring across a Hawaiian landscape. Proc. Natl Acad. Sci. USA 109, 13022–13027 (2012).
    CAS  PubMed  Article  Google Scholar 

    23.
    Higgins, K. L., Arnold, A. E., Coley, P. D. & Kursar, T. A. Communities of fungal endophytes in tropical forest grasses: highly diverse host- and habitat generalists characterized by strong spatial structure. Fungal Ecol. 8, 1–11 (2014).
    Article  Google Scholar 

    24.
    Darcy, J. L. et al. Fungal communities living within leaves of native Hawaiian dicots are structured by landscape-scale variables as well as by host plants. Mol. Ecol. 29, 3102–3115 (2020).
    Article  Google Scholar 

    25.
    Arnold, A. E. & Lutzoni, F. Diversity and host range of foliar fungal endophytes: are tropical leaves biodiversity hotspots? Ecology 88, 541–549 (2007).
    PubMed  Article  Google Scholar 

    26.
    Tellez, P. H. Tropical plants and fungal symbionts: Leaf functional traits as drivers of plant-fungal interactions. PhD dissertation (Tulane University, 2019).

    27.
    U’Ren, J. M. et al. Host availability drives distributions of fungal endophytes in the imperilled boreal realm. Nat. Ecol. Evol. 3, 1430–1437 (2019).
    PubMed  Article  Google Scholar 

    28.
    Arnold, A. E. & Herre, E. A. Canopy cover and leaf age affect colonization by tropical fungal endophytes: ecological pattern and process in Theobroma cacao (Malvaceae). Mycologia 95, 388–398 (2003).
    PubMed  Article  PubMed Central  Google Scholar 

    29.
    Rodríguez-Quiel, E. E., Mendieta-Leiva, G. & Bader, M. Y. Elevational patterns of bryophyte and lichen biomass differ among substrates in the tropical montane forest of Baru Volcano, Panama. J. Bryol. 41, 95–106 (2019).
    Article  Google Scholar 

    30.
    Magill, B., Solomon, J. & Stimmel, H. Tropicos Specimen Data. http://www.tropicos.org (2019).

    31.
    Arnold, A. E. et al. A phylogenetic estimation of trophic transition networks for ascomycetous fungi: are lichens cradles of symbiotrophic fungal diversification? Syst. Biol. 58, 283–297 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    32.
    U’Ren, J. M., Lutzoni, F., Miadlikowska, J., Laetsch, A. D. & Arnold, A. E. Host and geographic structure of endophytic and endolichenic fungi at a continental scale. Am. J. Bot. 99, 898–914 (2012).
    PubMed  Article  PubMed Central  Google Scholar 

    33.
    Arnold, A. E., Maynard, Z., Gilbert, G. S., Coley, P. D. & Kursar, T. A. Are tropical fungal endophytes hyperdiverse? Ecol. Lett. 3, 267–274 (2000).
    Article  Google Scholar 

    34.
    Phillips, O. L., Hall, P., Gentry, A. H., Sawyer, S. A. & Vásquez, R. Dynamics and species richness of tropical rain forests. Proc. Natl Acad. Sci. USA 91, 2805–2809 (1994).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    35.
    Usinowicz, J. et al. Temporal coexistence mechanisms contribute to the latitudinal gradient in forest diversity. Nature 550, 105–108 (2017).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    36.
    Pianka, E. R. Latitudinal gradients in species diversity: a review of concepts. Am. Nat. 100, 33–46 (1966).
    Article  Google Scholar 

    37.
    Coley, P. D. & Barone, J. A. Herbivory and plant defenses in tropical forests. Annu. Rev. Ecol. Syst. 27, 305–335 (1996).
    Article  Google Scholar 

    38.
    Thrall, P. H., Hochberg, M. E., Burdon, J. J. & Bever, J. D. Coevolution of symbiotic mutualists and parasites in a community context. Trends Ecol. Evol. 22, 120–126 (2007).
    PubMed  Article  PubMed Central  Google Scholar 

    39.
    Poisot, T., Bever, J. D., Nemri, A., Thrall, P. H. & Hochberg, M. E. A conceptual framework for the evolution of ecological specialisation. Ecol. Lett. 14, 841–851 (2011).
    PubMed  PubMed Central  Article  Google Scholar 

    40.
    Van Bael, S., Estrada, C. & Arnold, A. E. Chapter 6: foliar endophyte communities and leaf traits in tropical trees. In The Fungal Community: Its Organization and Role in the Ecosystem. (eds Dighton, J. & White, J. F.) 79–94 (CRC Press, 2017).

    41.
    Oono, R. et al. Species diversity of fungal endophytes across a stress gradient for plants. New Phytol. 228, 210–225 (2020).

    42.
    Top, S. M., Preston, C. M., Dukes, J. S. & Tharayil, N. Climate influences the content and chemical composition of foliar tannins in green and senesced tissues of Quercus rubra. Front. Plant Sci. 8, 423 (2017).
    PubMed  PubMed Central  Article  Google Scholar 

    43.
    Higginbotham, S. J. et al. Bioactivity of fungal endophytes as a function of endophyte taxonomy and the taxonomy and distribution of their host plants. PLoS ONE 8, e73192 (2013).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    44.
    Hijmans, R. J., Cameron, S. E., Parra, J. L., Jones, P. G. & Jarvis, A. Very high resolution interpolated climate surfaces for global land areas. Int. J. Climatol. 25, 1965–1978 (2005).
    Article  Google Scholar 

    45.
    Olson, D. M. et al. Terrestrial ecoregions of the world: a new map of life on earth: a new global map of terrestrial ecoregions provides an innovative tool for conserving biodiversity. Bioscience 51, 933–938 (2001).
    Article  Google Scholar 

    46.
    Prada, C. M. et al. Soils and rainfall drive landscape-scale changes in the diversity and functional composition of tree communities in premontane tropical forest. J. Veg. Sci. 28, 859–870 (2017).
    Article  Google Scholar 

    47.
    Walker, K. Capturing ephemeral forest dynamics with hybrid time-series and composite mapping in the Republic of Panama. Int. J. Appl. Earth Obs. Geoinf. 87, 102029 (2020).
    Article  Google Scholar 

    48.
    Leung, B., Hudgins, E. J., Potapova, A. & Ruiz-Jaen, M. C. A new baseline for countrywide α-diversity and species distributions: illustration using > 6,000 plant species in Panama. Ecol. Appl. 29, e01866 (2019).
    PubMed  Article  Google Scholar 

    49.
    Pyke, C. R., Condit, R., Aguilar, S. & Lao, S. Floristic composition across a climatic gradient in a neotropical lowland forest. J. Veg. Sci. 12, 553–566 (2001).
    Article  Google Scholar 

    50.
    Lieberman, D., Lieberman, M., Peralta, R. & Hartshorn, G. S. Tropical forest structure and composition on a large-scale altitudinal gradient in Costa Rica. J. Ecol. 84, 137–152 (1996).
    Article  Google Scholar 

    51.
    Bowman, E. A. & Arnold, A. E. Distributions of ectomycorrhizal and foliar endophytic fungal communities associated with Pinus ponderosa along a spatially constrained elevation gradient. Am. J. Bot. 105, 687–699 (2018).
    PubMed  Article  PubMed Central  Google Scholar 

    52.
    U’Ren, J. M. et al. Tissue storage and primer selection influence pyrosequencing-based inferences of diversity and community composition of endolichenic and endophytic fungi. Mol. Ecol. Resour. 14, 1032–1048 (2014).
    PubMed  PubMed Central  Google Scholar 

    53.
    U’Ren, J. M. & Arnold, A. E. 96 well DNA extraction protocol for plant and lichen tissue stored in CTAB. protocols.io. https://doi.org/10.17504/protocols.io.fscbnaw (2017).

    54.
    Daru, B. H., Bowman, E. A., Pfister, D. H. & Arnold, A. E. A novel proof of concept for capturing the diversity of endophytic fungi preserved in herbarium specimens. Philos. Trans. R. Soc. Lond. B Biol. Sci. 374, 20170395 (2018).

    55.
    Callahan, B. J., McMurdie, P. J. & Holmes, S. P. Exact sequence variants should replace operational taxonomic units in marker-gene data analysis. ISME J. 11, 2639–2643 (2017).
    PubMed  PubMed Central  Article  Google Scholar 

    56.
    Oksanen, J. et al. Vegan: community ecology package, version 2.5-2. https://CRAN.R-project.org/package=vegan (2018).

    57.
    Schoener, T. W. Food webs from the small to the large: the Robert H. MacArthur award lecture. Ecology 70, 1559–1589 (1989).
    Article  Google Scholar 

    58.
    Apigo, A. & Oono, R. Dimensions of host specificity in foliar fungal endophytes. In Endophytes of Forest Trees: Biology and Applications (eds Pirttilä, A. M. & Frank, A. C.) 15–42 (Springer International Publishing, 2018).

    59.
    Oita, S. et al. Data from: climate and seasonality drive the richness and composition of tropical fungal endophytes at a landscape scale. figshare https://doi.org/10.6084/m9.figshare.c.5084366.v1 (2020). More

  • in

    Neoptile feathers contribute to outline concealment of precocial chicks

    Experiment 1: proof of principle
    As a proof of principle, we designed the first experiment to test whether appendages may help to conceal the outline. We created an image of a uniformly light grey coloured circular object with a size of 2950 pixels (px)/250.0 mm circumference and 470 px/39.8 mm radius on a dark grey background using Adobe InDesign CS6 version 8.030. The initial setup started with no appendages added to the outline (Fig. 3a, ‘0’). We then added object-coloured appendages (i.e. lines of 1 Pt/4 px/0.4 mm thickness and 118 px/10.0 mm length) with regular intervals resembling protruding neoptile chick feathers orthogonally to the object outline (‘Basic Scenario’, Fig. 3a). The first image with appendages had 32 appendages added to the outline (Fig. 3a, ‘32′). We then doubled the number of appendages stepwise creating denser spaced appendages to the outline until the extended outline was completely filled (Fig. 3a, ‘full circle’). For the vision of a simulated predator, we used the spatial acuity from humans (Homo sapiens, 72 cycles per degree, cpd)36,45,46 in the basic scenario. The full details for the parameters are provided in Supplementary Table S1 (a–g).
    Figure 3

    (a) Basic Scenario: Seven stages of the artificial chick setup with varying number of thin, non-transparent appendages having all the same length. (b) Scenario 1: varying appendage thickness applied to the Basic Scenario. (c) Scenario 2: varying appendage transparency applied to the Basic Scenario. (d) Scenario 3: varying appendage length heterogeneity applied to the Basic Scenario. (e) Scenario 4: varying background complexity with chessboard backgrounds. (f) Scenario 5: high, medium and low spatial acuity applied to the Basic Scenario. (a–f) The analysed region of interest (ROI) is highlighted in red for clarification only. The figure was produced in Adobe Photoshop29 and InDesign30.

    Full size image

    To further explore the mechanism, we altered appendage characteristics, background and the spatial acuity of the predator. First, we increased appendage thickness to 2 Pt/8 pixels/ 0.7 mm (Scenario 1a) and 3 Pt/12 pixels/1.1 mm (Scenario 1b) resulting in decreased inter-appendage intervals (Fig. 3b and Supplementary Table S1, h–u). Second, we changed appendage transparency to 25% (Scenario 2a) and 50% transparency (Scenario 2b) (Fig. 3c and Supplementary Table S1, v to ai). Third, we varied the appendage length heterogeneity; half of the appendages having 50% of the length (Scenario 3a), and half of the appendages at 25% and one quarter at 50% of the original appendage length (Scenario 3b) (Fig. 3d and Supplementary Table S1, aj to aw). Fourth, we investigated the effect of background complexity on the detectability of the outline. As background, we used a chessboard pattern with large squares (346 pixels/29.3 mm, Scenario 4a) and with small squares (86 pixels/7.3 mm, Scenario 4b) (Fig. 3e and Supplementary Table S1, ax to bk). Fifth, we altered the spatial acuity to test whether or how the visual systems of different predators would affect detectability. We simulated the spatial acuity of a corvid predator (30 cpd, Scenario 5a) and canid predator (10 cpd, Scenario 5b) (Fig. 3f and Supplementary Table S1, bl to by), the two most common predators of ground-nesting plovers16,47,48. This range also covered other potential predators (Supplementary Table S2).
    We did not account for differences in colour vision between different predators as the setup mostly consists of greyscale images that predominantly differ in luminance. Note that in many animals, visual acuity is greater for achromatic than chromatic stimuli34,49.
    We conducted visual modelling and visual analysis using the Quantitative Colour Pattern Analysis (QCPA) framework27 integrated into the Multispectral Image Analysis and Calibration (MICA) toolbox50 for ImageJ version 1.52a51. We converted the generated images into multispectral images containing the red, green and blue channel in a stack and transformed them further into 32-bits/channel cone-catch images based on the human visual system, which are required by the framework. To create the luminance channel, we averaged the long and medium wave channel, which is thought to be representative of human vision52. We modelled the spatial acuity with Gaussian Acuity Control at a viewing distance of 1300 mm and a minimum resolvable angle (MRA) of 0.01389. To increase biological accuracy, we applied a Receptor Noise Limited (RNL) filter that reduces noise and reconstructs edges in the image. The RNL filter used the Weber fractions “Human 0.05” provided by the framework (longwave 0.05, mediumwave 0.07071, shortwave 0.1657), luminance 0.1, 5 iterations, a radius of 5 pixels and a falloff of 3 pixels as specified in van den Berg et al.27 (Supplementary Fig. S1).
    To test for the detectability of the outline, we used LEIA27, which is conceptually similar to the boundary strength analysis34. Boundary strength analysis requires an image with clearly delineated (clustered) colour and luminance pattern elements. However, a large degree of subthreshold details, which may be still perceived by the viewer gets lost in the clustering process. LEIA has the advantage of not requiring such a clustered input and therefore can be directly applied to RNL filtered images. LEIA measures the edge intensity (i.e. the luminance contrast) locally at each position in the image. The output image displays ΔS values in a 32-bit stack of four slices, where each slice shows the values measured in different angles (horizontal, vertical and the two diagonals, for more details, see van den Berg et al.27).
    We ran LEIA on the chosen region of interest (ROI) with the same Weber fractions used for the RNL filter. The ROI was a 180 pixel-wide band that included the area of the appendages extended by 30 pixels towards the object inside and towards the outside (Fig. 3a). We log-transformed the ΔS values as recommended for natural scenes53 to make the results comparable to the natural background images used in Experiment 2 (see below). To test whether the size of the ROI affected our results, we ran an additional analysis using a 1500 × 1500 pixel-wide rectangle surrounding the object as the ROI, which included a bigger area of the background and the full object inside (Supplementary Fig. S2).
    We extracted the luminance ΔS values from the four slices of the output image stack in ImageJ and stored them in separate matrices for further analysis using R version 3.5.328. ImageJ generally assigned values outside the chosen ROI to zero. Thus, we first discarded all values of zero. We then set all negative values that arose as artefacts in areas without any edges to zero, in order to make them biologically meaningful. We then identified the parallel maximum (R function pmax ()) of the four interrelated direction matrices and transferred this value to a new matrix.
    High luminance and colour contrasts imply high conspicuousness34. Consequently, a lower luminance contrast leads to lower conspicuousness and therefore, better camouflage. As the outline is an important cue for predators locating and identifying a prey item7, we assumed that especially low contrasts in the outline of an object improve camouflage. Thus, a reduction of edge intensity in the object outline by the appendages indicates a camouflage improvement. To test whether the object outline became less detectable we compared the edge intensity of the outline pixels in the basic scenario without appendages (Supplementary Table S1, a) with corresponding pixels from other scenarios. The outline pixels were characterised by high edge intensity and constituted a prominent peak. They comprised 1.59% of all pixels in the analysis focused on the contour region (see “Results”, Fig. 1a). For all scenarios, we calculated the mean edge intensity of the high edge intensity pixels (HEI pixels) and identified the changes with parameter variation. Unless otherwise stated we used R28 to produce graphs and panels.
    As an alternative mechanism, we tested whether appendages create a transition zone with intermediate luminance around the object (Mean Luminance Comparison (MLC), Supplementary material). We calculated the mean luminance of the object inside up to the border (object region), the area covered by appendages (appendage region) and the background (background region). We predicted that the appendage region would be characterised by intermediate luminance between object and background and therefore provide a luminance transition zone to conceal the object outline.
    Experiment 2: chick photographs
    Using pictures of young snowy plover chicks hiding when approached by a simulated predator, we tested if protruding neoptile feathers helped to conceal the chicks’ outline and therefore improve their camouflage.
    We studied snowy plovers in their natural environment at Bahía de Ceuta, Sinaloa, Mexico. Fieldwork permits were granted by the Secretaría de Medio Ambiente y Recursos Naturales (SEMARNAT). All field activities were performed in accordance with the approved ethical guidelines outlined by SEMARNAT. The breeding site consists of salt flats that are sparsely vegetated and surrounded by mangroves54. The predators of chicks are not well described but likely similar to the egg predator community that includes several mammalian predators such as racoon, opossum, coyote, bob cat, avian predators such as crested caracara Caracara cheriway and reptiles17. General field methodology is provided elsewhere55,56. In 2017, we took photographs of young (one to 3 days old) chicks hiding on the ground, that had already left the nest scrape. To photograph the chicks, two observers approached free-roaming families with two mobile hides within the period one hour after sunrise and one hour before sunset. At a distance of 100–200 m, one observer acted as ‘predator’, left the hide and openly approached the brood while the second observer kept watching the chicks. The chicks responded by crouching to the ground and staying motionless while the parents were alarming. The second observer directed the ‘predator’ to the approximate hiding place. When searching for the chicks, we took great care to reduce the number of steps to avoid modification of the ground through our tracks.
    Once the first chick had been found, the second observer joined the ‘predator’ and took the chick photographs. We used a Nikon D7000 camera converted to full spectrum including the UV range (Optic Makario GmbH, Germany) and a Nikkor macro 105 mm lens that allows transmission of light at low wavebands. The equipment was chosen because calibration data were available for this combination50. Each hiding background was photographed with and without the chick using a UV pass filter for the UV spectrum and a UV/IR blocking filter (“IR-Neutralisationsfilter NG”, Optic Makario GmbH, Germany) for the visible spectrum. The camera was set to an aperture of f/8, ISO 400 and the pictures were stored in “RAW” file format. We used exposure bracketing to produce three images to ensure that at least one picture was not over or underexposed. A 25% reflectance standard (Zenith Polymer Diffuse Reflectance Standard provided by SPHEREOPTICS, Germany) placed in the corner of each picture enabled a subsequent standardizing of light conditions.
    In total, we took pictures of 32 chicks from 15 families. For 21 chicks we obtained photographs suitable for further analyses with an unobstructed view to the entire chick and only one chick per photograph. Of these, we randomly selected pictures of 15 chicks. Unfortunately, it was not possible to obtain proper alignment of visual and UV pictures in ImageJ as either chick or camera moved slightly in the break between changing filters for the two settings. Therefore, we restricted our analyses to human colour vision and discarded the UV pictures for further analysis.
    In each picture, we manually selected the chick outline and the feather-boundary as a basis for the ROIs (Fig. 2a–c). The chick outline included bill, legs, rings and all areas densely covered by feathers without background shining through. We then marked the feather-boundary, i.e., the smoothened line created by the protruding neoptile feather tips. In the next step, we transferred images of chicks with or without protruding feathers, i.e. cropped at feather-boundary or chick outline, respectively, and inserted them into a uniform or the natural background. First, we cropped the chick without protruding feathers and transferred it into a uniform black background. Second, we cropped the chick including all feathers and inserted it into exactly the same hiding spot on the picture of the natural background (Fig. 2b). Third, we cropped the chick excluding the protruding feathers and transferred it into the natural background (Fig. 2c).
    We then proceeded with LEIA following the protocol of experiment 1 with the following changes. Again, the selected ROI was the contour region ranging from the chick outline extended by 30 pixels towards the chick inside to the feather-boundary extended by 30 pixels towards the outside. We excluded all areas of the ROI that showed a shadow of the chick as the chicks’ shadow was missing on the empty natural background images to which the cropped chicks were transferred to (Fig. 2a–c). We used the images of the cropped chicks on the black background to determine the threshold of the HEI pixels according to the protocol of experiment 1 for each chick separately. For each cropped chick that was transferred to the picture with the natural background, we compared the mean edge intensity of the HEI pixels provided by LEIA with and without protruding feathers (Fig. 2b,c) using a two-sided paired t-test.
    We also calculated mean luminance differences for chick photographs. Details for this MLC are given in the supplementary material. More

  • in

    Recovery of logged forest fragments in a human-modified tropical landscape during the 2015-16 El Niño

    1.
    Le Quéré, C. et al. Global Carbon Budget 2018. Earth Syst. Sci. Data 10, 2141–2194 (2018).
    ADS  Article  Google Scholar 
    2.
    Cook-Patton, S. C. et al. Mapping carbon accumulation potential from global natural forest regrowth. Nature 585, 545–550 (2020).
    ADS  CAS  PubMed  Article  Google Scholar 

    3.
    Houghton, R. A., Byers, B. & Nassikas, A. A. A role for tropical forests in stabilizing atmospheric CO2. Nat. Clim. Chang. 5, 1022–1023 (2015).
    ADS  Article  Google Scholar 

    4.
    Chazdon, R. L. et al. Carbon sequestration potential of second-growth forest regeneration in the Latin American tropics. Sci. Adv. 2, e1501639 (2016).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    5.
    Philipson, C. D. et al. Active restoration accelerates the carbon recovery of human-modified tropical forests. Science 369, 838–841 (2020).
    ADS  CAS  PubMed  Article  Google Scholar 

    6.
    Taubert, F. et al. Global patterns of tropical forest fragmentation. Nature 554, 519–522 (2018).
    ADS  CAS  PubMed  Article  Google Scholar 

    7.
    Tabarelli, M., Lopes, A. V. & Peres, C. A. Edge-effects drive tropical forest fragments towards an early-successional system. Biotropica 40, 657–661 (2008).
    Article  Google Scholar 

    8.
    Arroyo-Rodríguez, V. et al. Multiple successional pathways in human-modified tropical landscapes: new insights from forest succession, forest fragmentation and landscape ecology research. Biol. Rev. Camb. Philos. Soc. 92, 326–340 (2017).
    PubMed  Article  Google Scholar 

    9.
    Collins, C. D. et al. Fragmentation affects plant community composition over time. Ecography 40, 119–130 (2017).
    Article  Google Scholar 

    10.
    Hansen, M. C. et al. The fate of tropical forest fragments. Sci. Adv. 6 eaax8574 (2020).

    11.
    Qie, L. et al. Long-term carbon sink in Borneo’s forests halted by drought and vulnerable to edge effects. Nat. Commun. 8, 1966 (2017).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    12.
    Thirumalai, K., DiNezio, P. N., Okumura, Y. & Deser, C. Extreme temperatures in Southeast Asia caused by El Niño and worsened by global warming. Nat. Commun. 8, 15531 (2017).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    13.
    Cai, W. et al. Increased variability of eastern Pacific El Niño under greenhouse warming. Nature 564, 201–206 (2018).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    14.
    Vogel, M. M., Hauser, M. & Seneviratne, S. I. Projected changes in hot, dry and wet extreme events’ clusters in CMIP6 multi-model ensemble. Environ. Res. Lett. 15, 094021 (2020).
    ADS  Article  Google Scholar 

    15.
    McDowell, N. et al. Drivers and mechanisms of tree mortality in moist tropical forests. N. Phytol. https://doi.org/10.1111/nph.15027@10.1111/(ISSN)1469-8137.DroughtImpactsonTropicalForests (2018).

    16.
    Walker, A. P. et al. Decadal biomass increment in early secondary succession woody ecosystems is increased by CO2 enrichment. Nat. Commun. 10, 454 (2019).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    17.
    Riutta, T. et al. Logging disturbance shifts net primary productivity and its allocation in Bornean tropical forests. Glob. Chang. Biol. 24, 2913–2928 (2018).
    ADS  PubMed  Article  Google Scholar 

    18.
    Both, S. et al. Logging and soil nutrients independently explain plant trait expression in tropical forests. N. Phytol. 221, 1853–1865 (2019).
    CAS  Article  Google Scholar 

    19.
    Swinfield, T. et al. Imaging spectroscopy reveals the effects of topography and logging on the leaf chemistry of tropical forest canopy trees. Glob. Chang. Biol. 26, 989–1002 (2020).
    ADS  PubMed  Article  Google Scholar 

    20.
    Jotan, P., Maycock, C. R., Burslem, D., Berhaman, A. & Both, S. Comparative vessel traits of macaranga gigantea and vatica dulitensis from Malaysian Borneo. J. Trop. Sci. 32, 25–34 (2020).
    Google Scholar 

    21.
    Uriarte, M. et al. Impacts of climate variability on tree demography in second growth tropical forests: the importance of regional context for predicting successional trajectories. Biotropica 48, 780–797 (2016).
    Article  Google Scholar 

    22.
    Park Williams, A. et al. Temperature as a potent driver of regional forest drought stress and tree mortality. Nat. Clim. Chang. 3, 292–297 (2013).
    ADS  Article  Google Scholar 

    23.
    Wolfe, B. T., Sperry, J. S. & Kursar, T. A. Does leaf shedding protect stems from cavitation during seasonal droughts? A test of the hydraulic fuse hypothesis. N. Phytol. 212, 1007–1018 (2016).
    Article  Google Scholar 

    24.
    Slik, J. W. F. El Niño droughts and their effects on tree species composition and diversity in tropical rain forests. Oecologia 141, 114–120 (2004).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    25.
    Jucker, T. et al. Topography shapes the structure, composition and function of tropical forest landscapes. Ecol. Lett. 21, 989–1000 (2018).
    PubMed  PubMed Central  Article  Google Scholar 

    26.
    Itoh, A. et al. Importance of topography and soil texture in the spatial distribution of two sympatric dipterocarp trees in a Bornean rainforest. Ecol. Res. 18, 307–320 (2003).
    Article  Google Scholar 

    27.
    Stovall, A. E. L., Shugart, H. H. & Yang, X. Reply to ‘Height-related changes in forest composition explain increasing tree mortality with height during an extreme drought’. Nat. Commun. 11, 3401 (2020).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    28.
    Maréchaux, I. et al. Drought tolerance as predicted by leaf water potential at turgor loss point varies strongly across species within an Amazonian forest. Funct. Ecol. 29, 1268–1277 (2015).
    Article  Google Scholar 

    29.
    Cosme, L. H. M., Schietti, J., Costa, F. R. C. & Oliveira, R. S. The importance of hydraulic architecture to the distribution patterns of trees in a central Amazonian forest. N. Phytol. 215, 113–125 (2017).
    Article  Google Scholar 

    30.
    Bittencourt, P. R. L. et al. Amazonia trees have limited capacity to acclimate plant hydraulic properties in response to long-term drought. Glob. Chang. Biol. https://doi.org/10.1111/gcb.15040 (2020).

    31.
    Luke, S. H. et al. Riparian buffers in tropical agriculture: Scientific support, effectiveness and directions for policy. J. Appl. Ecol. 56, 85–92 (2019).
    Article  Google Scholar 

    32.
    Padfield, R. et al. Co-Producing a Research Agenda for Sustainable Palm Oil. https://doi.org/10.3389/ffgc.2019.00013 (2019).

    33.
    Zhao, K. et al. Utility of multitemporal lidar for forest and carbon monitoring: Tree growth, biomass dynamics, and carbon flux. Remote Sens. Environ. 204, 883–897 (2018).
    ADS  Article  Google Scholar 

    34.
    Simonson, W., Ruiz-Benito, P., Valladares, F. & Coomes, D. Modelling above-ground carbon dynamics using multi-temporal airborne lidar: insights from a Mediterranean woodland. Biogeosciences 13, 961–973 (2016).
    ADS  CAS  Article  Google Scholar 

    35.
    Leitold, V. et al. El Niño drought increased canopy turnover in Amazon forests. N. Phytol. 219, 959–971 (2018).
    CAS  Article  Google Scholar 

    36.
    Moura, Y. Mde et al. Carbon dynamics in a human-modified tropical forest: a case study using multi-temporal LiDAR data. Remote Sens. 12, 430 (2020).
    ADS  Article  Google Scholar 

    37.
    Simonson, W., Allen, H. & Coomes, D. Effect of tree phenology on LiDAR measurement of mediterranean forest structure. Remote Sens. 10, 659 (2018).
    ADS  Article  Google Scholar 

    38.
    Sullivan, M. J. P. et al. Long-term thermal sensitivity of Earth’s tropical forests. Science 368, 869–874 (2020).
    ADS  CAS  PubMed  Article  Google Scholar 

    39.
    Coomes, D. A. et al. Area-based vs tree-centric approaches to mapping forest carbon in Southeast Asian forests from airborne laser scanning data. Remote Sens. Environ. 194, 77–88 (2017).
    ADS  Article  Google Scholar 

    40.
    Asner, G. P. et al. Mapped aboveground carbon stocks to advance forest conservation and recovery in Malaysian Borneo. Biol. Conserv. 217, 289–310 (2018).
    Article  Google Scholar 

    41.
    Burton, C., Rifai, S. & Malhi, Y. Inter-comparison and assessment of gridded climate products over tropical forests during the 2015/2016 El Niño. Philosophical Transactions of the Royal Society B: Biological Sciences. 373, 1760 (2018).
    Article  Google Scholar 

    42.
    Ordway, E. M. & Asner, G. P. Carbon declines along tropical forest edges correspond to heterogeneous effects on canopy structure and function. Proc. Natl Acad. Sci. USA 117, 7863–7870 (2020).
    CAS  PubMed  Article  Google Scholar 

    43.
    Phillips, O. L. et al. Drought sensitivity of the Amazon rainforest. Science 323, 1344–1347 (2009).
    ADS  CAS  PubMed  Article  Google Scholar 

    44.
    Sevanto, S. Phloem transport and drought. J. Exp. Bot. 65, 1751–1759 (2014).
    CAS  PubMed  Article  Google Scholar 

    45.
    Aleixo, I. et al. Amazonian rainforest tree mortality driven by climate and functional traits. Nat. Clim. Chang. 9, 384–388 (2019).
    ADS  Article  Google Scholar 

    46.
    Woodgate, W. et al. Quantifying the impact of woody material on leaf area index estimation from hemispherical photography using 3D canopy simulations. Agric. Meteorol. 226-227, 1–12 (2016).
    Article  Google Scholar 

    47.
    Nunes, M. H. et al. Changes in leaf functional traits of rainforest canopy trees associated with an El Niño event in Borneo. Environ. Res. Lett. 14, 085005 (2019).
    ADS  CAS  Article  Google Scholar 

    48.
    Tang, H. & Dubayah, R. Light-driven growth in Amazon evergreen forests explained by seasonal variations of vertical canopy structure. Proc. Natl Acad. Sci. USA 114, 2640–2644 (2017).
    CAS  PubMed  Article  Google Scholar 

    49.
    Doughty, C. E. et al. Drought impact on forest carbon dynamics and fluxes in Amazonia. Nature 519, 78–82 (2015).
    ADS  CAS  PubMed  Article  Google Scholar 

    50.
    Marvin, D. C. & Asner, G. P. Branchfall dominates annual carbon flux across lowland Amazonian forests. Environ. Res. Lett. 11, 094027 (2016).
    ADS  Article  CAS  Google Scholar 

    51.
    Roussel, J.-R., Caspersen, J., Béland, M., Thomas, S. & Achim, A. Removing bias from LiDAR-based estimates of canopy height: accounting for the effects of pulse density and footprint size. Remote Sens. Environ. 198, 1–16 (2017).
    ADS  Article  Google Scholar 

    52.
    Sist, P. & Nguyen-Thé, N. Logging damage and the subsequent dynamics of a dipterocarp forest in East Kalimantan (1990–1996). Ecol. Manag. 165, 85–103 (2002).
    Article  Google Scholar 

    53.
    Rutishauser, E. et al. Rapid tree carbon stock recovery in managed Amazonian forests. Curr. Biol. 25, R787–R788 (2015).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    54.
    Giambelluca, T. W., Ziegler, A. D., Nullet, M. A., Truong, D. M. & Tran, L. T. Transpiration in a small tropical forest patch. Agric. Meteorol. 117, 1–22 (2003).
    Article  Google Scholar 

    55.
    Ewers, R. M. & Banks-Leite, C. Fragmentation impairs the microclimate buffering effect of tropical forests. PLoS ONE 8, e58093 (2013).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    56.
    Laurance, W. F. & Curran, T. J. Impacts of wind disturbance on fragmented tropical forests: a review and synthesis. Austral. Ecol. 33, 399–408 (2008).
    Article  Google Scholar 

    57.
    Jucker, T. et al. Canopy structure and topography jointly constrain the microclimate of human-modified tropical landscapes. Glob. Chang. Biol. 24, 5243–5258 (2018).
    ADS  PubMed  Article  PubMed Central  Google Scholar 

    58.
    Laurance, W. F. Theory meets reality: how habitat fragmentation research has transcended island biogeographic theory. Biol. Conserv. 141, 1731–1744 (2008).
    Article  Google Scholar 

    59.
    Muscarella, R., Kolyaie, S., Morton, D. C., Zimmerman, J. K. & Uriarte, M. Effects of topography on tropical forest structure depend on climate context. J. Ecol. 108, 145–159 (2020).
    Article  Google Scholar 

    60.
    Werner, F. A. & Homeier, J. Is tropical montane forest heterogeneity promoted by a resource‐driven feedback cycle? Evidence from nutrient relations, herbivory and litter decomposition along a topographical gradient. Funct. Ecol. 29, 430–440 (2015).
    Article  Google Scholar 

    61.
    Gonzalez‐Akre, E. et al. Patterns of tree mortality in a temperate deciduous forest derived from a large forest dynamics plot. Ecosphere 7, G04014 (2016).
    Article  Google Scholar 

    62.
    Hansen, M. C. et al. High-resolution global maps of 21st-century forest cover change. Science 342, 850–853 (2013).
    ADS  CAS  Article  Google Scholar 

    63.
    Williamson, J. et al. Riparian Buffers Act as Microclimatic Refugia in Oil Palm Landscapes. https://doi.org/10.17863/CAM.57796 (2020).

    64.
    Hurst, M. D., Mudd, S. M., Walcott, R., Attal, M. & Yoo, K. Using hilltop curvature to derive the spatial distribution of erosion rates: hilltop curvature predicts erosion rates. J. Geophys. Res. 117, F2 (2012).
    Google Scholar 

    65.
    Gaveau, D. L. A. et al. Rapid conversions and avoided deforestation: examining four decades of industrial plantation expansion in Borneo. Sci. Rep. 6, 32017 (2016).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    66.
    Ewers, R. M. et al. A large-scale forest fragmentation experiment: the Stability of Altered Forest Ecosystems Project. Philos. Trans. R. Soc. Lond. B Biol. Sci. 366, 3292–3302 (2011).
    PubMed  PubMed Central  Article  Google Scholar 

    67.
    Pfeifer, M. et al. Mapping the structure of Borneo’s tropical forests across a degradation gradient. Remote Sens. Environ. 176, 84–97 (2016).
    ADS  Article  Google Scholar 

    68.
    Reynolds, G., Payne, J., Sinun, W., Mosigil, G. & Walsh, R. P. D. Changes in forest land use and management in Sabah, Malaysian Borneo, 1990-2010, with a focus on the Danum Valley region. Philos. Trans. R. Soc. Lond. B Biol. Sci. 366, 3168–3176 (2011).
    PubMed  PubMed Central  Article  Google Scholar 

    69.
    Clarke, A. Principles of thermal ecology: temperature, energy, and life. Oxford Scholarship Online https://doi.org/10.1093/oso/9780199551668.001.0001 (2017).

    70.
    Bolton, D. The computation of equivalent potential temperature. Mon. Weather Rev. 108, 1046–1053 (1980).
    ADS  Article  Google Scholar 

    71.
    Asner, G. P. et al. Carnegie airborne observatory-2: increasing science data dimensionality via high-fidelity multi-sensor fusion. Remote Sens. Environ. 124, 454–465 (2012).
    ADS  Article  Google Scholar 

    72.
    Jucker, T. et al. Estimating aboveground carbon density and its uncertainty in Borneo’s structurally complex tropical forests using airborne laser scanning. Biogeosciences, 15, 3811–3830 (2018).
    ADS  Article  Google Scholar 

    73.
    Khosravipour, A., Skidmore, A. K., Isenburg, M., Wang, T. & Hussin, Y. A. Generating pit-free canopy height models from airborne lidar. Photogramm. Eng. Remote Sens. 80, 863–872 (2014).
    Article  Google Scholar 

    74.
    Asner, G. P. & Mascaro, J. Mapping tropical forest carbon: calibrating plot estimates to a simple LiDAR metric. Remote Sens. Environ. 140, 614–624 (2014).
    ADS  Article  Google Scholar 

    75.
    Pearson, T. R. H., Brown, S. & Casarim, F. M. Carbon emissions from tropical forest degradation caused by logging. Environ. Res. Lett. 9, 034017 (2014).
    ADS  Article  CAS  Google Scholar 

    76.
    Gobakken, T. G. & Næsset, E. N. Assessing effects of laser point density, ground sampling intensity, and field sample plot size on biophysical stand properties derived from airborne laser scanner data. Can. J. For. Res. https://doi.org/10.1139/X07-219 (2008).

    77.
    Gray, C. L., Slade, E. M., Mann, D. J. & Lewis, O. T. Do riparian reserves support dung beetle biodiversity and ecosystem services in oil palm-dominated tropical landscapes? Ecol. Evol. 4, 1049–1060 (2014).
    PubMed  PubMed Central  Article  Google Scholar 

    78.
    Metcalfe, P., Beven, K. & Freer, J. Dynamic TOPMODEL: a new implementation in R and its sensitivity to time and space steps. Environ. Model. Softw. 72, 155–172 (2015).
    Article  Google Scholar 

    79.
    Condit, R. et al. Tropical forest dynamics across a rainfall gradient and the impact of an El Niño Dry Season. J. Trop. Ecol. 20, 51–72 (2004).
    Article  Google Scholar 

    80.
    H’edl, R. et al. A new technique for inventory of permanent plots in tropical forests: a case study from lowland dipterocarp forest in Kuala Belalong, Brunei Darussalam. Blumea J. Biodivers. Evolut. Biogeogr. Plants 54, 124–130 (2009).
    Article  Google Scholar 

    81.
    Kent, R., Lindsell, J. A., Laurin, G. V., Valentini, R. & Coomes, D. A. Airborne LiDAR detects selectively logged tropical forest even in an advanced stage of recovery. Remote Sens. 7, 8348–8367 (2015).
    ADS  Article  Google Scholar 

    82.
    Gonsamo, A., Walter, J.-M. N. & Pellikka, P. Sampling gap fraction and size for estimating leaf area and clumping indices from hemispherical photographs. Can. J. Res. 40, 1588–1603 (2010).
    Article  Google Scholar 

    83.
    Kalacska, M. E. R., Sanchez-Azofeifa, G. A., Calvo-Alvarado, J. C., Rivard, B. & Quesada, M. Effects of season and successional stage on leaf area index and spectral vegetation indices in three mesoamerican tropical dry forests1. Biotropica 37, 486–496 (2005).
    Article  Google Scholar 

    84.
    Thimonier, A., Sedivy, I. & Schleppi, P. Estimating leaf area index in different types of mature forest stands in Switzerland: a comparison of methods. Eur. J. Res. 129, 543–562 (2010).
    Article  Google Scholar 

    85.
    Chen, J. M. & Cihlar, J. Quantifying the effect of canopy architecture on optical measurements of leaf area index using two gap size analysis methods. IEEE Trans. Geosci. Remote Sens. 33, 777–787 (1995).
    ADS  Article  Google Scholar 

    86.
    Schleppi, P., Conedera, M., Sedivy, I. & Thimonier, A. Correcting non-linearity and slope effects in the estimation of the leaf area index of forests from hemispherical photographs. Agric. Meteorol. 144, 236–242 (2007).
    Article  Google Scholar 

    87.
    Harmon, M. E., Whigham, D. F., Sexton, J. & Olmsted, I. Decomposition and mass of woody detritus in the dry tropical forests of the Northeastern Yucatan Peninsula, Mexico. Biotropica 27, 305–316 (1995).
    Article  Google Scholar 

    88.
    Team, R. C. R: A Language and Environment for Statistical Computing. (Foundation for Statistical Computing, Vienna, Austria. 2017). Available online: www.r-project.org (accessed 14 Febuary 2019) (2018).

    89.
    Ploton, P. et al. Spatial validation reveals poor predictive performance of large-scale ecological mapping models. Nat. Commun. 11, 4540 (2020).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    90.
    Wedeux, B. et al. Dynamics of a human-modified tropical peat swamp forest revealed by repeat lidar surveys. Glob. Chang. Biol. https://doi.org/10.1111/gcb.15108 (2020). More