More stories

  • in

    Landscapes of pesticide risk

    A large-scale field study finds that different bee species experience different levels of risk from pesticides, depending on how much land is farmed within their foraging range. For bumblebees and solitary bees, more seminatural habitat means less risk from pesticides, but this is not true for honeybees.In the discussion of how to protect bees from pesticides, bees are often treated as a monolith. It is assumed that what is good for one species is good for all, and that pesticides or changes to agricultural landscapes would affect all bee species equally. This is often taken one step further, with the western honeybee (Apis mellifera) being used as a surrogate species for all bees. Yet despite this simplification there are around 2,000 species of bee in Europe1 and 20,000 worldwide2 with a dazzling diversity of niches and life histories. With this in mind, the question arises of how valid the assumption is that honeybees represent a good surrogate species. In this issue of Nature Ecology & Evolution, Knapp et al.3 investigate this question by measuring how three species of bee with differing life histories respond to different agricultural land-use intensities, and find that a species’ foraging range plays a big part in pesticide exposure risk. More

  • in

    Better incentives are needed to reward academic software development

    Department of Ecology and Evolutionary Biology and Eversource Energy Center, University of Connecticut, Storrs, CT, USACory MerowDepartment of Ecology and Evolutionary Biology, University of Arizona, Tucson, AZ, USABrad Boyle & Brian J. EnquistDepartment of Geography, Florida State University, Tallahassee, FL, USAXiao FengBiodiversity and Biocomplexity Unit, Okinawa Institute of Science and Technology Graduate University, Onna, JapanJamie M. KassDepartment of Geography, University at Buffalo, Buffalo, NY, USABrian S. Maitner & Adam M. WilsonSchool of Biology and Ecology, University of Maine, Orono, ME, USABrian McGillMitchell Center for Sustainability Solutions, University of Maine, Orono, ME, USABrian McGillCenter for Macroecology, Evolution and Climate, Globe Institute, University of Copenhagen, Copenhagen, DenmarkHannah OwensFlorida Museum of Natural History, University of Florida, Gainesville, FL, USAHannah OwensDepartment of Biological Sciences, Purdue University, West Lafayette, IN, USADaniel S. ParkPurdue Center for Plant Biology, Purdue University, West Lafayette, IN, USADaniel S. ParkDepartment of Environmental Systems Science, Institute of Integrative Biology, ETH Zürich, Zurich, SwitzerlandAndrea PazDepartment of Biology, City College of the City University of New York, New York, NY, USAGonzalo E. Pinilla-BuitragoPhD Program in Biology, Graduate Center of the City University of New York, New York, NY, USAGonzalo E. Pinilla-BuitragoDepartment of Ecology and Evolutionary Biology, University of Connecticut, Storrs, CT, USAMark C. UrbanCenter of Biological Risk, University of Connecticut, Storrs, CT, USAMark C. UrbanDepartamento de Ecoloxía e Bioloxía Animal, Universidade de Vigo, Vigo, SpainSara Varela More

  • in

    Climate change as a global amplifier of human–wildlife conflict

    Abrahms, B. Human–wildlife conflict under climate change. Science 373, 484–485 (2021).Article 
    CAS 

    Google Scholar 
    Nyhus, P. J. Human–wildlife conflict and coexistence. Annu. Rev. Environ. Resour. 41, 143–171 (2016).Article 

    Google Scholar 
    Ripple, W. J. et al. Extinction risk is most acute for the world’s largest and smallest vertebrates. Proc. Natl Acad. Sci. USA 114, 10678–10683 (2017).Article 
    CAS 

    Google Scholar 
    Estes, J. A. et al. Trophic downgrading of planet Earth. Science 333, 301–306 (2011).Article 
    CAS 

    Google Scholar 
    Abrahms, B. et al. Data from: Climate change as an amplifier of human–wildlife conflict. Github https://github.com/Abrahms-Lab/Climate-Conflict-Review (2022).IPCC Climate Change 2021: The Physical Science Basis (eds Masson-Delmotte, V. et al.) (Cambridge Univ. Press, 2021).Sydeman, W. J., Santora, J. A., Thompson, S. A., Marinovic, B. & Lorenzo, E. D. Increasing variance in North Pacific climate relates to unprecedented ecosystem variability off California. Glob. Change Biol. 19, 1662–1675 (2013).Article 

    Google Scholar 
    Wang, G. et al. Continued increase of extreme El Niño frequency long after 1.5 °C warming stabilization. Nat. Clim. Change 7, 568–572 (2017).Article 

    Google Scholar 
    Filazzola, A., Blagrave, K., Imrit, M. A. & Sharma, S. Climate change drives increases in extreme events for lake ice in the Northern Hemisphere. Geophys. Res. Lett. 47, e2020GL089608 (2020).Marzeion, B., Cogley, J. G., Richter, K. & Parkes, D. Attribution of global glacier mass loss to anthropogenic and natural causes. Science 345, 919–921 (2014).Article 
    CAS 

    Google Scholar 
    Martin, J. T. et al. Increased drought severity tracks warming in the United States’ largest river basin. Proc. Natl Acad. Sci. USA 117, 11328–11336 (2020).Article 
    CAS 

    Google Scholar 
    Laufkötter, C., Zscheischler, J. & Frölicher, T. L. High-impact marine heatwaves attributable to human-induced global warming. Science 369, 1621–1625 (2020).Article 

    Google Scholar 
    Donat, M. G., Lowry, A. L., Alexander, L. V., O’Gorman, P. A. & Maher, N. More extreme precipitation in the world’s dry and wet regions. Nat. Clim. Change 6, 508–513 (2016).Article 

    Google Scholar 
    Walther, G.-R. et al. Ecological responses to recent climate change. Nature 416, 389–395 (2002).Article 
    CAS 

    Google Scholar 
    Pecl, G. T. et al. Biodiversity redistribution under climate change: impacts on ecosystems and human well-being. Science 355, eaai9214 (2017).Article 

    Google Scholar 
    Lin, D., Xia, J. & Wan, S. Climate warming and biomass accumulation of terrestrial plants: a meta‐analysis. New Phytol. 188, 187–198 (2010).Article 

    Google Scholar 
    Kharouba, H. M. & Wolkovich, E. M. Disconnects between ecological theory and data in phenological mismatch research. Nat. Clim. Change 10, 406–415 (2020).Article 

    Google Scholar 
    Marinovic, B. B., Croll, D. A., Gong, N., Benson, S. R. & Chavez, F. P. Effects of the 1997–1999 El Niño and La Niña events on zooplankton abundance and euphausiid community composition within the Monterey Bay coastal upwelling system. Prog. Oceanogr. 54, 265–277 (2002).Article 

    Google Scholar 
    Kardol, P. et al. Climate change effects on plant biomass alter dominance patterns and community evenness in an experimental old‐field ecosystem. Glob. Change Biol. 16, 2676–2687 (2010).Article 

    Google Scholar 
    Prugh, L. R. et al. Ecological winners and losers of extreme drought in California. Nat. Clim. Change 8, 819–824 (2018).Article 

    Google Scholar 
    Sorte, C. J. B., Williams, S. L. & Zerebecki, R. A. Ocean warming increases threat of invasive species in a marine fouling community. Ecology 91, 2198–2204 (2010).Article 

    Google Scholar 
    Muehlenbein, M. P. Human–environment interactions, current and future directions. Hum. Environ. Interact. 1, 79–94 (2012).
    Google Scholar 
    Sinervo, B. et al. Erosion of lizard diversity by climate change and altered thermal niches. Science 328, 894–899 (2010).Article 
    CAS 

    Google Scholar 
    Mason, T. H. E., Keane, A., Redpath, S. M. & Bunnefeld, N. The changing environment of conservation conflict: geese and farming in Scotland. J. Appl. Ecol. 55, 651–662 (2018).Article 

    Google Scholar 
    Pérez-Flores, J., Mardero, S., López-Cen, A., Contreras-Moreno, F. M. & Pérez-Flores, J. Human–wildlife conflicts and drought in the greater Calakmul Region, Mexico: implications for tapir conservation. Neotrop. Biol. Conserv. 16, 539–563 (2021).Article 

    Google Scholar 
    Mariki, S. B., Svarstad, H. & Benjaminsen, T. A. Elephants over the cliff: explaining wildlife killings in Tanzania. Land Use Policy 44, 19–30 (2015).Article 

    Google Scholar 
    Mukeka, J. M., Ogutu, J. O., Kanga, E. & Roskaft, E. Spatial and temporal dynamics of human–wildlife conflicts in the Kenya Greater Tsavo Ecosystem. Hum. Wildl. Interact. 14, 255–272 (2020).
    Google Scholar 
    Popp, J. N., Hamr, J., Chan, C. & Mallory, F. F. Elk (Cervus elaphus) railway mortality in Ontario. Can. J. Zool. 96, 1066–1070 (2018).Article 

    Google Scholar 
    Olson, D. D. et al. How does variation in winter weather affect deer–vehicle collision rates? Wildl. Biol. 21, 80–87 (2015).Article 

    Google Scholar 
    Nyhus, P. & Tilson, R. Agroforestry, elephants, and tigers: balancing conservation theory and practice in human-dominated landscapes of Southeast Asia. Agric. Ecosyst. Environ. 104, 87–97 (2004).Article 

    Google Scholar 
    Laufenberg, J. S., Johnson, H. E., Doherty, P. F. & Breck, S. W. Compounding effects of human development and a natural food shortage on a black bear population along a human development–wildland interface. Biol. Conserv 224, 188–198 (2018).Article 

    Google Scholar 
    Blondin, H., Abrahms, B., Crowder, L. B. & Hazen, E. L. Combining high temporal resolution whale distribution and vessel tracking data improves estimates of ship strike risk. Biol. Conserv. 250, 108757 (2020).Article 

    Google Scholar 
    Abrahms, B. et al. Dynamic ensemble models to predict distributions and anthropogenic risk exposure for highly mobile species. Divers. Distrib. 25, 1182–1193 (2019).Article 

    Google Scholar 
    Gaynor, K. M., Hojnowski, C. E., Carter, N. H. & Brashares, J. S. The influence of human disturbance on wildlife nocturnality. Science 360, 1232–1235 (2018).Article 
    CAS 

    Google Scholar 
    Kabir, M., Ghoddousi, A., Awan, M. S. & Awan, M. N. Assessment of human–leopard conflict in Machiara National Park, Azad Jammu and Kashmir, Pakistan. Eur. J. Wildl. Res. 60, 291–296 (2014).Article 

    Google Scholar 
    Soto, J. R. Patterns and Determinants of Human–Carnivore Conflicts in the Tropical Lowlands of Guatemala (Univ. of Florida, 2008).Honda, T. & Kozakai, C. Mechanisms of human–black bear conflicts in Japan: in preparation for climate change. Sci. Total Environ. 739, 140028 (2020).Article 
    CAS 

    Google Scholar 
    Mukeka, J. M., Ogutu, J. O., Kanga, E. & Røskaft, E. Human–wildlife conflicts and their correlates in Narok County, Kenya. Glob. Ecol. Conserv. 18, e00620 (2019).Article 

    Google Scholar 
    Kuiper, T. R. et al. Seasonal herding practices influence predation on domestic stock by African lions along a protected area boundary. Biol. Conserv. 191, 546–554 (2015).Article 

    Google Scholar 
    Funston, P. J., Mills, M. G. L. & Biggs, H. C. Factors affecting the hunting success of male and female lions in the Kruger National Park. J. Zool. 253, 419–431 (2001).Article 

    Google Scholar 
    Schiess-Meier, M., Ramsauer, S., Gabanapelo, T. & Konig, B. Livestock predation—insights from problem animal control registers in Botswana. J. Wildl. Manag. 71, 1267–1274 (2007).Article 

    Google Scholar 
    Wilder, J. M. et al. Polar bear attacks on humans: implications of a changing climate. Wildl. Soc. B 41, 537–547 (2017).Article 

    Google Scholar 
    Towns, L., Derocher, A. E., Stirling, I., Lunn, N. J. & Hedman, D. Spatial and temporal patterns of problem polar bears in Churchill, Manitoba. Polar Biol. 32, 1529–1537 (2009).Article 

    Google Scholar 
    Schmidt, A. & Clark, D. ‘It’s just a matter of time:’ lessons from agency and community responses to polar bear-inflicted human injury. Conserv. Soc. 16, 64 (2018).Article 

    Google Scholar 
    Koenig, J., Shine, R. & Shea, G. The dangers of life in the city: patterns of activity, injury and mortality in suburban lizards (Tiliqua scincoides). J. Herpetol. 36, 62–68 (2002).Article 

    Google Scholar 
    Whitaker, P. B. & Shine, R. Responses of free-ranging brownsnakes (Pseudonaja textilis: Elapidae) to encounters with humans. Wildl. Res. 26, 689–704 (1999).Article 

    Google Scholar 
    Saberwal, V., Gibbs, J., Chellam, R. & Johnsingh, A. Lion–human conflict in the Gir Forest, India. Conserv. Biol. 8, 501–507 (1994).Article 

    Google Scholar 
    Ferreira, S. M. & Viljoen, P. African large carnivore population changes in response to a drought. Afr. J. Wildl. Res. https://hdl.handle.net/10520/ejc-wild2-v52-n1-a1 (2022).Masiaine, S. et al. Landscape-level changes to large mammal space use in response to a pastoralist incursion. Ecol. Indic. 121, 107091 (2021).Article 

    Google Scholar 
    Kiria, E. A Spatial Multi-criteria Analysis of Land Use, Land Cover and Climate Changes on Wildlife Ecosystems Planning and Management in Meru Conservation Area (Chuka Univ., 2018).Benansio, J., Demaya, G., Dendi, D. & Luiselli, L. Attacks by Nile crocodiles (Crocodylus nilotticus) on humans and livestock in the Sudd wetlands, South Sudan. Russ. J. Herpetol. https://doi.org/10.30906/1026-2296-2022-29-4-199-205 (2022).Melia, N., Haines, K. & Hawkins, E. Sea ice decline and 21st century trans‐Arctic shipping routes. Geophys. Res. Lett. 43, 9720–9728 (2016).Article 

    Google Scholar 
    Ivanova, S. V. et al. Shipping alters the movement and behavior of Arctic cod (Boreogadus saida), a keystone fish in Arctic marine ecosystems. Ecol. Appl. 30, e02050 (2020).Article 

    Google Scholar 
    Hauser, D. D. W., Laidre, K. L. & Stern, H. L. Vulnerability of Arctic marine mammals to vessel traffic in the increasingly ice-free Northwest Passage and Northern Sea Route. Proc. Natl Acad. Sci. USA 5, 201803543–201803546 (2018).
    Google Scholar 
    Hovelsrud, G. K., McKenna, M. & Huntington, H. P. Marine mammal harvests and other interactions with humans. Ecol. Appl. 18, S135–S147 (2008).Article 

    Google Scholar 
    Santora, J. A. et al. Habitat compression and ecosystem shifts as potential links between marine heatwave and record whale entanglements. Nat. Commun. 11, 536 (2020).Samhouri, J. F. et al. Marine heatwave challenges solutions to human–wildlife conflict. Proc. R. Soc. B 288, 20211607 (2021).Article 

    Google Scholar 
    Chapman, B. K. & McPhee, D. Global shark attack hotspots: identifying underlying factors behind increased unprovoked shark bite incidence. Ocean Coast. Manag. 133, 72–84 (2016).Article 

    Google Scholar 
    Burgess, G., Buch, R., Carvalho, F., Garner, B. & Walker, C. in Sharks and Their Relatives II: Biodiversity, Adaptive Physiology, and Conservation (eds Carrier, J. C. et al.) 541–565 (CRC Press, 2010).Woodward, A. R., Leone, E. H., Dutton, H. J., Waller, J. E. & Hord, L. Characteristics of American alligator bites on people in Florida. J. Wildl. Manag. 83, 1437–1453 (2019).Article 

    Google Scholar 
    Khorozyan, I., Soofi, M., Ghoddousi, A., Hamidi, A. K. & Waltert, M. The relationship between climate, diseases of domestic animals and human–carnivore conflicts. Basic Appl. Ecol. 16, 703–713 (2015).Article 

    Google Scholar 
    Treves, A. & Bruskotter, J. Tolerance for predatory wildlife. Science 344, 476–477 (2014).Article 
    CAS 

    Google Scholar 
    Carpenter, S. Exploring the impact of climate change on the future of community‐based wildlife conservation. Conserv. Sci. Pract. 4, e585 (2022).Nisi, A. Cryptic Neighbors: Connecting Movement Ecology and Population Dynamics for a Large Carnivore in a Human-dominated Landscape (Univ. California, 2021). .Asiyanbi, A. P. A political ecology of REDD+: property rights, militarised protectionism, and carbonised exclusion in Cross River. Geoforum 77, 146–156 (2016).Article 

    Google Scholar 
    Dawson, N. M. et al. Barriers to equity in REDD+: deficiencies in national interpretation processes constrain adaptation to context. Environ. Sci. Policy 88, 1–9 (2018).Article 

    Google Scholar 
    Rabaiotti, D. et al. High temperatures and human pressures interact to influence mortality in an African carnivore. Ecol. Evol. 11, 8495–8506 (2021).Article 

    Google Scholar 
    Vargas, S. P., Castro-Carrasco, P. J., Rust, N. A. & F, J. L. R. Climate change contributing to conflicts between livestock farming and guanaco conservation in central Chile: a subjective theories approach. Oryx 55, 275–283 (2021).Article 

    Google Scholar 
    Heemskerk, S. et al. Temporal dynamics of human–polar bear conflicts in Churchill, Manitoba. Glob. Ecol. Conserv. 24, e01320 (2020).Article 

    Google Scholar 
    Schell, C. J. et al. The evolutionary consequences of human–wildlife conflict in cities. Evol. Appl. 14, 178–197 (2021).Article 

    Google Scholar 
    Clark, J. A. & May, R. M. Taxonomic bias in conservation research. Science 297, 191–192 (2002).Article 
    CAS 

    Google Scholar 
    Ravenelle, J. & Nyhus, P. J. Global patterns and trends in human–wildlife conflict compensation. Conserv. Biol. 31, 1247–1256 (2017).Article 

    Google Scholar 
    Zack, C. S., Milne, B. T. & Dunn, W. Southern oscillation index as an indicator of encounters between humans and black bears in New Mexico. Wildl. Soc. Bull. 31, 517–520 (2003).
    Google Scholar 
    Acosta-Jamett, G., Gutiérrez, J. R., Kelt, D. A., Meserve, P. L. & Previtali, M. A. El Niño Southern Oscillation drives conflict between wild carnivores and livestock farmers in a semiarid area in Chile. J. Arid. Environ. 126, 76–80 (2016).Article 

    Google Scholar 
    Timmermann, A. et al. El Niño–Southern Oscillation complexity. Nature 559, 535–545 (2018).Article 
    CAS 

    Google Scholar 
    Wittemyer, G., Elsen, P., Bean, W. T., Burton, A. C. O. & Brashares, J. S. Accelerated human population growth at protected area edges. Science 321, 123–126 (2008).Article 
    CAS 

    Google Scholar 
    Powell, G., Versluys, T. M. M., Williams, J. J., Tiedt, S. & Pooley, S. Using environmental niche modelling to investigate abiotic predictors of crocodilian attacks on people. Oryx 54, 639–647 (2020).Article 

    Google Scholar 
    Maxwell, S. M. et al. Dynamic ocean management: defining and conceptualizing real-time management of the ocean. Mar. Policy 58, 42–50 (2015).Article 

    Google Scholar 
    Maxwell, S. M., Gjerde, K. M., Conners, M. G. & Crowder, L. B. Mobile protected areas for biodiversity on the high seas. Science 367, 252–254 (2020).Article 
    CAS 

    Google Scholar 
    Pons, M. et al. Trade-offs between bycatch and target catches in static versus dynamic fishery closures. Proc. Natl Acad. Sci. USA 119, e2114508119 (2022).Article 

    Google Scholar 
    Oestreich, W. K., Chapman, M. S. & Crowder, L. B. A comparative analysis of dynamic management in marine and terrestrial systems. Front. Ecol. Environ. 18, 496–504 (2020).Article 

    Google Scholar 
    Mason, N., Ward, M., Watson, J. E. M., Venter, O. & Runting, R. K. Global opportunities and challenges for transboundary conservation. Nat. Ecol. Evol. 4, 694–701 (2020).Article 

    Google Scholar 
    Dickman, A. J., Macdonald, E. A. & Macdonald, D. W. A review of financial instruments to pay for predator conservation and encourage human–carnivore coexistence. Proc. Natl Acad. Sci. USA 108, 13937–13944 (2011).Article 
    CAS 

    Google Scholar 
    Ej, N. G. et al. A Future for All: The Need for Human–Wildlife Coexistence (UNEP, 2021).Lankford, A. J., Svancara, L. K., Lawler, J. J. & Vierling, K. Comparison of climate change vulnerability assessments for wildlife. Wildl. Soc. Bull. 38, 386–394 (2014).Article 

    Google Scholar 
    Syombua, M. An Analysis of Human–Wildlife Conflicts in Tsavo West-Amboseli Agro-Ecosystem Using an Integrated Geospatial Approach: A Case Study of Taveta District (Univ. of Nairobi, 2013).Malhi, Y. et al. The role of large wild animals in climate change mitigation and adaptation. Curr. Biol. 32, R181–R196 (2022).Article 
    CAS 

    Google Scholar 
    Aryal, A., Brunton, D. & Raubenheimer, D. Impact of climate change on human–wildlife–ecosystem interactions in the Trans-Himalaya region of Nepal. Theor. Appl. Climatol. 115, 517–529 (2013).Article 

    Google Scholar 
    Aryal, A., Brunton, D., Ji, W., Barraclough, R. K. & Raubenheimer, D. Human–carnivore conflict: ecological and economical sustainability of predation on livestock by snow leopard and other carnivores in the Himalaya. Sustain. Sci. 9, 321–329 (2014).Article 

    Google Scholar 
    Aryal, A. et al. Predicting the distributions of predator (snow leopard) and prey (blue sheep) under climate change in the Himalaya. Ecol. Evol. 6, 4065–4075 (2016).Article 

    Google Scholar 
    Nowell, K., Li, J., Paltsyn, M. & Sharma, R. An Ounce of Prevention: Snow Leopard Crime Revisited (Traffic Report, 2016). More

  • in

    Ecological traits interact with landscape context to determine bees’ pesticide risk

    Tilman, D., Cassman, K. G., Matson, P. A., Naylor, R. & Polasky, S. Agricultural sustainability and intensive production practices. Nature 418, 671–677 (2002).Article 
    CAS 
    PubMed 

    Google Scholar 
    Tilman, D. et al. Forecasting agriculturally driven global environmental change. Science 292, 281–284 (2001).Article 
    CAS 
    PubMed 

    Google Scholar 
    IPBES: Summary for Policymakers. In The Assessment Report on Pollinators, Pollination and Food Production (eds Potts, S. G. et al.) (IPBES, 2016).Potts, S. G. et al. Safeguarding pollinators and their values to human well-being. Nature 540, 220–229 (2016).Article 
    CAS 
    PubMed 

    Google Scholar 
    Sgolastra, F. et al. Synergistic mortality between a neonicotinoid insecticide and an ergosterol-biosynthesis-inhibiting fungicide in three bee species. Pest Manag Sci. 73, 1236–1243 (2016).Article 
    PubMed 

    Google Scholar 
    Whitehorn, P. R., O’Connor, S., Wackers, F. L. & Goulson, D. Neonicotinoid pesticide reduces bumble bee colony growth and queen production. Science 336, 351–352 (2012).Article 
    CAS 
    PubMed 

    Google Scholar 
    Rundlöf, M. et al. Seed coating with a neonicotinoid insecticide negatively affects wild bees. Nature 521, 77–80 (2015).Article 
    PubMed 

    Google Scholar 
    Woodcock, B. et al. Impacts of neonicotinoid use on long-term population changes in wild bees in England. Nat. Commun. 7, 12459 (2016).Stuligross, C. & Williams, N. Past insecticide exposure reduces bee reproduction and population growth rate. Proc. Natl Acad. Sci. USA 118, e2109909118 (2021).Stanley, D. A. et al. Neonicotinoid pesticide exposure impairs crop pollination services provided by bumblebees. Nature 528, 548–550 (2015).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Tamburini, G. et al. Fungicide and insecticide exposure adversely impacts bumblebees and pollination services under semi-field conditions. Environ. Int. 157, 106813 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Sponsler, D. B. et al. Pesticides and pollinators: a socioecological synthesis. Sci. Total Environ. 662, 1012–1027 (2019).Article 
    CAS 
    PubMed 

    Google Scholar 
    Meehan, T. D., Werling, B. P., Landis, D. A. & Gratton, C. Agricultural landscape simplification and insecticide use in the Midwestern United States. Proc. Natl Acad. Sci. USA 108, 11500–11505 (2011).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Nicholson, C. C. & Williams, N. M. Cropland heterogeneity drives frequency and intensity of pesticide use. Environ. Res. 16, 074008 (2021).CAS 

    Google Scholar 
    Böhme, F., Bischoff, G., Zebitz, C. P. W., Rosenkranz, P. & Wallner, K. Pesticide residue survey of pollen loads collected by honeybees (Apis mellifera) in daily intervals at three agricultural sites in South Germany. PLoS ONE 13, e0199995 (2018).Larsen, A. E. & Noack, F. Impact of local and landscape complexity on the stability of field-level pest control. Nat. Sustain. 4, 120–128 (2021).Article 

    Google Scholar 
    Botías, C. et al. Neonicotinoid residues in wildflowers, a potential route of chronic exposure for bees. Environ. Sci. Technol. 49, 12731–12740 (2015).Article 
    PubMed 

    Google Scholar 
    Krupke, C. H., Holland, J. D., Long, E. Y. & Eitzer, B. D. Planting of neonicotinoid-treated maize poses risks for honey bees and other non-target organisms over a wide area without consistent crop yield benefit. J. Appl. Ecol. 54, 1449–1458 (2017).Article 
    CAS 

    Google Scholar 
    Wintermantel, D. et al. Neonicotinoid-induced mortality risk for bees foraging on oilseed rape nectar persists despite EU moratorium. Sci. Total Environ. 704, 135400 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Krupke, C. H., Hunt, G. J., Eitzer, B. D., Andino, G. & Given, K. Multiple routes of pesticide exposure for honey bees living near agricultural fields. PLoS ONE 7, e29268 (2012).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Long, E. Y. & Krupke, C. H. Non-cultivated plants present a season-long route of pesticide exposure for honey bees. Nat. Commun. 7, 11629 (2016).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    David, A. et al. Widespread contamination of wildflower and bee-collected pollen with complex mixtures of neonicotinoids and fungicides commonly applied to crops. Environ. Int. 88, 169–178 (2016).Article 
    CAS 
    PubMed 

    Google Scholar 
    Heinrich, B. The foraging specializations of individual bumblebees. Ecol. Monogr. 46, 105–128 (1976).Article 

    Google Scholar 
    Bolin, A., Smith, H. G., Lonsdorf, E. V. & Olsson, O. Scale-dependent foraging tradeoff allows competitive coexistence. Oikos 127, 1575–1585 (2018).Article 

    Google Scholar 
    Cresswell, J. E., Osborne, J. L. & Goulson, D. An economic model of the limits to foraging range in central place foragers with numerical solutions for bumblebees. Ecol. Entomol. 25, 249–255 (2000).Article 

    Google Scholar 
    Rundlöf, M. et al. Flower plantings support wild bee reproduction and may also mitigate pesticide exposure effects. J. Appl. Ecol. 59, 2117–2127 (2022).Article 

    Google Scholar 
    Graham, K. K. et al. Identities, concentrations, and sources of pesticide exposure in pollen collected by managed bees during blueberry pollination. Sci. Rep. 11, 16857 (2021).Centrella, M. et al. Diet diversity and pesticide risk mediate the negative effects of land use change on solitary bee offspring production. J. Appl. Ecol. 57, 1031–1042 (2020).Article 
    CAS 

    Google Scholar 
    De Palma, A. et al. Ecological traits affect the sensitivity of bees to land-use pressures in European agricultural landscapes. J. Appl. Ecol. 52, 1567–1577 (2015).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Sponsler, D. B. & Johnson, R. M. Mechanistic modeling of pesticide exposure: the missing keystone of honey bee toxicology. Environ. Toxicol. Chem. 36, 871–881 (2017).Article 
    CAS 
    PubMed 

    Google Scholar 
    Holzschuh, A., Dormann, C. F., Tscharntke, T. & Steffan-Dewenter, I. Mass-flowering crops enhance wild bee abundance. Oecologia 172, 477–484 (2013).Article 
    PubMed 

    Google Scholar 
    McArt, S. H., Fersch, A. A., Milano, N. J., Truitt, L. L. & Böröczky, K. High pesticide risk to honey bees despite low focal crop pollen collection during pollination of a mass blooming crop. Sci. Rep. 7, 46554 (2017).Sanchez-Bayo, F. & Goka, K. Pesticide residues and bees—a risk assessment. PLoS ONE 9, e94482 (2014).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Zioga, E., Kelly, R., White, B. & Stout, J. C. Plant protection product residues in plant pollen and nectar: a review of current knowledge. Environ. Res. 189, 109873 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    The European Green Deal (European Commission, 2019).More, S. J., Auteri, D., Rortais, A. & Pagani, S. EFSA is working to protect bees and shape the future of environmental risk assessment. EFSA J. 19, e190101 (2021).Schmolke, A. et al. Assessment of the vulnerability to pesticide exposures across bee species. Environ. Toxicol. Chem. 40, 2640–2651 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Rollin, O. et al. Differences of floral resource use between honey bees and wild bees in an intensive farming system. Agric. Ecosyst. Environ. 179, 78–86 (2013).Article 

    Google Scholar 
    Persson, A. S. & Smith, H. G. Seasonal persistence of bumblebee populations is affected by landscape context. Agric. Ecosyst. Environ. 165, 201–209 (2013).Article 

    Google Scholar 
    Samuelson, A. E., Schürch, R. & Leadbeater, E. Dancing bees evaluate central urban forage resources as superior to agricultural land. J. Appl. Ecol. 59, 79–88 (2022).Article 

    Google Scholar 
    Milner, A. M. & Boyd, I. L. Toward pesticidovigilance. Science 357, 1232–1234 https://doi.org/10.1126/science.aan2683 (2017).Nowell, L. H., Norman, J. E., Moran, P. W., Martin, J. D. & Stone, W. W. Pesticide toxicity index—a tool for assessing potential toxicity of pesticide mixtures to freshwater aquatic organisms. Sci. Total Environ. 476–477, 144–157 (2014).Article 
    PubMed 

    Google Scholar 
    Mullin, C. A., Frazier, M., Frazier, J. L., Ashcraft, S. & Simonds, R. High levels of miticides and agrochemicals in North American apiaries: implications for honey bee health. PLoS ONE 5, 9754 (2010).Article 

    Google Scholar 
    Pettis, J. S. et al. Crop pollination exposes honey bees to pesticides which alters their susceptibility to the gut pathogen Nosema ceranae. PLoS ONE 8, e70182 (2013).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Végh, R., Sörös, C., Majercsik, N. & Sipos, L. Determination of pesticides in bee pollen: validation of a multiresidue high-performance liquid chromatography-mass spectrometry/mass spectrometry method and testing pollen samples of selected botanical origin. J. Agric. Food Chem. 70, 1507–1515 (2022).Article 
    PubMed 

    Google Scholar 
    Park, M. G., Blitzer, E. J., Gibbs, J., Losey, J. E. & Danforth, B. N. Negative effects of pesticides on wild bee communities can be buffered by landscape context. Proc. R. Soc. B 282, 20150299 (2015).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Graham, K. K. et al. Pesticide risk to managed bees during blueberry pollination is primarily driven by off-farm exposures. Sci. Rep. 12, 7189 (2022).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Yourstone, J., Karlsson, M., Klatt, B. K., Olsson, O. & Smith, H. G. Effects of crop and non-crop resources and competition: high importance of trees and oilseed rape for solitary bee reproduction. Biol. Conserv. 261, 109249 (2021).Persson, A. S., Mazier, F. & Smith, H. G. When beggars are choosers—how nesting of a solitary bee is affected by temporal dynamics of pollen plants in the landscape. Ecol. Evol. 8, 5777–5791 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wood, T. J., Holland, J. M. & Goulson, D. Providing foraging resources for solitary bees on farmland: current schemes for pollinators benefit a limited suite of species. J. Appl. Ecol. 54, 323–333 (2016).Garthwaite, D. et al. Collection of Pesticide Application Data in View of Performing Environmental Risk Assessments for Pesticides (EFSA, 2017).de Oliveira, R. C., Nascimento Queiroz, S. C., Pinto da Luz, C. F., Silveira Porto, R. & Rath, S. Bee pollen as a bioindicator of environmental pesticide contamination. Chemosphere 163, 525–534 (2016).Article 
    PubMed 

    Google Scholar 
    Arena, M. & Sgolastra, F. A meta-analysis comparing the sensitivity of bees to pesticides. Ecotoxicology 23, 324–334 (2014).Article 
    CAS 
    PubMed 

    Google Scholar 
    Douglas, M. R., Sponsler, D. B., Lonsdorf, E. V. & Grozinger, C. M. County-level analysis reveals a rapidly shifting landscape of insecticide hazard to honey bees (Apis mellifera) on US farmland. Sci. Rep. 10, 797 (2020).Commission Implementing Regulation (EU) 2021/2081 of 26 November 2021 concerning the non-renewal of approval of the active substance indoxacarb, in accordance with Regulation (EC) No 1107/2009 of the European Parliament and of the Council concerning the placing of plant protection products on the market, and amending Commission Implementing Regulation (EU) No 540/2011 (EUR-Lex, 2021); http://data.europa.eu/eli/reg_impl/2021/2081/ojCommission Implementing Regulation (EU) 2020/23 of 13 January 2020 concerning the non-renewal of the approval of the active substance thiacloprid, in accordance with Regulation (EC) No. 1107/2009 of the European Parliament and of the Council concerning the placing of plant protection products on the market, and amending the Annex to Commission Implementing Regulation (EU) No 540/2011 (EUR-Lex, 2020); http://data.europa.eu/eli/reg_impl/2020/23/ojCommission Implementing Regulation (EU) 2018/783 of 29 May 2018 amending Implementing Regulation (EU) No 540/2011 as regards the conditions of approval of the active substance imidacloprid (EUR-Lex, 2018); http://data.europa.eu/eli/reg_impl/2018/783/ojHerbertsson, L., Jonsson, O., Kreuger, J., Smith, H. G. & Rundlöf, M. Scientific note: imidacloprid found in wild plants downstream permanent greenhouses in Sweden. Apidologie 52, 946–949 (2021).Article 

    Google Scholar 
    Tosi, S. et al. Long-term field-realistic exposure to a next-generation pesticide, flupyradifurone, impairs honey bee behaviour and survival. Commun. Biol. 4, 805 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Siviter, H. & Muth, F. Do novel insecticides pose a threat to beneficial insects?: novel insecticides harm insects. Proc. R. Soc. B 287, 20201265 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    EFSA. Guidance on the risk assessment of plant protection products on bees (Apis mellifera, Bombus spp. and solitary bees). EFSA J. 11, 3295 (2013).Guidance for Assessing Pesticide Risks to Bees (US EPA, 2014).Boyle, N. K. et al. Workshop on pesticide exposure assessment paradigm for non-apis bees: foundation and summaries. Environ. Entomol. 48, 4–11 (2019).Article 
    PubMed 

    Google Scholar 
    EFSA. Analysis of the evidence to support the definition of specific protection goals for bumble bees and solitary bees. EFSA J. 19, EN-7125 (2022).Garibaldi, L. A. et al. Wild pollinators enhance fruit set of crops regardless of honey bee abundance. Science 339, 1608–1611 (2013).Article 
    CAS 
    PubMed 

    Google Scholar 
    Tscharntke, T., Grass, I., Wanger, T. C. & Westphal, C. Restoring biodiversity needs more than reducing pesticides. Trends Ecol. Evol. 37, 115–116 (2022).Article 
    PubMed 

    Google Scholar 
    Topping, C. J. et al. Holistic environmental risk assessment for bees. Science 37, 897 (2021).Article 

    Google Scholar 
    Tsvetkov, N. et al. Chronic exposure to neonicotinoids reduces honey bee health near corn crops. Science 356, 1395–1397 (2017).Article 
    CAS 
    PubMed 

    Google Scholar 
    Jonsson, O., Fries, I. & Kreuger, J. Utveckling av Analysmetoder och Screening av Växtskyddsmedel i bin och Pollen (CKB, 2013).Sawyer, R. Pollen Identification for Beekeepers (Univ. Cardiff Press, 1981).IUPAC Pesticide Properties Data Base (Univ. of Hertfordshire, 2022).EFSA Scientific Committee & More, S.J. et al. Guidance on harmonised methodologies for human health, animal health and ecological risk assessment of combined exposure to multiple chemicals. EFSA J. 17, e05634 (2019).Martin, O. et al. Ten years of research on synergisms and antagonisms in chemical mixtures: a systematic review and quantitative reappraisal of mixture studies. Environ. Int. 146, 106206 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    DiBartolomeis, M., Kegley, S., Mineau, P., Radford, R. & Klein, K. An assessment of acute insecticide toxicity loading (AITL) of chemical pesticides used on agricultural land in the United States. PLoS ONE 14, e0220029 (2019).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Test No. 213: Honeybees, Acute Oral Toxicity Test (OECD, 1998); https://doi.org/10.1787/9789264070165-enPrice, P. S. & Han, X. Maximum cumulative ratio (MCR) as a tool for assessing the value of performing a cumulative risk assessment. Int. J. Environ. Res. Public Health 8, 2212–2225 (2011).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bates, D., Mächler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1–48 (2015).Oksanen, J. et al. vegan community ecology package version 2.6-2 (2022).Lenth, R. emmeans: Estimated marginal means, aka least-squares means (2022).Lüdecke, D., Ben-shachar, M. S., Patil, I. & Makowski, D. performance: an R package for assessment, comparison and testing of statistical models statement of need. J. Open Source Softw. 6, 3139 (2021).Nakagawa, S. & Schielzeth, H. A general and simple method for obtaining R2 from generalized linear mixed-effects models. Methods Ecol. Evol. 4, 133–142 (2013).Article 

    Google Scholar 
    Kendall, L. K. et al. The potential and realized foraging movements of bees are differentially determined by body size and sociality. Ecology 103, e3809 (2022).Parreño, M. A. et al. Critical links between biodiversity and health in wild bee conservation. Trends Ecol. Evol. 37, 309–321 (2022).Article 
    PubMed 

    Google Scholar  More

  • in

    Microbial keystone taxa drive succession of plant residue chemistry

    Lal R, Bruce JP. The potential of world cropland soils to sequester C and mitigate the greenhouse effect. Environ Sci Policy. 1999;2:177–85.Article 
    CAS 

    Google Scholar 
    Wang J, Feng L, Palmer PI, Liu Y, Fang S, Bosch H, et al. Large Chinese land carbon sink estimated from atmospheric carbon dioxide data. Nature. 2020;586:720–3.Article 
    CAS 
    PubMed 

    Google Scholar 
    Lal R. Managing soils and ecosystems for mitigating anthropogenic carbon emissions and advancing global food security. Bioscience. 2010;60:708–21.Article 

    Google Scholar 
    Rumpel C, Lehmann J, Chabbi A. ‘4 per 1,000’ initiative will boost soil carbon for climate and food security. Nature. 2018;553:27–27.Article 
    CAS 
    PubMed 

    Google Scholar 
    Zhao Y, Wang M, Hu S, Zhang X, Ouyang Z, Zhang G, et al. Economics- and policy-driven organic carbon input enhancement dominates soil organic carbon accumulation in Chinese croplands. Proc Natl Acad Sci USA. 2018;115:4045–50.Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Yang F, Xu Y, Cui Y, Meng Y, Dong Y, Li R, et al. Variation of soil organic matter content in croplands of china over the last three decades (in Chinese). Acta Pedol Sin. 2017;5:1047–56.
    Google Scholar 
    Lehmann J, Hansel CM, Kaiser C, Kleber M, Maher K, Manzoni S, et al. Persistence of soil organic carbon caused by functional complexity. Nat Geosci. 2020;13:529–34.Article 
    CAS 

    Google Scholar 
    Schmidt M, Torn MS, Abiven S, Dittmar T, Guggenberger G, Janssens IA. et al. Persistence of soil organic matter as an ecosystem property. Nature. 2011;478:49–56.Article 
    CAS 
    PubMed 

    Google Scholar 
    Lehmann J, Kleber M. The contentious nature of soil organic matter. Nature. 2015;528:60–68.Article 
    CAS 
    PubMed 

    Google Scholar 
    Cotrufo MF, Soong JL, Horton AJ, Campbell EE, Haddix ML, Wall DH, et al. Formation of soil organic matter via biochemical and physical pathways of litter mass loss. Nat Geosci. 2015;8:776–9.Article 
    CAS 

    Google Scholar 
    Schnitzer M, Monreal CM. Quo vadis soil organic matter research? A biological link to the chemistry of humification. Adv Agron. 2011;113:139–213.
    Google Scholar 
    Wang X, Sun B, Mao J, Sui Y, Cao X. Structural convergence of maize and wheat straw during two-year decomposition under different climate conditions. Environ Sci Technol. 2012;46:7159–65.Article 
    CAS 
    PubMed 

    Google Scholar 
    Liang C, Schimel JP, Jastrow JD. The importance of anabolism in microbial control over soil carbon storage. Nat Microbiol. 2017;2:17105.Article 
    CAS 
    PubMed 

    Google Scholar 
    Wickings K, Grandy AS, Reed SC, Cleveland CC. The origin of litter chemical complexity during decomposition. Ecol Lett. 2012;15:1180–8.Article 
    PubMed 

    Google Scholar 
    Grandy AS, Neff JC. Molecular C dynamics downstream: the biochemical decomposition sequence and its impact on soil organic matter structure and function. Sci Total Environ. 2008;404:297–307.Article 
    CAS 
    PubMed 

    Google Scholar 
    Jenkinson DS, Ayanaba A. Decomposition of 14C labeled plant material under tropical conditions. Soil Sci Soc Am J. 1977;41:912–5.Article 
    CAS 

    Google Scholar 
    Li Y, Chen N, Harmon ME, Li Y, Cao X, Chappell MA, et al. Plant species rather than climate greatly alters the temporal pattern of litter chemical composition during long-term decomposition. Sci Rep. 2015;5:15783.Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Preston CM, Nault JR, Trofymow JA, Smyth C, Grp CW. Chemical changes during 6 years of decomposition of 11 litters in some Canadian forest sites. Part 1. Elemental composition, tannins, phenolics, and proximate fractions. Ecosystems. 2009;12:1053–77.Article 
    CAS 

    Google Scholar 
    Kallenbach CM, Frey SD, Grandy AS. Direct evidence for microbial-derived soil organic matter formation and its ecophysiological controls. Nat Commun. 2016;7:13630.Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wickings K, Stuart Grandy A, Reed S, Cleveland C. Management intensity alters decomposition via biological pathways. Biogeochemistry. 2011;104:365–79.Article 

    Google Scholar 
    Schimel JP, Schaeffer SM. Microbial control over carbon cycling in soil. Front Microbiol. 2012;3:348.Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Sun B, Wang X, Wang F, Jiang Y, Zhang X-X. Assessing the relative effects of geographic location and soil type on microbial communities associated with straw decomposition. Appl Environ Microbiol. 2013;79:3327.Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Balser TC, Firestone MK. Linking microbial community composition and soil processes in a California annual grassland and mixed-conifer forest. Biogeochemistry. 2005;73:395–415.Article 
    CAS 

    Google Scholar 
    Grandy AS, Neff JC, Weintrau MN. Carbon structure and enzyme activities in alpine and forest ecosystems. Soil Biol Biochem. 2007;39:2701–11.Article 
    CAS 

    Google Scholar 
    Maynard DS, Crowther TW, Bradford MA. Competitive network determines the direction of the diversity-function relationship. Proc Natl Acad Sci USA. 2017;114:11464–9.Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wagg C, Bender SF, Widmer F, van der Heijden MGA. Soil biodiversity and soil community composition determine ecosystem multifunctionality. Proc Natl Acad Sci USA. 2014;111:5266–70.Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Snajdr J, Cajthaml T, Valaskova V, Merhautova V, Petrankova M, Spetz P, et al. Transformation of Quercus petraea litter: successive changes in litter chemistry are reflected in differential enzyme activity and changes in the microbial community composition. FEMS Microbiol Ecol. 2011;75:291–303.Article 
    CAS 
    PubMed 

    Google Scholar 
    Banerjee S, Kirkby CA, Schmutter D, Bissett A, Kirkegaard JA, Richardson AE. Network analysis reveals functional redundancy and keystone taxa amongst bacterial and fungal communities during organic matter decomposition in an arable soil. Soil Biol Biochem. 2016;97:188–98.Article 
    CAS 

    Google Scholar 
    Banerjee S, Schlaeppi K, van der Heijden MGA. Keystone taxa as drivers of microbiome structure and functioning. Nat Rev Microbiol. 2018;16:567–76.Article 
    CAS 
    PubMed 

    Google Scholar 
    Carrias J-F, Gerphagnon M, Rodríguez-Pérez H, Borrel G, Loiseau C, Corbara B, et al. Resource availability drives bacterial succession during leaf-litter decomposition in a bromeliad ecosystem. FEMS Microbiol Ecol. 2020;96:fiaa045.Article 
    CAS 
    PubMed 

    Google Scholar 
    Zhan P, Liu Y, Wang H, Wang C, Xia M, Wang N, et al. Plant litter decomposition in wetlands is closely associated with phyllospheric fungi as revealed by microbial community dynamics and co-occurrence network. Sci Total Environ. 2021;753:142194.Article 
    CAS 
    PubMed 

    Google Scholar 
    Panettieri M, Knicker H, Murillo JM, Madejon E, Hatcher PG. Soil organic matter degradation in an agricultural chronosequence under different tillage regimes evaluated by organic matter pools, enzymatic activities and CPMAS C-13 NMR. Soil Biol Biochem. 2014;78:170–81.Article 
    CAS 

    Google Scholar 
    Skjemstad JO, Clarke P, Taylor JA, Oades JM, Newman RH. The removal of magnetic-materials from surface soils – a solid state 13C CP/MAS NMR study. Aust J Soil Res. 1994;32:1215–29.Article 
    CAS 

    Google Scholar 
    Sokolenko S, Jézéquel T, Hajjar G, Farjon J, Akoka S, Giraudeau P. Robust 1D NMR lineshape fitting using real and imaginary data in the frequency domain. J Magn Reson. 2019;298:91–100.Article 
    CAS 
    PubMed 

    Google Scholar 
    Grandy AS, Strickland MS, Lauber CL, Bradford MA, Fierer N. The influence of microbial communities, management, and soil texture on soil organic matter chemistry. Geoderma.2009;150:278–86.Article 
    CAS 

    Google Scholar 
    Saiya-Cork KR, Sinsabaugh RL, Zak DR. The effects of long term nitrogen deposition on extracellular enzyme activity in an Acer saccharum forest soil. Soil Biol Biochem. 2002;34:1309–15.Article 
    CAS 

    Google Scholar 
    Allison SD, Jastrow JD. Activities of extracellular enzymes in physically isolated fractions of restored grassland soils. Soil Biol Biochem. 2006;38:3245–56.Article 
    CAS 

    Google Scholar 
    Zhang XD, Amelung W. Gas chromatographic determination of muramic acid, glucosamine, mannosamine, and galactosamine in soils. Soil Biol Biochem. 1996;28:1201–6.Article 
    CAS 

    Google Scholar 
    Lee CK, Barbier BA, Bottos EM, McDonald IR, Cary SC. The inter-valley soil comparative survey: the ecology of dry valley edaphic microbial communities. ISME J. 2012;6:1046–57.Article 
    CAS 
    PubMed 

    Google Scholar 
    Degnan PH, Ochman H. Illumina-based analysis of microbial community diversity. ISME J. 2012;6:183–94.Article 
    CAS 
    PubMed 

    Google Scholar 
    Pruesse E, Quast C, Knittel K, Fuchs BM, Ludwig W, Peplies J, et al. SILVA: a comprehensive online resource for quality checked and aligned ribosomal RNA sequence data compatible with ARB. Nucleic Acids Res. 2007;35:7188–96.Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kõljalg U, Nilsson RH, Abarenkov K, Tedersoo L, Taylor AFS, Bahram M, et al. Towards a unified paradigm for sequence-based identification of fungi. Mol Ecol. 2013;22:5271–7.Article 
    PubMed 

    Google Scholar 
    Faust K, Sathirapongsasuti JF, Izard J, Segata N, Gevers D, Raes J, et al. Microbial co-occurrence relationships in the human microbiome. PLoS Comp Biol. 2012;8:e1002606.Article 
    CAS 

    Google Scholar 
    Chong IG, Jun CH. Performance of some variable selection methods when multicollinearity is present. Chemometrics Intell Lab Syst. 2005;78:103–12.Article 
    CAS 

    Google Scholar 
    Strukelj M, Brais S, Mazerolle MJ, Pare D, Drapeau P. Decomposition patterns of foliar litter and deadwood in managed and unmanaged stands: A 13-year experiment in boreal mixedwoods. Ecosystems. 2018;21:68–84.Article 
    CAS 

    Google Scholar 
    Manzoni S, Piñeiro G, Jackson RB, Jobbágy EG, Kim JH, Porporato A. Analytical models of soil and litter decomposition: Solutions for mass loss and time-dependent decay rates. Soil Biol Biochem. 2012;50:66–76.Article 
    CAS 

    Google Scholar 
    Dixon P. VEGAN, a package of R functions for community ecology. J Veg Sci. 2003;14:927–30.Article 

    Google Scholar 
    Grace JB (ed). Structural Equation Modeling and Natural Systems. Cambridge University Press, Cambridge, 2006.Shen Y, Cheng R, Xiao W, Yang S, Guo Y, Wang N, et al. Labile organic carbon pools and enzyme activities of Pinus massoniana plantation soil as affected by understory vegetation removal and thinning. Sci Rep. 2018;8:573.Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Gallo ME, Lauber CL, Cabaniss SE, Waldrop MP, Sinsabaugh RL, Zak DR. Soil organic matter and litter chemistry response to experimental N deposition in northern temperate deciduous forest ecosystems. Glob Change Biol. 2005;11:1514–21.Article 

    Google Scholar 
    Wilhelm RC, Singh R, Eltis LD, Mohn WW. Bacterial contributions to delignification and lignocellulose degradation in forest soils with metagenomic and quantitative stable isotope probing. ISME J. 2019;13:413–29.Article 
    CAS 
    PubMed 

    Google Scholar 
    Sahay H, Yadav AN, Singh AK, Singh S, Kaushik R, Saxena AK. Hot springs of Indian Himalayas: potential sources of microbial diversity and thermostable hydrolytic enzymes. 3 Biotech. 2017;7:118.Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Robledo M, Rivera L, Jimenez-Zurdo JI, Rivas R, Dazzo F, Velazquez E, et al. Role of Rhizobium endoglucanase CelC2 in cellulose biosynthesis and biofilm formation on plant roots and abiotic surfaces. Micro Cell Factories. 2012;11:125.Article 
    CAS 

    Google Scholar 
    Wang X, Bian Q, Jiang Y, Zhu L, Chen Y, Liang Y, et al. Organic amendments drive shifts in microbial community structure and keystone taxa which increase C mineralization across aggregate size classes. Soil Biol Biochem. 2021;153:108062.Article 
    CAS 

    Google Scholar 
    Joergensen RG. Amino sugars as specific indices for fungal and bacterial residues in soil. Biol Fert Soils. 2018;54:559–68.Article 
    CAS 

    Google Scholar 
    Chen Y, Sun R, Sun T, Chen P, Yu Z, Ding L, et al. Evidence for involvement of keystone fungal taxa in organic phosphorus mineralization in subtropical soil and the impact of labile carbon. Soil Biol Biochem. 2020;148:107900.Article 
    CAS 

    Google Scholar 
    Puentes-Tellez PE, Salles JF. Construction of effective minimal active microbial consortia for lignocellulose degradation. Micro Ecol. 2018;76:419–29.Article 
    CAS 

    Google Scholar 
    Zark M, Dittmar T. Universal molecular structures in natural dissolved organic matter. Nat Commun. 2018;9:3178.Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Lynch LM, Sutfin NA, Fegel TS, Boot CM, Covino TP, Wallenstein MD. River channel connectivity shifts metabolite composition and dissolved organic matter chemistry. Nat Commun. 2019;10:459.Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Filley TR, Boutton TW, Liao JD, Jastrow JD, Gamblin DE. Chemical changes to nonaggregated particulate soil organic matter following grassland-to-woodland transition in a subtropical savanna. J Geophys Res Biogeosci. 2008;113:G03009.Article 

    Google Scholar 
    Stewart CE, Neff JC, Amatangelo KL, Vitousek PM. Vegetation effects on soil organic matter chemistry of aggregate fractions in a Hawaiian forest. Ecosystems. 2011;14:382–97.Article 
    CAS 

    Google Scholar  More

  • in

    Composition, structure and robustness of Lichen guilds

    Spribille, T. et al. Basidiomycete yeasts in the cortex of ascomycete macrolichens. Science 353, 488–492 (2016).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hawksworth, D. L. & Grube, M. Lichens redefined as complex ecosystems. New Phytol. 227, 1281 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Jung, P. et al. Lichens bite the dust-a bioweathering scenario in the atacama desert. iScience 23, 101647 (2020).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Seneviratne, G. & Indrasena, I. Nitrogen fixation in lichens is important for improved rock weathering. J. Biosci. 31, 639–643 (2006).Article 
    PubMed 

    Google Scholar 
    Nybakken, L., Solhaug, K. A., Bilger, W. & Gauslaa, Y. The lichens Xanthoria elegans and Cetraria islandica maintain a high protection against uv-b radiation in arctic habitats. Oecologia 140, 211–216 (2004).Article 
    ADS 
    PubMed 

    Google Scholar 
    Peksa, O. & Škaloud, P. Do photobionts influence the ecology of lichens? A case study of environmental preferences in symbiotic green alga asterochloris (trebouxiophyceae). Mol. Ecol. 20, 3936–3948 (2011).Article 
    PubMed 

    Google Scholar 
    Friedmann, E. I. & Galun, M. Desert algae, lichens and fungi. Desert Biol. 2, 165–212 (1974).Article 

    Google Scholar 
    Conti, M. E. & Cecchetti, G. Biological monitoring: Lichens as bioindicators of air pollution assessment—a review. Environ. Pollut. 114, 471–492 (2001).Article 
    CAS 
    PubMed 

    Google Scholar 
    Van Herk, C., Mathijssen-Spiekman, E. & De Zwart, D. Long distance nitrogen air pollution effects on lichens in Europe. Lichenologist 35, 347–359 (2003).Article 

    Google Scholar 
    Osyczka, P., Lenart-Boroń, A., Boroń, P. & Rola, K. Lichen-forming fungi in postindustrial habitats involve alternative photobionts. Mycologia 113, 43–55 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Margulis, L. & Fester, R. Symbiosis as a Source of Evolutionary Innovation: Speciation and Morphogenesis (MIT press, 1991).
    Google Scholar 
    Solé, R. et al. Synthetic collective intelligence. Biosystems 148, 47–61 (2016).Article 
    PubMed 

    Google Scholar 
    Dal Forno, M. et al. Extensive photobiont sharing in a rapidly radiating cyanolichen clade. Mol. Ecol. 30, 1755–1776 (2021).Article 

    Google Scholar 
    Nishiguchi, M. K. Cospeciation between hosts and symbionts. In Symbiosis 757–774 (Springer, 2001).Hill, D. J. Asymmetric co-evolution in the lichen symbiosis caused by a limited capacity for adaptation in the photobiont. Bot. Rev. 75, 326–338 (2009).Article 

    Google Scholar 
    Muggia, L., Pérez-Ortega, S., Fryday, A., Spribille, T. & Grube, M. Global assessment of genetic variation and phenotypic plasticity in the lichen-forming species Tephromela atra. Fungal Divers. 64, 233–251 (2014).Article 

    Google Scholar 
    Vančurová, L., Muggia, L., Peksa, O., Řídká, T. & Škaloud, P. The complexity of symbiotic interactions influences the ecological amplitude of the host: A case study in stereocaulon (lichenized ascomycota). Mol. Ecol. 27, 3016–3033 (2018).Article 
    PubMed 

    Google Scholar 
    Peksa, O., Gebouská, T., Škvorová, Z., Vančurová, L. & Škaloud, P. The guilds in green algal lichens-an insight into the life of terrestrial symbiotic communities. FEMS Microbiol. Ecol. 98, fiac008 (2022).Article 
    PubMed 

    Google Scholar 
    Wagner, M. et al. Macroclimatic conditions as main drivers for symbiotic association patterns in lecideoid lichens along the transantarctic mountains, ross sea region, antarctica. Sci. Rep. 11, 1–15 (2021).Article 
    MathSciNet 

    Google Scholar 
    Nascimbene, J. & Marini, L. Epiphytic lichen diversity along elevational gradients: Biological traits reveal a complex response to water and energy. J. Biogeogr. 42, 1222–1232 (2015).Article 

    Google Scholar 
    Galloway, D. Lichen biogeography. Lichen Biol. 2, 315–35 (1996).
    Google Scholar 
    Vančurová, L., Malíček, J., Steinová, J. & Škaloud, P. Choosing the right life partner: Ecological drivers of lichen symbiosis. Front. Microbiol. 12, 769304 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Škvorová, Z. et al. Promiscuity in lichens follows clear rules: Partner switching in cladonia is regulated by climatic factors and soil chemistry. Front. Microbiol. 12, 56 (2021).
    Google Scholar 
    Medeiros, I. D. et al. Turnover of lecanoroid mycobionts and their trebouxia photobionts along an elevation gradient in bolivia highlights the role of environment in structuring the lichen symbiosis. Front. Microbiol. 2021, 3859 (2021).
    Google Scholar 
    Marini, L., Nascimbene, J. & Nimis, P. L. Large-scale patterns of epiphytic lichen species richness: Photobiont-dependent response to climate and forest structure. Sci. Total Environ. 409, 4381–4386 (2011).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Saini, K. C., Nayaka, S. & Bast, F. Diversity of lichen photobionts: Their coevolution and bioprospecting potential. Microb. Divers. Ecosyst. Sustain. Biotechnol. Appl. 2019, 307–323 (2019).
    Google Scholar 
    Ivens, A. B., von Beeren, C., Blüthgen, N. & Kronauer, D. J. Studying the complex communities of ants and their symbionts using ecological network analysis. Annu. Rev. Entomol. 61, 353–371 (2016).Article 
    CAS 
    PubMed 

    Google Scholar 
    Ziegler, M., Eguíluz, V. M., Duarte, C. M. & Voolstra, C. R. Rare symbionts may contribute to the resilience of coral-algal assemblages. ISME J. 12, 161–172 (2018).Article 
    PubMed 

    Google Scholar 
    Rikkinen, J. et al. Ecological and evolutionary role of photobiont-mediated guilds in lichens. Symbiosis 2003, 256 (2003).
    Google Scholar 
    Belinchón, R., Yahr, R. & Ellis, C. J. Interactions among species with contrasting dispersal modes explain distributions for epiphytic lichens. Ecography 38, 762–768 (2015).Article 

    Google Scholar 
    Muggia, L. et al. The symbiotic playground of lichen thalli-a highly flexible photobiont association in rock-inhabiting lichens. FEMS Microbiol. Ecol. 85, 313–323 (2013).Article 
    CAS 
    PubMed 

    Google Scholar 
    Rikkinen, J., Oksanen, I. & Lohtander, K. Lichen guilds share related cyanobacterial symbionts. Science 297, 357 (2002).Article 
    CAS 
    PubMed 

    Google Scholar 
    Kaasalainen, U., Tuovinen, V., Mwachala, G., Pellikka, P. & Rikkinen, J. Complex interaction networks among cyanolichens of a tropical biodiversity hotspot. Front. Microbiol. 12, 1246 (2021).Article 

    Google Scholar 
    Werth, S. Fungal-algal interactions in Ramalina menziesii and its associated epiphytic lichen community. Lichenologist 44, 543–560 (2012).Article 

    Google Scholar 
    O’Brien, H. E., Miadlikowska, J. & Lutzoni, F. Assessing population structure and host specialization in lichenized cyanobacteria. New Phytol. 198, 557–566 (2013).Article 
    PubMed 

    Google Scholar 
    Pino-Bodas, R. & Stenroos, S. Global biodiversity patterns of the photobionts associated with the genus cladonia (lecanorales, ascomycota). Microb. Ecol. 82, 173–187 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Miadlikowska, J. et al. New insights into classification and evolution of the lecanoromycetes (pezizomycotina, ascomycota) from phylogenetic analyses of three ribosomal rna-and two protein-coding genes. Mycologia 98, 1088–1103 (2006).Article 
    CAS 
    PubMed 

    Google Scholar 
    Chagnon, P.-L., Magain, N., Miadlikowska, J. & Lutzoni, F. Species diversification and phylogenetically constrained symbiont switching generated high modularity in the lichen genus peltigera. J. Ecol. 107, 1645–1661 (2019).Article 

    Google Scholar 
    Bascompte, J. & Jordano, P. Plant-animal mutualistic networks: The architecture of biodiversity. Annu. Rev. Ecol. Evol. Syst. 38, 567–593 (2007).Article 
    MATH 

    Google Scholar 
    Olesen, J. M., Bascompte, J., Dupont, Y. L. & Jordano, P. The modularity of pollination networks. Proc. Natl. Acad. Sci. 104, 19891–19896 (2007).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 
    MATH 

    Google Scholar 
    Weitz, J. S. et al. Phage-bacteria infection networks. Trends Microbiol. 21, 82–91 (2013).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Maliet, O., Loeuille, N. & Morlon, H. An individual-based model for the eco-evolutionary emergence of bipartite interaction networks. Ecol. Lett. 23, 1623–1634 (2020).Article 
    PubMed 

    Google Scholar 
    Fortuna, M. A. et al. Nestedness versus modularity in ecological networks: Two sides of the same coin?. J. Anim. Ecol. 2010, 811–817 (2010).
    Google Scholar 
    Mariani, M. S., Ren, Z.-M., Bascompte, J. & Tessone, C. J. Nestedness in complex networks: Observation, emergence, and implications. Phys. Rep. 813, 1–90 (2019).Article 
    ADS 
    MathSciNet 

    Google Scholar 
    Almeida-Neto, M., Guimaraes, P., Guimaraes, P. R. Jr., Loyola, R. D. & Ulrich, W. A consistent metric for nestedness analysis in ecological systems: Reconciling concept and measurement. Oikos 117, 1227–1239 (2008).Article 

    Google Scholar 
    Flores, C. O., Valverde, S. & Weitz, J. S. Multi-scale structure and geographic drivers of cross-infection within marine bacteria and phages. ISME J. 7, 520–532 (2013).Article 
    PubMed 

    Google Scholar 
    Sanders, W. B. & Masumoto, H. Lichen algae: The photosynthetic partners in lichen symbioses. Lichenologist 53, 347–393 (2021).Article 

    Google Scholar 
    Duran-Nebreda, S. & Bassel, G. W. Bridging scales in plant biology using network science. Trends Plant Sci. 22, 1001–1003 (2017).Article 
    CAS 
    PubMed 

    Google Scholar 
    Galiana, N. et al. Ecological network complexity scales with area. Nature Ecol. Evol. 2022, 1–8 (2022).
    Google Scholar 
    Solé, R. V. & Valverde, S. Spontaneous emergence of modularity in cellular networks. J. R. Soc. Interface 5, 129–133 (2008).Article 
    PubMed 

    Google Scholar 
    Jackson, M. D., Duran-Nebreda, S. & Bassel, G. W. Network-based approaches to quantify multicellular development. J. R. Soc. Interface 14, 20170484 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Jackson, M. D., Xu, H., Duran-Nebreda, S., Stamm, P. & Bassel, G. W. Topological analysis of multicellular complexity in the plant hypocotyl. Elife 6, e26023 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Jackson, M. D. et al. Global topological order emerges through local mechanical control of cell divisions in the arabidopsis shoot apical meristem. Cell Syst. 8, 53–65 (2019).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Miadlikowska, J. et al. A multigene phylogenetic synthesis for the class lecanoromycetes (ascomycota): 1307 fungi representing 1139 infrageneric taxa, 317 genera and 66 families. Mol. Phylogenet. Evol. 79, 132–168 (2014).Article 
    PubMed 

    Google Scholar 
    Perez-Lamarque, B., Selosse, M.-A., Öpik, M., Morlon, H. & Martos, F. Cheating in arbuscular mycorrhizal mutualism: A network and phylogenetic analysis of mycoheterotrophy. New Phytol. 226, 1822–1835 (2020).Article 
    PubMed 

    Google Scholar 
    Jaccard, P. Étude comparative de la distribution florale dans une portion des alpes et des jura. Bull. Soc. Vaudoise Sci. Naturelles 37, 547–579 (1901).
    Google Scholar 
    Müllner, D. fastcluster: Fast hierarchical, agglomerative clustering routines for r and python. J. Stat. Softw. 53, 1–18 (2013).Article 

    Google Scholar 
    Sole, R. V. & Montoya, M. Complexity and fragility in ecological networks. Proc. R. Soc. Lond. Ser. B Biol. Sci. 268, 2039–2045 (2001).Article 
    CAS 

    Google Scholar 
    Guimaraes, P. R. Jr. The structure of ecological networks across levels of organization. Annu. Rev. Ecol. Evol. Syst. 51, 433–460 (2020).Article 

    Google Scholar 
    Nash, T. H. Lichen Biology (Cambridge University Press, 1996).
    Google Scholar 
    Hawksworth, D. The variety of fungal-algal symbioses, their evolutionary significance, and the nature of lichens. Bot. J. Linn. Soc. 96, 3–20 (1988).Article 

    Google Scholar 
    Richardson, D. H. War in the world of lichens: Parasitism and symbiosis as exemplified by lichens and lichenicolous fungi. Mycol. Res. 103, 641–650 (1999).Article 
    ADS 

    Google Scholar 
    Lücking, R. et al. Do lichens domesticate photobionts like farmers domesticate crops? Evidence from a previously unrecognized lineage of filamentous cyanobacteria. Am. J. Bot. 96, 1409–1418 (2009).Article 
    PubMed 

    Google Scholar 
    Kaasalainen, U., Schmidt, A. R. & Rikkinen, J. Diversity and ecological adaptations in palaeogene lichens. Nature Plants 3, 1–8 (2017).Article 

    Google Scholar 
    Piercey-Normore, M. D. The lichen-forming ascomycete evernia mesomorpha associates with multiple genotypes of Trebouxia jamesii. New Phytol. 169, 331–344 (2006).Article 
    CAS 
    PubMed 

    Google Scholar 
    Rudgers, J. A. & Strauss, S. Y. A selection mosaic in the facultative mutualism between ants and wild cotton. Proc. R. Soc. Lond. Ser. B Biol. Sci. 271, 2481–2488 (2004).Article 

    Google Scholar 
    Spribille, T., Resl, P., Stanton, D. E. & Tagirdzhanova, G. Evolutionary biology of lichen symbioses. New Phytol. 2022, 25 (2022).
    Google Scholar 
    Thébault, E. & Fontaine, C. Stability of ecological communities and the architecture of mutualistic and trophic networks. Science 329, 853–856 (2010).Article 
    ADS 
    PubMed 

    Google Scholar 
    Stouffer, D. B. & Bascompte, J. Compartmentalization increases food-web persistence. Proc. Natl. Acad. Sci. 108, 3648–3652 (2011).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Guimaraes, P. R. Jr. et al. Interaction intimacy affects structure and coevolutionary dynamics in mutualistic networks. Curr. Biol. 17, 1797–1803 (2007).Article 
    CAS 
    PubMed 

    Google Scholar 
    Valverde, S. et al. The architecture of mutualistic networks as an evolutionary spandrel. Nature Ecol. Evol. 2, 94–99 (2018).Article 

    Google Scholar 
    Staniczenko, P. P., Kopp, J. C. & Allesina, S. The ghost of nestedness in ecological networks. Nat. Commun. 4, 1–6 (2013).Article 

    Google Scholar 
    Vázquez, D. P., Blüthgen, N., Cagnolo, L. & Chacoff, N. P. Uniting pattern and process in plant-animal mutualistic networks: A review. Ann. Bot. 103, 1445–1457 (2009).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Mello, M. A. et al. Insights into the assembly rules of a continent-wide multilayer network. Nature Ecol. Evol. 3, 1525–1532 (2019).Article 

    Google Scholar 
    Felix, G. M., Pinheiro, R. B., Jorge, L. R. & Lewinsohn, T. M. A framework for hierarchical compound topologies in species interaction networks. Oikos 2022, 9538 (2022).Article 

    Google Scholar 
    Valverde, S. et al. Coexistence of nestedness and modularity in host-pathogen infection networks. Nature Ecol. Evol. 4, 568–577 (2020).Article 

    Google Scholar 
    Hui, C. & Richardson, D. M. How to invade an ecological network. Trends Ecol. Evol. 34, 121–131 (2019).Article 
    PubMed 

    Google Scholar 
    Layman, C. A., Quattrochi, J. P., Peyer, C. M. & Allgeier, J. E. Niche width collapse in a resilient top predator following ecosystem fragmentation. Ecol. Lett. 10, 937–944 (2007).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Vidiella, B., Fontich, E., Valverde, S. & Sardanyés, J. Habitat loss causes long extinction transients in small trophic chains. Thyroid Res. 14, 641–661 (2021).
    Google Scholar 
    Donohue, I. et al. Navigating the complexity of ecological stability. Ecol. Lett. 19, 1172–1185 (2016).Article 
    PubMed 

    Google Scholar 
    Krause, A. E., Frank, K. A., Mason, D. M., Ulanowicz, R. E. & Taylor, W. W. Compartments revealed in food-web structure. Nature 426, 282–285 (2003).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Hagberg, A., Swart, P. & S Chult, D. Exploring network structure, dynamics, and function using networkx. In Tech. Rep., Los Alamos National Lab.(LANL), Los Alamos (2008).Chimani, M. et al. The open graph drawing framework (ogdf). Handb. Graph Draw. Visual. 2011, 543–569 (2013).
    Google Scholar 
    Hachul, S. & Jünger, M. Drawing large graphs with a potential-field-based multilevel algorithm. In Graph Drawing: 12th International Symposium, GD 2004, New York, NY, USA, September 29-October 2, 2004, Revised Selected Papers 12, 285–295 (Springer, 2005).Raghavan, U. N., Albert, R. & Kumara, S. Near linear time algorithm to detect community structures in large-scale networks. Phys. Rev. E 76, 036106 (2007).Article 
    ADS 

    Google Scholar 
    Newman, M. E. & Girvan, M. Finding and evaluating community structure in networks. Phys. Rev. E 69, 026113 (2004).Article 
    ADS 
    CAS 

    Google Scholar 
    Barber, M. J. Modularity and community detection in bipartite networks. Phys. Rev. E 76, 066102 (2007).Article 
    ADS 
    MathSciNet 

    Google Scholar 
    Pesántez-Cabrera, P. & Kalyanaraman, A. Efficient detection of communities in biological bipartite networks. IEEE/ACM Trans. Comput. Biol. Bioinf. 16, 258–271 (2017).Article 

    Google Scholar 
    Almeida-Neto, M. & Ulrich, W. A straightforward computational approach for measuring nestedness using quantitative matrices. Environ. Model. Softw. 26, 173–178 (2011).Article 

    Google Scholar 
    Latora, V. & Marchiori, M. Efficient behavior of small-world networks. Phys. Rev. Lett. 87, 198701 (2001).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Borgatti, S. P. & Halgin, D. S. Analyzing affiliation networks. Sage Handb. Soc. Netw. Anal. 1, 417–433 (2011).
    Google Scholar 
    Roopnarine, P. D. Extinction cascades and catastrophe in ancient food webs. Paleobiology 32, 1–19 (2006).Article 

    Google Scholar 
    Pires, M. M. et al. The indirect paths to cascading effects of extinctions in mutualistic networks. Ecology 101(7), e03080 https://doi.org/10.1002/ecy.3080 (2020).Article 
    PubMed 

    Google Scholar 
    Mestres, J., Gregori-Puigjane, E., Valverde, S. & Sole, R. V. Data completeness-the achilles heel of drug-target networks. Nat. Biotechnol. 26, 983–984 (2008).Article 
    CAS 
    PubMed 

    Google Scholar 
    Milo, R. et al. Network motifs: Simple building blocks of complex networks. Science 298, 824–827 (2002).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Sarzynska, M., Leicht, E. A., Chowell, G. & Porter, M. A. Null models for community detection in spatially embedded, temporal networks. J. Complex Netw. 4, 363–406 (2016).Article 
    MathSciNet 

    Google Scholar  More

  • in

    Horses discriminate human body odors between fear and joy contexts in a habituation-discrimination protocol

    Semin, G. R., Scandurra, A., Baragli, P., Lanatà, A. & D’Aniello, B. Inter- and intra-species communication of emotion: Chemosignals as the neglected medium. Animals 9, 887 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Désiré, L., Boissy, A. & Veissier, I. Emotions in farm animals: A new approach to animal welfare in applied ethology. Behav. Processes 60, 165–180 (2002).Article 
    PubMed 

    Google Scholar 
    Briefer, E. F. & Le Comber, S. Vocal expression of emotions in mammals: mechanisms of production and evidence. J. Zool. 288, 1–20 (2012).Article 

    Google Scholar 
    Jardat, P. & Lansade, L. Cognition and the human–animal relationship: a review of the sociocognitive skills of domestic mammals toward humans. Anim. Cogn. 25, 369–384 (2022).Article 
    PubMed 

    Google Scholar 
    Galvan, M. & Vonk, J. Man’s other best friend: domestic cats (F. silvestris catus) and their discrimination of human emotion cues. Anim. Cogn. 19, 193–205 (2016).Nawroth, C., Albuquerque, N., Savalli, C., Single, M. S. & McElligott, A. G. Goats prefer positive human emotional facial expressions. R. Soc. Open Sci. 5, 180491 (2018).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Proops, L., Grounds, K., Smith, A. V. & McComb, K. Animals remember previous facial expressions that specific humans have exhibited. Curr. Biol. 28, 1428-1432.e4 (2018).Article 
    CAS 
    PubMed 

    Google Scholar 
    Smith, A. V., Proops, L., Grounds, K., Wathan, J. & McComb, K. Functionally relevant responses to human facial expressions of emotion in the domestic horse (Equus caballus). Biol. Lett. 12, 20150907 (2016).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Siniscalchi, M., D’Ingeo, S., Fornelli, S. & Quaranta, A. Lateralized behavior and cardiac activity of dogs in response to human emotional vocalizations. Sci. Rep. 8, 77 (2018).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Siniscalchi, M., D’Ingeo, S. & Quaranta, A. Orienting asymmetries and physiological reactivity in dogs’ response to human emotional faces. Learn. Behav. 46, 574–585 (2018).Article 
    PubMed 

    Google Scholar 
    Smith, A. V. et al. Domestic horses (Equus caballus) discriminate between negative and positive human nonverbal vocalisations. Sci. Rep. 8, 13052 (2018).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Trösch, M. et al. Horses categorize human emotions cross-modally based on facial expression and non-verbal vocalizations. Animals 9, 862 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Quaranta, A., D’ingeo, S., Amoruso, R. & Siniscalchi, M. Emotion recognition in cats. Animals 10, 1107 (2020).Nakamura, K., Takimoto-Inose, A. & Hasegawa, T. Cross-modal perception of human emotion in domestic horses (Equus caballus). Sci. Rep. 8, 8660 (2018).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Albuquerque, N. et al. Dogs recognize dog and human emotions. Biol. Lett. 12, 20150883 (2016).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Briefer, E. F. Vocal contagion of emotions in non-human animals. Proc. R. Soc. B Biol. Sci. 285 (2018).Sabiniewicz, A., Tarnowska, K., Świątek, R., Sorokowski, P. & Laska, M. Olfactory-based interspecific recognition of human emotions: Horses (Equus ferus caballus) can recognize fear and happiness body odour from humans (Homo sapiens). Appl. Anim. Behav. Sci. 230, 105072 (2020).Article 

    Google Scholar 
    Baba, C., Kawai, M. & Takimoto-Inose, A. Are horses (Equus caballus) sensitive to human emotional cues?. Animals 9, 630 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Brennan, P. A. & Kendrick, K. M. Mammalian social odours: attraction and individual recognition. Philos. Trans. R. Soc. B Biol. Sci. 361, 2061–2078 (2006).Saslow, C. A. Understanding the perceptual world of horses. Appl. Anim. Behav. Sci. 78, 209–224 (2002).Article 

    Google Scholar 
    Péron, F., Ward, R. & Burman, O. Horses (Equus caballus) discriminate body odour cues from conspecifics. Anim. Cogn. 17, 1007–1011 (2014).Article 
    PubMed 

    Google Scholar 
    Krueger, K. & Flauger, B. Olfactory recognition of individual competitors by means of faeces in horse (Equus caballus). Anim. Cogn. 14, 245–257 (2011).Article 
    PubMed 

    Google Scholar 
    Boissy, A., Terlouw, C. & Le Neindre, P. Presence of cues from stressed conspecifics increases reactivity to aversive events in cattle: evidence for the existence of alarm substances in urine. Physiol. Behav. 63, 489–495 (1998).Article 
    CAS 
    PubMed 

    Google Scholar 
    Vieuille-Thomas, C. & Signoret, J. P. Pheromonal transmission of an aversive experience in domestic pig. J. Chem. Ecol. 18, 1551–1557 (1992).Article 
    CAS 
    PubMed 

    Google Scholar 
    Siniscalchi, M., D’Ingeo, S. & Quaranta, A. The dog nose ‘KNOWS’ fear: Asymmetric nostril use during sniffing at canine and human emotional stimuli. Behav. Brain Res. 304, 34–41 (2016).Article 
    PubMed 

    Google Scholar 
    Calvi, E. et al. The scent of emotions: A systematic review of human intra- and interspecific chemical communication of emotions. Brain Behav. 10 (2020).de Groot, J. H. B., Semin, G. R. & Smeets, M. A. M. On the communicative function of body odors: A theoretical integration and review. Perspect. Psychol. Sci. 12, 306–324 (2017).Article 
    PubMed 

    Google Scholar 
    de Groot, J. H. B. et al. A sniff of happiness. Psychol. Sci. 26, 684–700 (2015).Article 
    PubMed 

    Google Scholar 
    Destrez, A. et al. Male mice and cows perceive human emotional chemosignals: A preliminary study. Anim. Cogn. 24, 1205–1214 (2021).Article 
    PubMed 

    Google Scholar 
    Wilson, Id. Dogs can discriminate between human baseline and psychological stress condition odours. PLoS ONE 17, e0274143 (2022).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    D’Aniello, B., Semin, G. R., Alterisio, A., Aria, M. & Scandurra, A. Interspecies transmission of emotional information via chemosignals: from humans to dogs (Canis lupus familiaris). Anim. Cogn. 21, 67–78 (2018).Article 
    PubMed 

    Google Scholar 
    D’Aniello, B. et al. Sex differences in the behavioral responses of dogs exposed to human chemosignals of fear and happiness. Anim. Cogn. 24, 299–309 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Siniscalchi, M., d’Ingeo, S., Quaranta, A., D’Ingeo, S. & Quaranta, A. Lateralized emotional functioning in domestic animals. Appl. Anim. Behav. Sci. 237, 105282 (2021).Article 

    Google Scholar 
    Rogers, L. & Vallortigara, G. Lateralized Brain Functions: Methods in Human and Non-Human Species. vol. 122 (2017).D’Ingeo, S. et al. Horses associate individual human voices with the valence of past interactions: A behavioural and electrophysiological study. Sci. Rep. 9, 11568 (2019).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Siniscalchi, M., Padalino, B., Aubé, L. & Quaranta, A. Right-nostril use during sniffing at arousing stimuli produces higher cardiac activity in jumper horses. Laterality 20, (2015).De Boyer Des Roches, A., Richard-Yris, M.-A. A., Henry, S., Ezzaouïa, M. & Hausberger, M. Laterality and emotions: Visual laterality in the domestic horse (Equus caballus) differs with objects’ emotional value. Physiol. Behav. 94, 487–490 (2008).Albrecht, J. et al. Smelling chemosensory signals of males in anxious versus nonanxious condition increases state anxiety of female subjects. Chem. Senses 36, 19–27 (2011).Article 
    CAS 
    PubMed 

    Google Scholar 
    Derrickson, S. Sinister (VF)—YouTube. (Wild Bunch SA, 2001).Friard, O. & Gamba, M. BORIS: a free, versatile open-source event-logging software for video/audio coding and live observations. Methods Ecol. Evol. 7, 1325–1330 (2016).Article 

    Google Scholar 
    R Core Team. R: A language and environment for statistical computing (2021).Wickham, H. Ggplot2: Elegant graphics for data analysis. (2016).Hothorn, T., Winell, H., Hornik, K., van de Wiel, M. A. & Zeileis, A. coin: Conditional inference procedures in a permutation test framework (2021).Brooks, M. E. et al. glmmTMB balances speed and flexibility among packages for zero-inflated generalized linear mixed modeling. R J. 9, 378–400 (2017).Article 

    Google Scholar 
    Hartig, F. DHARMa: Residual diagnostics for hierarchical (multi-level/mixed) regression models (2021).Lenth, R. V. emmeans: Estimated Marginal Means, aka Least-Squares Means. (2022).Hothersall, B., Harris, P., Sörtoft, L. & Nicol, C. J. Discrimination between conspecific odour samples in the horse (Equus caballus). Appl. Anim. Behav. Sci. 126, 37–44 (2010).Article 

    Google Scholar 
    Smeets, M. A. M. et al. Chemical fingerprints of emotional body odor. Metabolites 10, (2020).Sabiniewicz, A. et al. A preliminary investigation of interspecific chemosensory communication of emotions: Can Humans (Homo sapiens) recognise fear- and non-fear body odour from horses (Equus ferus caballus). Animal 11, 3499 (2021).Article 

    Google Scholar 
    Zhou, W. & Chen, D. Entangled chemosensory emotion and identity: Familiarity enhances detection of chemosensorily encoded emotion. Soc. Neurosci. 6, 270–276 (2011).Article 
    PubMed 

    Google Scholar 
    Starling, M., McLean, A. & McGreevy, P. The contribution of equitation science to minimising horse-related risks to humans. Animal 6, 15 (2016).Article 

    Google Scholar 
    Basile, M. et al. Socially dependent auditory laterality in domestic horses (Equus caballus). Anim. Cogn. 12, 611–619 (2009).Article 
    PubMed 

    Google Scholar 
    Hatfield, E., Cacioppo, J. T. & Rapson, R. L. Emotional contagion. Curr. Dir. Psychol. Sci. 2, 96–100 (1993).Article 

    Google Scholar 
    Austin, N. P. & Rogers, L. J. Limb preferences and lateralization of aggression, reactivity and vigilance in feral horses Equus caballus. Anim. Behav. 83, 239–247 (2012).Article 

    Google Scholar 
    Larose, C., Richard-Yris, M.-A., Hausberger, M. & Rogers, L. J. Laterality of horses associated with emotionality in novel situations Laterality Asymmetries Body. Brain Cogn. 11, 355–367 (2006).
    Google Scholar 
    Farmer, K., Krüger, K., Byrne, R. W. & Marr, I. Sensory laterality in affiliative interactions in domestic horses and ponies (Equus caballus). Anim. Cogn. 21, 631–637 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Chen, D. & Haviland-Jones, J. Human olfactory communication of emotion. Percept. Mot. Skills 91, 771–781 (2000).Article 
    CAS 
    PubMed 

    Google Scholar 
    de Groot, J. H. B., Semin, G. R. & Smeets, M. A. M. Chemical communication of fear: A case of male-female asymmetry. J. Exp. Psychol. Gen. 143, 1515–1525 (2014).Article 
    PubMed 

    Google Scholar 
    Marinier, S. L., Alexander, A. J. & Waring, G. H. Flehmen behaviour in the domestic horse: Discrimination of conspecific odours. Appl. Anim. Behav. Sci. 19, 227–237 (1988).Article 

    Google Scholar 
    Lansade, L., Pichard, G. & Leconte, M. Sensory sensitivities: Components of a horse’s temperament dimension. Appl. Anim. Behav. Sci. 114, 534–553 (2008).Article 

    Google Scholar 
    Lansade, L., Bouissou, M. F. & Erhard, H. W. Fearfulness in horses: A temperament trait stable across time and situations. Appl. Anim. Behav. Sci. 115, 182–200 (2008).Article 

    Google Scholar 
    Hoenen, M., Wolf, O. T. & Pause, B. M. The impact of stress on odor perception. https://doi.org/10.1177/030100661668870746,366-376 (2017).Article 
    PubMed 

    Google Scholar 
    Rørvang, M. V., Nicova, K. & Yngvesson, J. Horse odor exploration behavior is influenced by pregnancy and age. Front. Behav. Neurosci. 16, 295 (2022).Article 

    Google Scholar 
    Doty, R. L. & Cameron, E. L. Sex differences and reproductive hormone influences on human odor perception. Physiol. Behav. 97, 213–228 (2009).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar  More

  • in

    Climate-induced range shifts drive adaptive response via spatio-temporal sieving of alleles

    Study populations and sequencing strategyDNA libraries were prepared for 1261 D. sylvestris individuals from 115 populations (5–20 individuals per population) under a modified protocol49 of the Illumina Nextera DNA library preparation kit (Supplementary Methods S1.1, Supplementary Data 1). Individuals were indexed with unique dual-indexes (IDT Illumina Nextera 10nt UDI – 384 set) from Integrated DNA Technologies Co, to avoid index-hopping50. Libraries were sequenced (150 bp paired-end sequencing) in four lanes of an Illumina NovaSeq 6000 machine at Novogene Co. This resulted in an average coverage of ca. 2x per individual. Sequenced individuals were trimmed for adapter sequences (Trimmomatic version 0.3551), mapped (BWA-MEM version 0.7.1752,53) against a reference assembly54 (ca. 440 Mb), had duplicates marked and removed (Picard Toolkit version 2.0.1; http://broadinstitute.github.io/picard), locally realigned around indels (GATK version 3.555), recalibrated for base quality scores (ATLAS version 0.956) and had overlapping read pairs clipped (bamUtil version 1.0.1457) (Supplementary Methods S1.1). Population genetic analyses were performed on the resultant BAM files via genotype likelihoods (ANGSD version 0.93358 and ATLAS versions 0.9–1.056), to accommodate the propagation of uncertainty from the raw sequence data to population genetic inference.Population genetic structure and biogeographic barriersTo investigate the genetic structure of our samples (Fig. 2A, Supplementary Fig. S2), we performed principal component analyses (PCA) on all 1261 samples (“full” dataset) via PCAngsd version 0.9859, following conversion of the mapped sequence data to ANGSD genotype likelihoods in Beagle format (Supplementary Methods S1.2). To visualise PCA results in space (Supplementary Fig. S4), individuals’ principal components were projected on a map, spatially interpolated (linear interpolation, akima R package version 0.6.260) and had the first two principal components represented as green and blue colour channels. Given that uneven sampling can bias the inference of structure in PCA, PCA was also performed on a balanced dataset comprising a common, down-sampled size of 125 individuals per geographic region (“balanced” dataset; Fig. 2B, Supplementary Fig. S3; Supplementary Methods S1.2; Supplementary Data 1). Individual admixture proportions and ancestral allele frequencies were estimated using PCAngsd (-admix model) for K = 2–6, using the balanced dataset to avoid potential biases related to imbalanced sampling22,23 and an automatic search for the optimal sparseness regularisation parameter (alpha) soft-capped to 10,000 (Supplementary Methods S1.2). To visualise ancestry proportions in space, population ancestry proportions were spatially interpolated (kriging) via code modified from Ref. 61 (Supplementary Fig. S5).To test if between-lineage admixture underlies admixture patterns inferred by PCAngsd or if the data is better explained by alternative scenarios such as recent bottlenecks, we used chromosome painting and patterns of allele sharing to construct painting palettes via the programmes MixPainter and badMIXTURE (unlinked model)28 and compared this to the PCAngsd-inferred palettes (Fig. 2B, C; Supplementary Methods S1.2). We referred to patterns of residuals between these palettes to inform of the most likely underlying demographic scenario. For assessing Alpine–Balkan palette residuals (and hence admixture), 65 individuals each from the French Alps (inferred as pure Alpine ancestry in PCAngsd), Monte Baldo (inferred with both Alpine and Balkan ancestries in PCAngsd) and Julian Alps (inferred as pure Balkan ancestry in PCAngsd) were analysed under K = 2 in PCAngsd and badMIXTURE (Fig. 2C). For assessing Apennine–Balkan admixture, 22 individuals each from the French pre-Alps (inferred as pure Apennine ancestry in PCAngsd), Tuscany (inferred with both Apennine and Balkan ancestries in PCAngsd) and Julian Alps (inferred as pure Balkan ancestry in PCAngsd) were analysed under K = 2 in PCAngsd and badMIXTURE.To construct a genetic distance tree (Supplementary Fig. S1), we first calculated pairwise genetic distances between 549 individuals (5 individuals per population for all populations) using ATLAS, employing a distance measure (weight) reflective of the number of alleles differing between the genotypes (Supplementary Methods S1.2; Supplementary Data 1). A tree was constructed from the resultant distance matrix via an initial topology defined by the BioNJ algorithm with subsequent topological moves performed via Subtree Pruning and Regrafting (SPR) in FastME version 2.1.6.162. This matrix of pairwise genetic distances was also used as input for analyses of effective migration and effective diversity surfaces in EEMS25. EEMS was run setting the number of modelled demes to 1000 (Fig. 2A, Supplementary Fig. S8). For each case, ten independent Markov chain Monte Carlo (MCMC) chains comprising 5 million iterations each were run, with a 1 million iteration burn-in, retaining every 10,000th iteration. Biogeographic barriers (Fig. 2A, Supplementary Fig. S7) were further identified via applying Monmonier’s algorithm24 on a valuated graph constructed via Delauney triangulation of population geographic coordinates, with edge values reflecting population pairwise FST; via the adegenet R package version 2.1.163. FST between all population pairs were calculated via ANGSD, employing a common sample size of 5 individuals per population (Supplementary Fig. S6; Supplementary Methods S1.2; Supplementary Data 1). 100 bootstrap runs were performed to generate a heatmap of genetic boundaries in space, from which a weighted mean line was drawn (Supplementary Fig. S7). All analyses in ANGSD were performed with the GATK (-GL 2) model, as we noticed irregularities in the site frequency spectra (SFS) with the SAMtools (-GL 1) model similar to that reported in Ref. 58 with particular BAM files. All analyses described above were performed on the full genome.Ancestral sequence reconstructionTo acquire ancestral states and polarise site-frequency spectra for use in the directionality index ψ and demographic inference, we reconstructed ancestral genome sequences at each node of the phylogenetic tree of 9 Dianthus species: D. carthusianorum, D. deltoides, D. glacialis, D. sylvestris (Apennine lineage), D. lusitanus, D. pungens, D. superbus alpestris, D. superbus superbus, and D. sylvestris (Alpine lineage). This tree topology was extracted from a detailed reconstruction of Dianthus phylogeny based on 30 taxa by Fior et al. (Fior, Luqman, Scharmann, Zemp, Zoller, Pålsson, Gargano, Wegmann & Widmer; paper in preparation) (Supplementary Methods S1.3). For ancestral sequence reconstruction, one individual per species was sequenced at medium coverage (ca. 10x), trimmed (Trimmomatic), mapped against the D. sylvestris reference assembly (BWA-MEM) and had overlapping read pairs clipped (bamUtil) (Supplementary Methods S1.3). For each species, we then generated a species-specific FASTA using GATK FastaAlternateReferenceMaker. This was achieved by replacing the reference bases at polymorphic sites with species-specific variants as identified by freebayes64 (version 1.3.1; default parameters), while masking (i.e., setting as “N”) sites (i) with zero depth and (ii) that didn’t pass the applied variant filtering criteria (i.e., that are not confidently called as polymorphic; Supplementary Methods S1.3). Species FASTA files were then combined into a multi-sample FASTA. Using this, we probabilistically reconstructed ancestral sequences at each node of the tree via PHAST (version 1.4) prequel65, using a tree model produced by PHAST phylofit under a REV substitution model and the specified tree topology (Supplementary Methods S1.3). Ancestral sequence FASTA files were then generated from the prequel results using a custom script.Expansion signalTo calculate the population pairwise directionality index ψ for the Alpine lineage, we utilised equation 1b from Peter and Slatkin (2013)31, which defines ψ in terms of the two-population site frequency spectrum (2D-SFS) (Supplementary Methods S1.4). 2D-SFS between all population pairs (10 individuals per population; Supplementary Data 1) were estimated via ANGSD and realSFS66 (Supplementary Methods S1.4), for unfolded spectra. Unfolding of spectra was achieved via polarisation with respect to the ancestral state of sites defined at the D. sylvestris (Apennine lineage) – D. sylvestris (Alpine lineage) ancestral node. Correlation of pairwise ψ and (great-circle) distance matrices was tested via a Mantel test (10,000 permutations). To infer the geographic origin of the expansion (Fig. 3), we employed a time difference of arrival (TDOA) algorithm following Peter and Slatkin (2013);31 performed via the rangeExpansion R package version 0.0.0.900031,67. We further estimated the strength of the founder of this expansion using the same package.Demographic inferenceTo evaluate the demographic history of D. sylvestris, a set of candidate demographic models was formulated. To constrain the topology of tested models, we first inferred the phylogenetic tree of the three identified evolutionary lineages of D. sylvestris (Alpine, Apennine and Balkan) as embedded within the larger phylogeny of the Eurasian Dianthus clade (note that the phylogeny from Fior et al. (Fior, Luqman, Scharmann, Zemp, Zoller, Pålsson, Gargano, Wegmann & Widmer; paper in preparation) excludes Balkan representatives of D. sylvestris). Trees were inferred based on low-coverage whole-genome sequence data of 1–2 representatives from each D. sylvestris lineage, together with whole-genome sequence data of 7 other Dianthus species, namely D. carthusianorum, D. deltoides, D. glacialis, D. lusitanus, D. pungens, D. superbus alpestris and D. superbus superbus, that were used to root the D. sylvestris clade (Supplementary Methods S1.5). We estimated distance-based phylogenies using ngsDist68 that accommodates genotype likelihoods in the estimation of genetic distances (Supplementary Methods S1.5). Genetic distances were calculated via two approaches: (i) genome-wide and (ii) along 10 kb windows. For the former, 110 bootstrap replicates were calculated by re-sampling over similar-sized genomic blocks. For the alternative strategy based on 10 kb windows, window trees were combined using ASTRAL-III version 5.6.369 to generate a genome-wide consensus tree accounting for potential gene tree discordance (Supplementary Methods S1.5). Trees were constructed from matrices of genetic distances from initial topologies defined by the BioNJ algorithm with subsequent topological moves performed via Subtree Pruning and Regrafting (SPR) in FastME version 2.1.6.162. We rooted all resultant phylogenetic trees with D. deltoides as the outgroup70. Both approaches recovered a topology with the Balkan lineage diverging prior to the Apennine and Alpine lineages (Supplementary Fig. S9). This taxon topology for D. sylvestris was supported by high ASTRAL-III posterior probabilities ( >99%), ASTRAL-III quartet scores ( >0.5) and bootstrap values ( >99%). Topologies deeper in the tree were less well-resolved (with quartet scores More