More stories

  • in

    The influence of social cues on timing of animal migrations

    Alerstam, T., Hedenström, A. & Åkesson, S. Long-distance migration: evolution and determinants. Oikos 103, 247–260 (2003).Article 

    Google Scholar 
    Bauer, S., Lisovski, S. & Hahn, S. Timing is crucial for consequences of migratory connectivity. Oikos 125, 605–612 (2016).Article 

    Google Scholar 
    Bauer, S. & Hoye, B. J. Migratory animals couple biodiversity and ecosystem functioning worldwide. Science 344, 1242552 (2014).Article 
    CAS 
    PubMed 

    Google Scholar 
    Fricke, E. C., Ordonez, A., Rogers, H. S. & Svenning, J. C. The effects of defaunation on plants’ capacity to track climate change. Science 214, 210–214 (2022).Article 

    Google Scholar 
    Tucker, M. A. et al. Moving in the Anthropocene: global reductions in terrestrial mammalian movements. Science 359, 466–469 (2018).Article 
    CAS 
    PubMed 

    Google Scholar 
    Wilcove, D. S. & Wikelski, M. Going, going, gone: is animal migration disappearing? PLoS Biol. 6, e188 (2008).Article 
    PubMed 
    PubMed Central 

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

    Google Scholar 
    Walther, G. et al. Ecological responses to recent climate change. Nature 4126, 389–395 (2002).Article 

    Google Scholar 
    Teitelbaum, C. S. et al. Experience drives innovation of new migration patterns of whooping cranes in response to global change. Nat. Commun. 7, 12793 (2016).Article 
    CAS 
    PubMed 
    PubMed Central 

    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 
    Senzaki, M. et al. Sensory pollutants alter bird phenology and fitness across a continent. Nature 587, 605–609 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Guerra, A. S. Wolves of the sea: managing human–wildlife conflict in an increasingly tense ocean. Mar. Policy 99, 369–373 (2019).Article 

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

    Google Scholar 
    Both, C., Bouwhuis, S., Lessells, C. M. & Visser, M. E. Climate change and population declines in a long-distance migratory bird. Nature 441, 81–83 (2006).Article 
    CAS 
    PubMed 

    Google Scholar 
    Post, E. & Forchhammer, M. C. Climate change reduces reproductive success of an Arctic herbivore through trophic mismatch. Phil. Trans. R. Soc. B Biol. Sci. 363, 2369–2375 (2008).Article 

    Google Scholar 
    Winkler, D. W. et al. Cues, strategies, and outcomes: how migrating vertebrates track environmental change. Mov. Ecol. 2, 10 (2014).Article 

    Google Scholar 
    Xu, W. et al. The plasticity of ungulate migration in a changing world. Ecology 102, e03293 (2021).Article 
    PubMed 

    Google Scholar 
    McNamara, J. M., Barta, Z., Klaassen, M. & Bauer, S. Cues and the optimal timing of activities under environmental change. Ecol. Lett. 14, 1183–1190 (2011).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bauer, S., McNamara, J. M. & Barta, Z. Environmental variability, reliability of information and the timing of migration. Proc. R. Soc. B Biol. Sci. 287, 20200622 (2020).Article 

    Google Scholar 
    Abrahms, B. et al. Emerging perspectives on resource tracking and animal movement ecology. Trends Ecol. Evol. 36, 308–320 (2020).Article 
    PubMed 

    Google Scholar 
    Visser, M. E., Holleman, L. J. M. & Gienapp, P. Shifts in caterpillar biomass phenology due to climate change and its impact on the breeding biology of an insectivorous bird. Oecologia 147, 164–172 (2006).Article 
    PubMed 

    Google Scholar 
    Aikens, E. O. et al. Wave-like patterns of plant phenology determine ungulate movement tactics. Curr. Biol. 30, 3444–3449 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Abrahms, B. et al. Memory and resource tracking drive blue whale migrations. Proc. Natl Acad. Sci. USA 116, 5582–5587 (2019).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Lank, D. B., Butler, R. W., Ireland, J. & Ydenberg, R. C. Effects of predation danger on migration strategies of sandpipers. Oikos 103, 303–319 (2003).Article 

    Google Scholar 
    Sabal, M. C. et al. Predation landscapes influence migratory prey ecology and evolution. Trends Ecol. Evol. 36, 737–749 (2021).Article 
    PubMed 

    Google Scholar 
    Furey, N. B., Armstrong, J. B., Beauchamp, D. A. & Hinch, S. G. Migratory coupling between predators and prey. Nat. Ecol. Evol. 2, 1846–1853 (2018).Article 
    PubMed 

    Google Scholar 
    Altizer, S., Bartel, R. & Han, B. A. Animal migration and infectious disease risk. Science 331, 296–302 (2011).Article 
    CAS 
    PubMed 

    Google Scholar 
    Gunnarsson, T., Gill, J., Sigurbjörnsson, T. & Sutherland, W. Arrival synchrony in migratory birds. Nature 431, 646 (2004).Article 
    CAS 
    PubMed 

    Google Scholar 
    Beltran, R. S. et al. Elephant seals time their long-distance migration using a map sense. Curr. Biol. 32, R156–R157 (2022).Article 
    CAS 
    PubMed 

    Google Scholar 
    Yang, L. H. & Rudolf, V. H. W. Phenology, ontogeny and the effects of climate change on the timing of species interactions. Ecol. Lett. 13, 1–10 (2010).Article 
    CAS 
    PubMed 

    Google Scholar 
    Visser, M. E. & Gienapp, P. Evolutionary and demographic consequences of phenological mismatches. Nat. Ecol. Evol. 3, 879–885 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Furey, N. B. et al. Predator swamping reduces predation risk during nocturnal migration of juvenile salmon in a high-mortality landscape. J. Anim. Ecol. 85, 948–959 (2016).Article 
    PubMed 

    Google Scholar 
    Rickbeil, G. J. M. et al. Plasticity in elk migration timing is a response to changing environmental conditions. Glob. Change Biol. 25, 2368–2381 (2019).Article 

    Google Scholar 
    Schmaljohann, H. & Both, C. The limits of modifying migration speed to adjust to climate change. Nat. Clim. Change 7, 573–576 (2017).Article 

    Google Scholar 
    Gwinner, E. Circadian and circannual programmes in avian migration. J. Exp. Biol. 199, 39–48 (1996).Article 
    CAS 
    PubMed 

    Google Scholar 
    Liedvogel, M., Åkesson, S. & Bensch, S. The genetics of migration on the move. Trends Ecol. Evol. 26, 561–569 (2011).Article 
    PubMed 

    Google Scholar 
    Hauser, D. D. W. et al. Decadal shifts in autumn migration timing by Pacific Arctic beluga whales are related to delayed annual sea ice formation. Glob. Change Biol. 23, 2206–2217 (2017).Article 

    Google Scholar 
    Palacín, C., Alonso, J. C., Alonso, J. A., Magaña, M. & Martín, C. A. Cultural transmission and flexibility of partial migration patterns in a long-lived bird, the great bustard Otis tarda. J. Avian Biol. 42, 301–308 (2011).Article 

    Google Scholar 
    Couzin, I. D. Collective animal migration. Curr. Biol. 28, R976–R980 (2018).Article 
    CAS 
    PubMed 

    Google Scholar 
    Guttal, V. & Couzin, I. D. Social interactions, information use, and the evolution of collective migration. Proc. Natl Acad. Sci. USA 107, 16172–16177 (2010).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Berdahl, A. M. et al. Collective animal navigation and migratory culture: from theoretical models to empirical evidence. Phil. Trans. R. Soc. B Biol. Sci. 373, 20170009 (2018).Article 

    Google Scholar 
    Cohen, E. B. & Satterfield, D. A. ‘Chancing on a spectacle:’ co-occurring animal migrations and interspecific interactions. Ecography 43, 1657–1671 (2020).Article 

    Google Scholar 
    Berdahl, A., Torney, C. J., Ioannou, C. C., Faria, J. J. & Couzin, I. D. Emergent sensing of complex environments by mobile animal groups. Science 339, 574–576 (2013).Article 
    CAS 
    PubMed 

    Google Scholar 
    Abrahms, B., Teitelbaum, C. S., Mueller, T. & Converse, S. J. Ontogenetic shifts from social to experiential learning drive avian migration timing. Nat. Commun. 12, 7326 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Sasaki, T. & Biro, D. Cumulative culture can emerge from collective intelligence in animal groups. Nat. Commun. 8, 15049 (2017).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Helm, B., Piersma, T. & van der Jeugd, H. Sociable schedules: interplay between avian seasonal and social behaviour. Anim. Behav. 72, 245–262 (2006).Article 

    Google Scholar 
    Piersma, T., Zwarts, L. & Bruggemann, J. H. Behavioural aspects of the departure of waders before long-distance flights: flocking, vocalizations, flight paths and diurnal timing. Ardea 78, 157–184 (1990).
    Google Scholar 
    Dingle, H. & Drake, V. A. What is migration? BioScience 57, 113–121 (2007).Article 

    Google Scholar 
    Oestreich, W. K. & Aiu, K. M. Code and data from: The influence of social cues on timing of animal migrations. Zenodo https://zenodo.org/record/6574762 (2022).Furey, N. B., Martins, E. G. & Hinch, S. G. Migratory salmon smolts exhibit consistent interannual depensatory predator swamping: effects on telemetry-based survival estimates. Ecol. Freshw. Fish 30, 18–30 (2021).Article 

    Google Scholar 
    Berdahl, A., Westley, P. A. H. & Quinn, T. P. Social interactions shape the timing of spawning migrations in an anadromous fish. Anim. Behav. 126, 221–229 (2017).Article 

    Google Scholar 
    Louca, V., Lindsay, S. W. & Lucas, M. C. Factors triggering floodplain fish emigration: importance of fish density and food availability. Ecol. Freshw. Fish 18, 60–64 (2009).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bastille-Rousseau, G. et al. Migration triggers in a large herbivore: Galápagos giant tortoises navigating resource gradients on volcanoes. Ecology 100, e02658 (2019).Article 
    PubMed 

    Google Scholar 
    Bracis, C. & Mueller, T. Memory, not just perception, plays an important role in terrestrial mammalian migration. Proc. R. Soc. B Biol. Sci. 284, 20170449 (2017).Article 

    Google Scholar 
    Barrett, B., Zepeda, E., Pollack, L., Munson, A. & Sih, A. Counter-culture: does social learning help or hinder adaptive response to human-induced rapid environmental change? Front. Ecol. Evol. 7, 183 (2019).Article 

    Google Scholar 
    Merkle, J. A. et al. Site fidelity as a maladaptive behavior in the Anthropocene. Front. Ecol. Environ. 20, 187–194 (2022).Article 

    Google Scholar 
    Teske, P. R. et al. The sardine run in southeastern Africa is a mass migration into an ecological trap. Sci. Adv. 7, eabf4514 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Corten, A. The role of ‘conservatism’ in herring migrations. Rev. Fish Biol. Fish. 11, 339–361 (2002).Article 

    Google Scholar 
    Mukhin, A., Chernetsov, N. & Kishkinev, D. Acoustic information as a distant cue for habitat recognition by nocturnally migrating passerines during landfall. Behav. Ecol. 19, 716–723 (2008).Article 

    Google Scholar 
    Barker, K. J. et al. Toward a new framework for restoring lost wildlife migrations. Conserv. Lett. 15, e12850 (2022).Article 

    Google Scholar 
    Teitelbaum, C. S., Converse, S. J. & Mueller, T. The importance of early life experience and animal cultures in reintroductions. Conserv. Lett. 12, e12599 (2019).Article 

    Google Scholar 
    Hughey, L. F., Hein, A. M., Strandburg-Peshkin, A., Jensen, F. H. & Hughey, L. F. Challenges and solutions for studying collective animal behaviour in the wild. Phil. Trans. R. Soc. B 373, 20170005 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Calabrese, J. M. et al. Disentangling social interactions and environmental drivers in multi-individual wildlife tracking data. Phil. Trans. R. Soc. B 373, 20170007 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Jesmer, B. R. et al. Is ungulate migration culturally transmitted? Evidence of social learning from translocated animals. Science 361, 1023–1025 (2018).Article 
    CAS 
    PubMed 

    Google Scholar 
    Bousquet, C. A. H., Sumpter, D. J. T. & Manser, M. B. Moving calls: a vocal mechanism underlying quorum decisions in cohesive groups. Proc. R. Soc. B Biol. Sci. 278, 1482–1488 (2011).Article 

    Google Scholar 
    Dibnah, A. J. et al. Vocally mediated consensus decisions govern mass departures from jackdaw roosts. Curr. Biol. 32, R455–R456 (2022).Article 
    CAS 
    PubMed 

    Google Scholar 
    Robart, A. R., Zuñiga, H. X., Navarro, G. & Watts, H. E. Social environment influences termination of nomadic migration. Biol. Lett. 18, 20220006 (2022).Article 
    PubMed 

    Google Scholar 
    Dodson, S., Abrahms, B., Bograd, S. J., Fiechter, J. & Hazen, E. L. Disentangling the biotic and abiotic drivers of emergent migratory behavior using individual-based models. Ecol. Modell. 432, 109225 (2020).Article 

    Google Scholar 
    Kays, R., Crofoot, M. C., Jetz, W. & Wikelski, M. Terrestrial animal tracking as an eye on life and planet. Science 348, aaa2478 (2015).Article 
    PubMed 

    Google Scholar 
    Hussey, N. E. et al. Aquatic animal telemetry: a panoramic window into the underwater world. Science 348, 1255642 (2015).Article 
    PubMed 

    Google Scholar 
    Oestreich, W. K. et al. Acoustic signature reveals blue whale tune life history transitions to oceanographic conditions. Funct. Ecol. 36, 882–895 (2022).Article 
    CAS 

    Google Scholar 
    Chapman, J. W., Reynolds, D. R. & Smith, A. D. Vertical-looking radar: a new tool for monitoring high-altitude insect migration. BioScience 53, 503–511 (2003).Article 

    Google Scholar 
    Oestreich, W. K. et al. Animal-borne metrics enable acoustic detection of blue whale migration. Curr. Biol. 30, 4773–4779 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Fraser, K. C., Shave, A., de Greef, E., Siegrist, J. & Garroway, C. J. Individual variability in migration timing can explain long-term, population-level advances in a songbird. Front. Ecol. Evol. 7, 324 (2019).Article 

    Google Scholar 
    Byholm, P., Beal, M., Isaksson, N., Lötberg, U. & Åkesson, S. Paternal transmission of migration knowledge in a long-distance bird migrant. Nat. Commun. 13, 1566 (2022).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Schneider, S. S. & McNally, L. C. Waggle dance behavior associated with seasonal absconding in colonies of the African honey bee, Apis mellifera scutellata. Insectes Soc. 41, 115–127 (1994).Article 

    Google Scholar 
    Raveling, D. G. Preflight and flight behavior of Canada geese. Auk 86, 671–681 (1969).Article 

    Google Scholar 
    Tennessen, J. B., Parks, S. E. & Langkilde, T. Traffic noise causes physiological stress and impairs breeding migration behaviour in frogs. Conserv. Physiol. 2, cou032 (2014).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Lagarde, A., Lagarde, F. & Piersma, T. Vocal signalling by Eurasian spoonbills Platalea leucorodia in flocks before migratory departure. Ardea 109, 243–250 (2021).Article 

    Google Scholar 
    Rees, E. C. Conflict of choice within pairs of Bewick’s swans regarding their migratory movement to and from the wintering grounds. Anim. Behav. 35, 1685–1693 (1987).Article 

    Google Scholar 
    Mazeroll, A. I. & Montgomery, W. L. Daily migrations of a coral reef fish in the Red Sea (Gulf of Aqaba, Israel). Copiea 1998, 893–905 (1998).Article 

    Google Scholar 
    Méndez, V. et al. Paternal effects in the initiation of migratory behaviour in birds. Sci. Rep. 11, 2782 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Nelson, M. E. Development of migratory behavior in northern white-tailed deer. Can. J. Zool. 76, 426–432 (1998).Article 

    Google Scholar 
    Sweanor, P. Y. & Sandgren, F. Winter-range philopatry of seasonally migratory moose. J. Appl. Ecol. 26, 25–33 (1989).Article 

    Google Scholar 
    Rees, E. C. Consistency in the timing of migration for individual Bewick’s swans. Anim. Behav. 38, 384–393 (1989).Article 

    Google Scholar 
    Corten, A. A possible adaptation of herring feeding migrations to a change in timing of the Calanus finmarchicus season in the eastern North Sea. ICES J. Mar. Sci. 57, 1261–1270 (2000).Article 

    Google Scholar 
    Loonstra, A. J. et al. Individual black-tailed godwits do not stick to single routes: a hypothesis on how low population densities might decrease social conformity. Ardea 107, 251–261 (2020).Article 

    Google Scholar 
    Hake, M., Kjellén, N. & Alerstam, T. Age‐dependent migration strategy in honey buzzards Pernis apivorus tracked by satellite. Oikos 103, 385–396 (2003).Article 

    Google Scholar 
    Gupte, P. R., Koffijberg, K., Müskens, G. J. D. M., Wikelski, M. & Kölzsch, A. Family size dynamics in wintering geese. J. Ornithol. 160, 363–375 (2019).Article 

    Google Scholar 
    Gonçalves, M. I. C. et al. Movement patterns of humpback whales (Megaptera novaeangliae) reoccupying a Brazilian breeding ground. Biota Neotrop. 18, e20180567 (2018).Article 

    Google Scholar 
    Trudelle, L. et al. First insights on spatial and temporal distribution patterns of humpback whales in the breeding ground at Sainte Marie Channel, Madagascar. Afr. J. Mar. Sci. 40, 75–86 (2018).Article 

    Google Scholar 
    De La Gala-Hernández, S. R., Heckel, G. & Sumich, J. L. Comparative swimming effort of migrating gray whales (Eschrichtius robustus) and calf cost of transport along Costa Azul, Baja California, Mexico. Can. J. Zool. 86, 307–313 (2008).Article 

    Google Scholar 
    Sword, G. A. Local population density and the activation of movement in migratory band-forming Mormon crickets. Anim. Behav. 69, 437–444 (2005).Article 

    Google Scholar 
    Buhl, J. et al. From disorder to order in marching locusts. Science 312, 1402–1406 (2006).Article 
    CAS 
    PubMed 

    Google Scholar 
    Mysterud, A., Loe, L. E., Zimmermann, B., Bischof, R. & Meisingset, E. Partial migration in expanding red deer populations at northern latitudes—a role for density dependence? Oikos 120, 1817–1825 (2011).Article 

    Google Scholar 
    Bukreeva, O. M. & Lidzhi-garyaeva, G. V. Mass migration of social voles (Microtus socialis Pallas, 1773) in the Northwestern Caspian region. Arid Ecosyst. 8, 147–151 (2018).Article 

    Google Scholar 
    Eggeman, S. L., Hebblewhite, M., Bohm, H., Whittington, J. & Merrill, E. H. Behavioural flexibility in migratory behaviour in a long-lived large herbivore. J. Anim. Ecol. 85, 785–797 (2016).Article 
    PubMed 

    Google Scholar 
    Weithman, C. et al. Senescence and carryover effects of reproductive performance influence migration, condition, and breeding propensity in a small shorebird. Ecol. Evol. 7, 11044–11056 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Rappole, J. H. & Warner, D. W. Relationships between behavior, physiology and weather in avian transients at a migration stopover site. Oecologia 212, 193–212 (1976).Article 

    Google Scholar 
    Fauchald, P., Mauritzen, M. & Gjøsæter, H. Density‐dependent migratory waves in the marine pelagic ecosystem. Ecology 87, 2915–2924 (2006).Article 
    PubMed 

    Google Scholar 
    Makris, N. C. et al. Critical population density triggers rapid formation of vast oceanic fish shoals. Science 323, 1734–1737 (2009).Article 
    CAS 
    PubMed 

    Google Scholar 
    Tøttrup, A. P. & Thorup, K. Sex-differentiated migration patterns, protandry and phenology in North European songbird populations. J. Ornithol. 149, 161–167 (2008).Article 

    Google Scholar 
    Francis, C. M. & Cooke, C. F. Differential timing of spring migration in rose-breasted grosbeaks. J. Field Ornithol. 61, 404–412 (1990).
    Google Scholar 
    Corgos, A., Verísimo, P. & Freire, J. Timing and seasonality of the terminal molt and mating migration in the spider crab, Maja brachydactyla: evidence of alternative mating strategies. J. Shellfish Res. 25, 577–587 (2006).Article 

    Google Scholar 
    Gordo, O., Sanz, J. J. & Lobo, J. M. Spatial patterns of white stork (Ciconia ciconia) migratory phenology in the Iberian Peninsula. J. Ornithol. 148, 293–308 (2007).Article 

    Google Scholar 
    Sergio, F. et al. Individual improvements and selective mortality shape lifelong migratory performance. Nature 515, 410–413 (2014).Article 
    CAS 
    PubMed 

    Google Scholar 
    Manica, L. T., Graves, J. A., Podos, J. & Macedo, R. H. Hidden leks in a migratory songbird: mating advantages for earlier and more attractive males. Behav. Ecol. 31, 1180–1191 (2020).Article 

    Google Scholar 
    Cade, D. E. et al. Social exploitation of extensive, ephemeral, environmentally controlled prey patches by supergroups of rorqual whales. Anim. Behav. 182, 251–266 (2021).Article 

    Google Scholar 
    Urbanek, R. P., Fondow, L. E. A., Zimorski, S. E., Wellington, M. A. & Nipper, M. A. Winter release and management of reintroduced migratory whooping cranes Grus americana. Bird Conserv. Int. 20, 43–54 (2010).Article 

    Google Scholar 
    Németh, Z. & Moore, F. R. Information acquisition during migration: a social perspective. Auk 131, 186–194 (2014).Article 

    Google Scholar  More

  • in

    Dual ancestries and ecologies of the Late Glacial Palaeolithic in Britain

    Housley, R. A., Gamble, C. S., Street, M. & Pettitt, P. Radiocarbon evidence for the lateglacial human recolonisation of Northern Europe. Proc. Prehist. Soc. 63, 25–54 (1997).
    Google Scholar 
    Blockley, S. P. E., Donahue, R. E. & Pollard, A. M. Radiocarbon calibration and Late Glacial occupation in northwest Europe. Antiquity 74, 112–119 (2000).
    Google Scholar 
    Terberger, T., Barton, N. & Street, M. in Humans, Environment and Chronology of the Late Glacial of the North European Plain (eds Street, M. et al.) 189–207 (Romisch-Germanisches Zentralmuseum, 2009).Miller, R. Mapping the expansion of the Northwest Magdalenian. Quat. Int. 272–273, 209–230 (2012).
    Google Scholar 
    Riede, F. & Pedersen, J. B. Late Glacial human dispersals in Northern Europe and disequilibrium dynamics. Hum. Ecol. 46, 621–632 (2018).
    Google Scholar 
    Lazaridis, I. et al. Ancient human genomes suggest three ancestral populations for present-day Europeans. Nature 513, 409–413 (2014).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Jones, E. R. et al. Upper Palaeolithic genomes reveal deep roots of modern Eurasians. Nat. Commun. 6, 8912 (2015).CAS 
    PubMed 

    Google Scholar 
    Fu, Q. et al. The genetic history of Ice Age Europe. Nature 534, 200–205 (2016).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Villalba-Mouco, V. et al. Survival of Late Pleistocene hunter-gatherer ancestry in the Iberian Peninsula. Curr. Biol. 29, 1169–1177 (2019).Willis, K. J. & Whittaker, R. J. Perspectives: paleoecology. The refugial debate. Science 287, 1406–1407 (2000).CAS 
    PubMed 

    Google Scholar 
    Sommer, R. S. & Nadachowski, A. Glacial refugia of mammals in Europe: evidence from fossil records. Mamm. Rev. 36, 251–265 (2006).
    Google Scholar 
    Bennett, K. D. & Provan, J. What do we mean by ‘refugia’? Quat. Sci. Rev. 27, 2449–2455 (2008).
    Google Scholar 
    Terberger, T. & Street, M. Hiatus or continuity? New results for the question of pleniglacial settlement in Central Europe. Antiquity 76, 691–698 (2002).
    Google Scholar 
    Maier, A. in The Central European Magdalenian. Vertebrate Paleobiology and Paleoanthropology (ed. Maier, A.) 231–241 (Springer, 2015).Reade, H. et al. Radiocarbon chronology and environmental context of Last Glacial Maximum human occupation in Switzerland. Sci. Rep. 10, 4694 (2020).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Stevens, R. E., Hermoso-Buxán, X. L., Marín-Arroyo, A. B., González-Morales, M. R. & Straus, L. G. Investigation of Late Pleistocene and Early Holocene palaeoenvironmental change at El Mirón cave (Cantabria, Spain): insights from carbon and nitrogen isotope analyses of red deer. Palaeogeogr. Palaeoclimatol. Palaeoecol. 414, 46–60 (2014).
    Google Scholar 
    Clark, C. D., Hughes, A. L. C., Greenwood, S. L., Jordan, C. & Sejrup, H. P. Pattern and timing of retreat of the last British–Irish Ice Sheet. Quat. Sci. Rev. 44, 112–146 (2012).
    Google Scholar 
    Currant, A. P. & Jacobi, R. in The Ancient Human Occupation of Britain Vol. 14 (eds Ashton, N. et al.) 165–180 (Elsevier, 2011).Walker, M. J. C. et al. Devensian lateglacial environmental changes in Britain: a multi-proxy environmental record from Llanilid, South Wales, UK. Quat. Sci. Rev. 22, 475–520 (2003).
    Google Scholar 
    Hill, T. C. B. et al. Devensian late-glacial environmental change in the Gordano Valley, North Somerset, England: a rare archive for southwest Britain. J. Paleolimnol. 40, 431–444 (2008).
    Google Scholar 
    Jacobi, R. M. & Higham, T. F. G. The early Lateglacial re-colonization of Britain: new radiocarbon evidence from Gough’s Cave, southwest England. Quat. Sci. Rev. 28, 1895–1913 (2009).
    Google Scholar 
    Jacobi, R. & Higham, T. in The Ancient Human Occupation of Britain Vol. 14 (eds Ashton, N. M. et al.) 223–247 (Elsevier, 2011).Grimm, S. B. & Weber, M.-J. The chronological framework of the Hamburgian in the light of old and new 14C dates. Quartär. 55, 17–40 (2008).
    Google Scholar 
    Olalde, I. et al. The Beaker phenomenon and the genomic transformation of northwest Europe. Nature 555, 190–196 (2018).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Brace, S. et al. Ancient genomes indicate population replacement in Early Neolithic Britain. Nat. Ecol. Evol. 3, 765–771 (2019).PubMed 
    PubMed Central 

    Google Scholar 
    Jacobi, R. M. & Higham, T. F. G. The ‘Red Lady’ ages gracefully: new ultrafiltration AMS determinations from Paviland. J. Hum. Evol. 55, 898–907 (2008).CAS 
    PubMed 

    Google Scholar 
    Schulting, R. J. et al. A mid-upper Palaeolithic human humerus from Eel Point, South Wales, UK. J. Hum. Evol. 48, 493–505 (2005).PubMed 

    Google Scholar 
    Richards, M. P., Hedges, R. E. M., Jacobi, R., Current, A. & Stringer, C. FOCUS: Gough’s Cave and Sun Hole Cave human stable isotope values indicate a high animal protein diet in the British Upper Palaeolithic. J. Archaeol. Sci. 27, 1–3 (2000).
    Google Scholar 
    Proctor, C., Douka, K., Proctor, J. W. & Higham, T. The age and context of the KC4 Maxilla, Kent’s Cavern, UK. Eur. J. Archaeol. 20, 74–97 (2017).
    Google Scholar 
    Richards, M. P., Jacobi, R., Cook, J., Pettitt, P. B. & Stringer, C. B. Isotope evidence for the intensive use of marine foods by Late Upper Palaeolithic humans. J. Hum. Evol. 49, 390–394 (2005).CAS 
    PubMed 

    Google Scholar 
    Bello, S. M., Saladié, P., Cáceres, I., Rodríguez-Hidalgo, A. & Parfitt, S. A. Upper Palaeolithic ritualistic cannibalism at Gough’s Cave (Somerset, UK): the human remains from head to toe. J. Hum. Evol. 82, 170–189 (2015).PubMed 

    Google Scholar 
    Andrews, P. & Fernández-Jalvo, Y. Cannibalism in Britain: taphonomy of the Creswellian (Pleistocene) faunal and human remains from Gough’s Cave (Somerset, England). Bull. Nat. Hist. Mus. Geol. 58, 59–81 (2003).
    Google Scholar 
    Bello, S. M., Parfitt, S. A. & Stringer, C. B. Earliest directly-dated human skull-cups. PLoS ONE 6, e17026 (2011).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Currant, A. P., Jacobi, R. M. & Stringer, C. B. Excavations at Gough’s Cave, Somerset 1986–7. Antiquity 63, 131–136 (1989).
    Google Scholar 
    Davies, M. in Limestones and Caves of Wales (ed. Ford, T. D.) 92–101 (Cambridge Univ. Press, 1989).Dawkins, W. B. Memorandum on the remains from the cave at the Great Ormes Head. Proc. Liverp. Geol. Soc. 4, 156–159 (1880).
    Google Scholar 
    Sieveking, G. & de, G. The Kendrick’s Cave mandible. Br. Mus. Q. 35, 230–250 (1971).
    Google Scholar 
    Pettitt, P. B. Discovery, nature and preliminary thoughts about Britain’s first cave art.Capra 5,1–12 (2003).
    Google Scholar 
    Bello, S. M., Wallduck, R., Parfitt, S. A. & Stringer, C. B. An Upper Palaeolithic engraved human bone associated with ritualistic cannibalism. PLoS ONE 12, e0182127 (2017).PubMed 
    PubMed Central 

    Google Scholar 
    Bocherens, H. & Drucker, D. Isotope evidence for paleodiet of late Upper Paleolithic humans in Great Britain: a response to Richards et al. 2005. J. Hum. Evol. 51, 440–442 (2006).PubMed 

    Google Scholar 
    Fernandes, R., Millard, A. R., Brabec, M., Nadeau, M.-J. & Grootes, P. Food reconstruction using isotopic transferred signals (FRUITS): a Bayesian model for diet reconstruction. PLoS ONE 9, e87436 (2014).PubMed 
    PubMed Central 

    Google Scholar 
    Rasmussen, S. O. et al. A stratigraphic framework for abrupt climatic changes during the Last Glacial period based on three synchronized Greenland ice-core records: refining and extending the INTIMATE event stratigraphy. Quat. Sci. Rev. 106, 14–28 (2014).
    Google Scholar 
    Kloss-Brandstätter, A. et al. HaploGrep: a fast and reliable algorithm for automatic classification of mitochondrial DNA haplogroups. Hum. Mutat. 32, 25–32 (2011).PubMed 

    Google Scholar 
    Skoglund, P., Storå, J., Götherström, A. & Jakobsson, M. Accurate sex identification of ancient human remains using DNA shotgun sequencing. J. Archaeol. Sci. 40, 4477–4482 (2013).CAS 

    Google Scholar 
    Haak, W. et al. Massive migration from the steppe was a source for Indo-European languages in Europe. Nature 522, 207–211 (2015).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Fu, Q. et al. An early modern human from Romania with a recent Neanderthal ancestor. Nature 524, 216–219 (2015).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Patterson, N., Price, A. L. & Reich, D. Population structure and eigenanalysis. PLoS Genet. 2, e190 (2006).PubMed 
    PubMed Central 

    Google Scholar 
    Price, A. L. et al. Principal components analysis corrects for stratification in genome-wide association studies. Nat. Genet. 38, 904–909 (2006).CAS 
    PubMed 

    Google Scholar 
    Mallick, S. et al. The Simons Genome Diversity Project: 300 genomes from 142 diverse populations. Nature 538, 201–206 (2016).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Patterson, N. et al. Ancient admixture in human history. Genetics 192, 1065–1093 (2012).PubMed 
    PubMed Central 

    Google Scholar 
    Harney, É., Patterson, N., Reich, D. & Wakeley, J. Assessing the performance of qpAdm: a statistical tool for studying population admixture. Genetics 217, iyaa045 (2021).PubMed 
    PubMed Central 

    Google Scholar 
    Currant, A. & Jacobi, R. A formal mammalian biostratigraphy for the Late Pleistocene of Britain. Quat. Sci. Rev. 20, 1707–1716 (2001).
    Google Scholar 
    Pickard, C. & Bonsall, C. Post-glacial hunter-gatherer subsistence patterns in Britain: dietary reconstruction using FRUITS. Archaeol. Anthropol. Sci. 12, 142 (2020).
    Google Scholar 
    Stevens, R. E., Jacobi, R. M. & Higham, T. F. G. Reassessing the diet of Upper Palaeolithic humans from Gough’s Cave and Sun Hole, Cheddar Gorge, Somerset, UK. J. Archaeol. Sci. 37, 52–61 (2010).
    Google Scholar 
    Sala, N. & Conard, N. Taphonomic analysis of the hominin remains from Swabian Jura and their implications for the mortuary practices during the Upper Paleolithic. Quat. Sci. Rev. 150, 278–300 (2016).
    Google Scholar 
    Saladié, P. & Rodríguez-Hidalgo, A. Archaeological evidence for cannibalism in prehistoric Western Europe: from Homo antecessor to the Bronze Age. J. Archaeol. Method Theory 24, 1034–1071 (2017).
    Google Scholar 
    Cook, J. Ice Age Art: Arrival of the Modern Mind (British Museum Press, 2013).Gupta, S., Collier, J. S., Palmer-Felgate, A. & Potter, G. Catastrophic flooding origin of shelf valley systems in the English Channel. Nature 448, 342–345 (2007).CAS 
    PubMed 

    Google Scholar 
    Mills, W. in From the Atlantic to Beyond the Bug River. Finding and Defining the Federmesser-Gruppen/Azilian (eds Grimm, S. B. et al.) 1–24 (Propylaeum, 2020).Amkreutz, L. et al. What lies beneath … Late Glacial human occupation of the submerged North Sea landscape. Antiquity 92, 22–37 (2018).
    Google Scholar 
    Ward, I., Larcombe, P. & Lillie, M. The dating of Doggerland—post-glacial geochronology of the southern North Sea. Environ. Archaeol. 11, 207–218 (2006).
    Google Scholar 
    Brock, F., Higham, T., Ditchfield, P. & Ramsey, C. B. Current pretreatment methods for AMS radiocarbon dating at the Oxford Radiocarbon Accelerator Unit (ORAU). Radiocarbon 52, 103–112 (2010).CAS 

    Google Scholar 
    Dabney, J. et al. Complete mitochondrial genome sequence of a Middle Pleistocene cave bear reconstructed from ultrashort DNA fragments. Proc. Natl Acad. Sci. USA 110, 15758–15763 (2013).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Meyer, M. & Kircher, M. Illumina sequencing library preparation for highly multiplexed target capture and sequencing. Cold Spring Harb. Protoc. 2010, pdb.prot5448 (2010).Rohland, N., Harney, E., Mallick, S., Nordenfelt, S. & Reich, D. Partial uracil–DNA–glycosylase treatment for screening of ancient DNA. Philos. Trans. R. Soc. Lond. B 370, 20130624 (2015).
    Google Scholar 
    Kircher, M., Sawyer, S. & Meyer, M. Double indexing overcomes inaccuracies in multiplex sequencing on the Illumina platform. Nucleic Acids Res. 40, e3 (2012).CAS 
    PubMed 

    Google Scholar 
    Briggs, A. W. et al. Patterns of damage in genomic DNA sequences from a Neandertal. Proc. Natl Acad. Sci. USA 104, 14616–14621 (2007).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Skoglund, P. et al. Separating endogenous ancient DNA from modern day contamination in a Siberian Neandertal. Proc. Natl Acad. Sci. USA 111, 2229–2234 (2014).CAS 
    PubMed 
    PubMed Central 

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

    Google Scholar 
    Petr, M., Vernot, B. & Kelso, J. admixr—R package for reproducible analyses using ADMIXTOOLS. Bioinformatics 35, 3194–3195 (2019).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Busing, F. M., Meijer, E. & Van Der Leeden, R. Delete-m jackknife for unequal m. Stat. Comput. 9, 3–8 (1999).
    Google Scholar 
    Fu, Q. et al. Genome sequence of a 45,000-year-old modern human from western Siberia. Nature 514, 445–449 (2014).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Raghavan, M. et al. Upper Palaeolithic Siberian genome reveals dual ancestry of Native Americans. Nature 505, 87–91 (2014).PubMed 

    Google Scholar 
    Lazaridis, I. et al. Genomic insights into the origin of farming in the ancient Near East. Nature 536, 419–424 (2016).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Lipson, M. et al. Parallel palaeogenomic transects reveal complex genetic history of early European farmers. Nature 551, 368–372 (2017).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Gallego Llorente, M. et al. Ancient Ethiopian genome reveals extensive Eurasian admixture throughout the African continent. Science 350, 820–822 (2015).CAS 
    PubMed 

    Google Scholar 
    Villalba-Mouco, V. et al. Survival of Late Pleistocene hunter-gatherer ancestry in the Iberian Peninsula. Curr. Biol. 29, 1169–1177 (2019).CAS 
    PubMed 

    Google Scholar  More

  • in

    Deforestation slowed last year — but not enough to meet climate goals

    Deforested areas rim a highway running through the state of Amazonas, in Brazil.Credit: Michael Dantas/AFP/Getty

    Countries are failing to meet international targets to stop global forest loss and degradation by 2030, according to a report. It is the first to measure progress since world leaders set the targets last year at the 26th United Nations Climate Change Conference of the Parties (COP26) in Glasgow, UK. Preserving forests, which can store carbon and, in some cases, provide local cooling, is a crucial part of a larger strategy to curb global warming.
    Tropical forests have big climate benefits beyond carbon storage
    The analysis, called the Forest Declaration Assessment, shows that the rate of global deforestation slowed by 6.3% in 2021, compared with the baseline average for 2018–20. But this “modest” progress falls short of the annual 10% cut needed to end deforestation by 2030, says Erin Matson, a consultant at Climate Focus, an advisory company headquartered in Amsterdam, and author of the assessment, published on 24 October.“It’s a good start, but we are not on track,” Matson said at a press briefing, although she cautioned that the assessment looks at only one year’s worth of data. A clearer picture of deforestation trends will emerge in successive years, she added.The assessment, which was carried out by a number of civil-society and research groups, including the World Resources Institute, an environmental think tank in Washington DC, comes as nations gear up for the next big climate summit (COP27), to be held in November in Sharm El-Sheikh, Egypt. Scientists agree that in order to limit global warming to 1.5–2 °C above preindustrial levels — a threshold beyond which Earth’s climate will become profoundly disrupted — deforestation must end.Tropical forests are keyTo track deforestation over the past year, the groups analysed indicators such as changes in forest canopy, as measured by satellite data, and the forest landscape integrity index, which is a measure of the ecological health of forests. The slow progress they found is mainly attributable to a few tropical countries where deforestation is highest (see ‘Progress report’). Among them is Brazil — the world’s largest contributor to tree loss — which saw a 3% rise in the rate of deforestation in 2021, compared with the baseline years. Rates also rose in heavy deforesters Bolivia and the Democratic Republic of the Congo, by 6% and 3%, respectively, over the same period.

    Adapted from the 2022 Forest Declaration Assessment

    The loss of tropical forests, in particular, is worrisome because a growing body of research shows that besides sequestering carbon, these forests can physically cool nearby areas by creating clouds, humidifying the air and releasing certain cooling molecules. Keeping tropical forests standing provides a massive boost to global cooling that current policies ignore, says a report, “Not Just Carbon”, released alongside the Forest Declaration Assessment.A region made up of tropical countries in Asia is the only one on track to halt deforestation by 2030, according to the assessment (see ‘Movement towards goal’). The region cut the rate at which it lost humid, old-growth forests last year by 20% from the 2018–20 baseline, mostly thanks to large strides made by Indonesia — normally one of the world’s largest contributors to deforestation — where the loss of old-growth forests fell by 25% in 2021 compared with the previous year.

    Adapted from the 2022 Forest Declaration Assessment

    “The progress we see is driven by exceptional results in some countries,” Matson said.Efforts by the government and corporations in Indonesia to address the environmental harms of palm-oil production were key to progress, the assessment says. For example, as of 2020, more than 80% of palm-oil refiners had promised not to cut down or degrade any more forests. And in 2018, the Indonesian government imposed a moratorium on new palm-oil plantations. But the ban expired last year, raising concerns that progress might eventually be reversed.Finance laggingGlobal demand for commodities such as beef, fossil fuels and timber drive much of the forest loss that occurs today, as industry seeks to clear trees for new pastures and resource extraction. Matson said that many governments haven’t introduced reforms, such as protected-area regulations or fiscal incentives to encourage the private sector to safeguard forests, and that this is stalling progress.“Stronger mandatory action is needed,” she said.
    How much can forests fight climate change?
    In particular, nations are lagging behind in terms of fiscal support for forest protection and restoration. On the basis of previous assessments, the report estimates that forest conservation efforts require somewhere between US$45 billion and $460 billion per year if nations are to meet the 2030 goal. At present, commitments average less than 1% of what is needed per year, it concludes.Matson said that nations need to improve transparency on financing by setting interim milestones and publicly reporting progress. Michael Wolosin, a climate-solutions adviser at Conservation International, a non-profit environmental organization headquartered in Arlington, Virginia, would like to see donor countries recommit to their forest finance pledges at COP27 this year.However, Constance McDermott, an environmental-change researcher at the University of Oxford, UK, cautions against focusing too much on “estimates of forest cover change and dollars spent”. Social equity for Indigenous people and those in local communities should be part of discussions relating to deforestation, but is mostly missing, she says. These communities are the best forest stewards, and more effort is needed to support them by strengthening land rights and addressing land-use challenges that they identify, she says.Otherwise, McDermott warns that “global efforts to stop deforestation are more than likely to reinforce global, national and local inequalities”. More

  • in

    Reply To: Relative tree cover does not indicate a lagged Holocene forest response to monsoon rainfall

    replying to Y. Cheng et al. Nature Communications https://doi.org/10.1038/s41467-022-33958-7 (2022)We welcome the comment from Matters arising from Cheng Y et al.1, which provides us an opportunity for further clarification of some of our points2. The Comment raised interesting and important issues about our paper, that undoubtably could enhance our understanding for the Holocene vegetation evolution in the northern China and its relationship with East Asian Summer Monsoon (EASM). In particular, the results from Dali lake pose the questions on the timing of the peak of tree cover, that invokes the further investigation to understand this complex tree changes over Holocene period. However, these comments do not impact the key result in our original study2, that is the periodical asynchronous evolutions between EASM and northern China ecosystem under specific conditions.Main points of our paper are: First, we propose that the EASM and its rainfall over northern China mainly followed the variation of the summer insolation and peaked in the early Holocene, while the relative tree cover of temperate deciduous broadleaf tree peaked in the mid-Holocene; the delayed tree cover peak is caused by the winter warming, and peak soil moisture also in the mid-Holocene, which could be related to a hydrological impact from vegetation shift from grass to tree and the positive feedback between this vegetation shift and soil wetting.Second, this asynchronous evolution between the EASM rainfall, which peaks in the early Holocene, and the northern China ecosystem, which peaks in the mid-Holocene, is caused mainly by the opposing effect of residual ice sheet retreat on the decreasing summer insolation. The declining summer insolation does cause a substantial decrease of EASM rainfall from the early to mid-Holocene. However, 2/3 of this rainfall decrease is canceled by the rainfall increase forced by the retreat of residual Laurentide ice sheet, resulting in a weak decreasing trend of rainfall over this period.Third, under this background of weak rainfall changes, winter warming, induced by increased winter insolation and ice sheet retreat, raised the coldest temperature to above −17 °C, the threshold for the survival of temperate deciduous broadleaf tree3, and then favored an increase in tree, meanwhile induced a decrease in grass for reasons of its lower competitiveness than that of tree. This vegetation shift then supported the wetting of northern China through its hydrological effect2. The vegetation shift and soil wetting could reinforce each other. Furthermore, the dominant effect of winter warming on vegetation from the early to mid-Holocene is supported by our sensitive experiments with an off-line land-vegetation model.As stated in Cheng Y’s Comment1, the land cover in northern China includes forests, grass and bare land. In our interpretation, the process of “the vegetation feedback to climate” is mentioned as a possible feedback that enhances this asynchronous response, but is not critically involved in the mechanism. As such, whether the absolute or relative vegetation cover is not a major issue in our discussion. It’s sure that the reconstructed absolute tree cover, which based on pollen concentration, could enrich our understanding of the vegetation changes over the Holocene period in northern China. Indeed, the hydrological impact of bare land (evaporation) had been considered in our hydrological analysis of northern China soil moisture, and the results indicated its impact is important but not critical to the Holocene long-term changes of soil moisture over this region. The relative tree cover, the percentage of cool mixed tree (COMX4) in fossil pollen which is consistent with that of temperate deciduous broadleaf tree in simulation2, that we cited4 is a synthesis of 31 records, which represented the general evolution of vegetation over a large part of northern China, and its main result is consistent with records from other part of northern China such as the 6 ka peak in Gonghai Lake5. In spite of its low time resolution, the general trend over the millennium scale seems to us clear.It’s true that the −17 °C of the coldest month temperature is the survival threshold for the temperate deciduous broadleaved tree. While, the temperature threshold for C3 grass and C3 arctic grass are complex, its direct impact on the changes of grass proposed in our paper is somewhat not strict. However, considering the different competitiveness between tree and grass, increased temperate deciduous broadleaved tree, which derived by the winter warming, could induce a decrease in the grass from the early to mid-Holocene. Indeed, summer temperature, annual rainfall and fire incidents are all the important factors determining the Holocene changes of vegetation over northern China, but series of sensitivity experiment proposed the key impact of winter temperature on the vegetation shift and soil moisture evolution, which is consistent with the results of transient coupled climate simulation and geological records. This grass-to-tree shift for this period is evidenced in the pollen percentages and well simulated by the climate model shown in our paper2.Fire is an important factor for the long-term changes of vegetation cover over semi-arid regions, and its emergence and impact on vegetation are already incorporated into our model6, then, in turn, the simulation. Future works could assess the impact of fire on the long-term changes of semi-arid vegetation through the combination of reconstruction and process-based simulation of fire7.Focusing on the contrary views of Holocene EASM within proxy records, we proposed an asynchronous evolution of EASM rainfall and northern China ecosystem for the period of early to mid-Holocene. Our proposal is based on a state-of-the-art transient climate simulation, which reproduced the diverse evolution of EASM proxies reasonably well. The mechanisms proposed for this asynchronous evolution appear to us consistent with the current evidences available. There are, however, uncertainties in models and proxies. Meanwhile, the northern China is a broad region with large gradient in rainfall and ecosystem, that could induce the possible diverse evolutions in climate and ecosystem under Holocene climate change. Therefore, we believe further studies using other models and new proxies are important to further improve our understanding of this issue. More

  • in

    Old trees have much to teach us

    Elderflora: A Modern History of Ancient Trees Jared Farmer Basic (2022)About 45 million years ago, when the Arctic was ice-free, the world’s earliest known mummified trees flourished on what is now Axel Heiberg Island in Canada’s Qikiqtaaluk Region. In 1986, palaeobotanists identified the megaflora as members of Metasequoia occidentalis, an extinct redwood species. They had been buried in silt, then frozen, their wood preserved.The lead palaeontologist “celebrated his eureka by kindling a fire with 45-million-year-old twigs and boiling water for tea time,” writes historian Jared Farmer in Elderflora, his expansive global history of grand and venerable trees. Granted, these plants had been dead since the Eocene epoch. Nevertheless, as the author describes, the incident is part of a troubling pattern in which scientists rejoice at their discovery of the ‘oldest’ tree of their time — and then destroy it.In 1957, for example, Edmund Schulman at the University of Arizona in Tucson spent the summer seeking ancient bristlecone pines in California’s White Mountains. He found three more than 4,000 years old, and named them Alpha, Beta and Gamma. Then, in the interests of tree-ring science, he chose to “sacrifice” Alpha, taking snapshots as his nephew and a colleague sawed it down. When the University of Arizona issued a press release titled ‘UA Finds Oldest Living Thing’, Farmer writes, “they say nothing about the thing being dead”.Schulman’s aim was dendroclimatology — the reconstruction of climates using tree-ring data. That lofty motive cannot be ascribed to those who, in 1881, bored a tunnel into the 2,000-year-old Wawona tree in Yosemite National Park, allowing tourists to drive their cars through the 71.3-metre-high giant sequoia (Sequoiadendron giganteum), since toppled.Arboreal legendsAs Elderflora shows, big, old trees are objects of veneration and vandalism, appearing “in the oldest surviving mythologies and the earliest extant texts”. They were associated with gods and heroes, prophets and gurus: they had pivotal roles in the Mesopotamian Epic of Gilgamesh and in the Polynesian legend of Rātā, who fells a noble tree to carve a canoe. In more recent times, European settlers “dispossessed Indigenous peoples and cleared forests with abandon”. Research shows that, for 8,000 years after the glaciers of the last ice age retreated, forests in the Midwestern United States doubled in biomass (A. M. Raiho et al. Science 376, 1491–1495; 2022). Just 150 years of industrial logging and agriculture erased this carbon accumulation.
    It takes a wood to raise a tree: a memoir
    “Imperial conquests and industrial revolutions relied on timber,” Farmer writes. “Wood-stock long guns for capturing lands and peoples; naval vessels with mighty masts for transporting the enslaved and the harvests of their labor.” In New Zealand, European settlers decimated the majestic kauri trees, which can live for up to 2,000 years and that once covered 1.2 million hectares of land. The trees’ 50-metre-trunks became ships’ masts; their resin was made into varnish and linoleum.Like pines, firs, spruces, cedars, cypresses and redwoods, kauri (Agathis australis) is a gymnosperm. These flowerless plants with naked seeds tend to grow slower and live longer than angiosperms, flowering plants that bear fruit. About 25 plant species — most of them conifers — can live for more than a millennium without human assistance, surviving in restricted, vulnerable habitats.Farmer also offers a global survey of ancient trees that have been protected and exalted. They include olive trees of the Levant (Olea europaea); research published this year shows that these were domesticated about 7,000 years ago for their fruit and oil (D. Langgut and Y. Garfinkel Sci. Rep. 12, 7463; 2022). In Africa, the baobab (Adansonia sp.) is both the longest-lived tree and the largest, offering shade and shelter, foods, medicines and textiles. Enslaved Africans planted baobabs in the Caribbean; some survive still. Ginkgo biloba, a species that dates back 390,000 years, survived only in China, whence it was spread around the world in the past millennium. A grove of ginkgo trees survived the atomic bombing of Hiroshima in Japan in August 1945, pushing out new buds the following spring.The planet’s current tree cover, Farmer writes, includes 3 trillion large plants covering about 30% of all land. It is, in fact, expanding. But the new cover consists mostly of shelter belts (trees planted to protect crops or animals), temperate-zone timber crops and tropical plantations of eucalyptus and palm oil. A shrinking proportion of tree cover is made up of species-rich old-growth communities.Epic loss“What would humans and nonhumans stand to lose if these survivors all died prematurely? A world of things,” Farmer writes. “Old trees sustain forest communities” with their seeds and litter. Other plants grow on them, and animals live in them. Their roots share nutrients with other organisms via underground fungi. Groups of “Old Ones” are carbon sinks. Large-scale monocultures are shorter-lived and take less greenhouse gas out of circulation.But even bygone trees of the once-tropical Arctic might offer lessons for a warming world. Palaeobotanist Hope Jahren, in her 2016 memoir Lab Girl, describes how she spent three summers on Axel Heiberg Island, digging “through a hundred vertical feet of time”. Fir, cypress, larch, redwood, spruce, pine and hemlock trees populated this lush conifer forest, with an understory of angiosperms: maple, alder, birch, hickory, chestnut, beech, ash, holly, walnut, sweetgum, sycamore, oak, willow and elm. These plants thrived even through three months of winter darkness and three of constant summer light.“Here stood one of the great forests of all time,” Farmer writes. Today, as the Arctic warms nearly four times as fast as any other place on Earth, the genomes of species related to the trees of this mummified forest might be adaptable enough for the trees to flourish in a rewarmed planet, he says. Old trees have much to teach us: we would be wise to listen. More

  • in

    Induced pluripotent stem cells of endangered avian species

    Animal experimentsTeratoma formation experiments were performed at Iwate University. All surgical procedures and animal husbandry were performed in accordance with the international guidelines of the Animal Experiments of Iwate University and were approved by the university’s Animal Research Committee (approval number A201734).Chicken embryonic fibroblasts (Rhode Island Red) were obtained from a primary culture of chicken embryonic tissue provided by Prof. Atsushi Tajima, Tsukuba University. Chicken culture cells were obtained from chicken embryos, and the acquisition of these cells did not require approval. Mouse embryonic fibroblasts (CF-1 strain) were purchased from a manufacturer (CMPMEFCFL; DS Pharma Biomedical, Osaka, Japan). Approval was not required to obtain these cells.Somatic cells were obtained from wild animals (ex., Okinawa rail). The sampling details described below do not include the exact location of sampling to protect against poaching.Fibroblast cells from Okinawa rail and Japanese ptarmigan were obtained from dead animals, such as those killed by vehicles (Fig. 1A and Supplementary Fig. 1). Approval was not required to obtain these samples.Dead Okinawa rail were found on May 21, 2008, by the Okinawa Wildlife Federation, a nonprofit organization that focuses on the conservation of wild animals in the Okinawa area in the southwest region of Japan. The organization has permission from the Japanese Ministry of the Environment (MOE) to handle and perform first aid activities on endangered animals. The dead birds were transferred the following day to the National Institute for Environmental Studies (NIES). Primary cell culture was carried out from muscle tissue and skin of the dead birds (NIES ID: 715A).On July 8, 2004, tissues recovered from dead Japanese ptarmigan (e.g., skin and retina tissues) were also transferred to NIES from Gifu University Department of Veterinary Medicine. Primary cell culture from this tissue was performed (NIES ID: 22A).Somatic cells from Blakiston’s fish owl and Japanese golden eagle were obtained from emerging pinfeathers. Concerning the Blakiston’s fish owl, the MOE carries out bird banding, of wild birds with identification tags. The emerging pinfeathers we used had been accidentally release during banding. The banding had been performed by a veterinarian at the Institute for Raptor Biomedicine Japan (IRBJ) in the Hokkaido area on June 2, 2006. IRBJ is a private organization that primarily focuses on emergency medicine first aid and care for wild avians in Hokkaido region of Japan. IRBJ is contracted to MOE to handle and administer first aid for endangered animals. The MOE banding ring was 14C0242. Since banding was carried out with the permission of MOE for capturing wildlife, we did not require the approval to obtain these avian somatic cells. On July 8, 2006, Blakiston’s fish owl pinfeathers were transferred to from IRBJ to NIES, where primary cell culture was performed (NIES ID: 215A).Concerning the Japanese golden eagle, an emerging pinfeather accidentally fell off a bird during blood collection at the Yagiyama Zoo in Sendai, Japan on July 11, 2018. Dr. Yukiko Watanabe, an IRBJ veterinarian, collected the emerging pinfeather. The sample was shipped the following day to NIES where primary cell culture was performed (NIES ID: 5228).In addition to these birds, we obtained somatic cells emerging avian pinfeathers of Steller’s sea eagle, white-tail eagle, mountain hawk-eagle, northern goshawk, Taiga bean goose, and Latham’s snipe. These samples were provided by IRBJ.Concerning the Steller’s sea eagle, an injured individual was found in Hokkaido on July 11, 2006 (ID: 06-NE-SSE-1). The eagle was transferred to IRBJ. On December 4, 2006, IRBJ veterinarian Dr. Keisuke Saito collected fallen pinfeathers. Primary cell culture was performed at NIES on December 8, 2006 (NIES ID: 369A).Concerning the white-tailed eagle, an injured individual was found in Hokkaido, Japan, on July 12, 2007 (ID: 07-NE-WTE-4). The bird was transferred to IRBJ the same day for emergency treatment. On January 15, 2008, Dr. Saito collected fallen pinfeathers. Primary cell culture was performed on January 18, 2008 at NIES (NIES ID: 492A).Concerning the mountain hawk-eagle, an injured individual was found in the Hokkaido area on August 10, 2008 (ID: 08-Tokachi-HHE-2). The bird was transferred to IRBJ the same day. The bird was treated by an IRBJ veterinarian, but died on September 8, 2008. Emerging pinfeathers were collected from the dead bird by Dr. Saito. Primary cell culture was performed on September 11, 2008 at NIES (NIES ID: 847A).Concerning the Northern Goshawk, IRBJ accepted an injured bird for treatment on June 12, 2006. Following treatment and recovery, the bird was released into the wild in the Hokkaido area on August 1, 2006. During the treatment (July 4, 2006), Dr. Saito collected fallen pinfeathers. The primary cell culture was performed at NIES on July 6, 2006 (NIES ID: 222A).Concerning the Taiga bean geese, an injured individual was found in Hokkaido on September 15, 2016 (ID: 13B8005). The injured bird was transferred to IRBJ the same day for emergency treatment. On September 16, 2016, IRBJ veterinarian Dr. Yukiko Watanabe collected fallen emerging pinfeathers. Primary cell culture was performed on September 20, 2016 (NIES ID: 4420A).Finally, concerning the Latham’s snipe, fallen pinfeathers were collected during MOE approved bird banding performed on September 17, 2006, by Dr. Saito. Dr. Saito also collected fallen emerging pinfeathers (ID: 6A22598). The samples were transferred to NIES on September 20, 2006, for primary cell culture (NIES ID: 338A).All records are available at NIES.Cell culture and preservationOkinawa rail, Japanese ptarmigan, and Blakiston’s fish owl-derived fibroblasts were preserved in liquid nitrogen for 8–12 years (Fig. 1f). The preservation solution contained 90% fetal bovine serum (FBS) and 10% dimethyl sulfoxide. Cells were preserved at a cell density of 1 × 106–4 × 106 cell/mL. During the freezing period, the cells were maintained at minus The cells were frozen at a temperature of −135 °C. Japanese golden eagle fibroblasts were used without freezing.Avian-derived fibroblasts were cultured with Kuwana’s modified avian culture medium-1 (KAv-1), which is based on alpha-MEM containing 5% FBS and 5% chicken serum23. Mouse embryonic fibroblasts were cultured in Dulbecco’s modified Eagle’s medium (DMEM) containing 10% FBS and 1% antibiotic–antimycotic mixed stock solution (161–23181; Wako Pure Chemical Industries, Osaka, Japan). All avian and mouse cells were cultured at 37 °C under 5% CO2.Reprogramming vectorWe chemically synthesized an expression cassette that included seven reprogramming factors (MyoD-derived transactivation domain-linked Oct3/4, Sox2, Klf4, c-Myc, Klf2, Lin28, and Nanog; all genes derived from mice). The self-cleaving 2A peptide was inserted at the junction of the coding region (Fig. 1g). We transferred the complementary DNA (cDNA) insert from the shuttle vector to the PiggyBac transposon vector containing green fluorescent protein (PB-CAG-GFP). Although the original transposon vector drive the expression of cDNA with the elongation factor-1 (EF1) promoter (PJ547-17; DNA 2.0, Menlo Park, CA, USA), we replaced the EF1 promoter to CAG promoter in our previous study22,24. The reprogramming vector was designated PB-TAD-7F (Fig. 1g).In addition to the PB-TAD-7F reprogramming vector, we used the PB-DDR-8F reprogramming vector to establish Japanese golden eagle iPSCs. The complete coding sequence of DDR-8F (DDR-Oct3/4, Sox2, Klf4, c-Myc, Klf2, Nanog, Lin28, and Yap) was chemically synthesized. The expression cassettes containing the eight reprogramming factors were excised from the shuttle vector using restriction enzymes. The cDNA fragments were transferred to the PB-CAG-GFP PiggyBac transposon-based vector22,24. Detailed information regarding the PB-DDR-8F reprogramming vectors is shown in Fig. 10a.Establishment of iPSCsWe transfected PB-R6F or PB-TAD-7F reprogramming vectors into mouse, chicken, Okinawa rail, Japanese ptarmigan, and Blakiston’s fish owl-derived fibroblasts using Lipofectamine 2000 transfection reagent (Thermo Fisher Scientific, Waltham, MA, USA). After hygromycin selection (Wako Pure Chemical Industries), the cells were reseeded onto a mouse embryonic fibroblast (MEF) feeder layer. On days 14–32, we picked primary iPSC-like colonies and seeded them on new MEF feeder cell plates. The detailed protocol is shown in Fig. 1h.To establish Japanese golden eagle-derived iPSCs, we transduced PB-TAD-7F or PB-DDR-8F reprogramming vectors into Japanese golden eagle pinfeather-derived somatic cells. Transfection was performed using Lipofectamine 2000 transduction reagent (11668019; Thermo Fisher Scientific) according to the manufacturer’s instructions. After hygromycin selection (Wako Pure Chemical Industries), cells were seeded onto feeder culture plates. The golden eagle iPSCs were cultured in KAv-1-based medium5.The medium used to establish avian iPSCs was supplemented with 1000 × human Leukemia Inhibitory Factor (LIF) (125–05603; Wako Pure Chemical Industries), 4.0 ng/ml basic FGF (064–04541; Wako Pure Chemical Industries), 0.75 μM CHIR99021 glycogen synthase kinase-3 inhibitor (034–23103; Wako Pure Chemical Industries), 0.25 μM PD0325901 mitogen-activated protein kinase inhibitor (163–24001; Wako Pure Chemical Industries). In addition to those supplements, 0.25 μM thiazovivin (202–18011; Wako Pure Chemical Industries) was added in the media used to generate Okinawa rail, Japanese ptarmigan, Blakiston’s fish owl, and chicken iPSCs. In the medium used to generate mouse iPSCs, we added 1000 × LIF, 0.75 μM CHIR99021, and 0.25 μM PD0325901.iPSC culture conditionsTwo types of cell culture media were used: KAv-1 for avian iPSCs and DMEM for mouse iPSCs. The composition of KAv-1 for avian iPSCs was as follows: alpha-MEM containing 5% FBS and 5% chicken serum 1% antibiotic–antimycotic mixed solution, 1% nonessential amino acids (Wako Pure Chemical Industries), and 2 mM glutamic acid was added (Nacalai Tesque, Kyoto, Japan). The composition of DMEM for mouse was follows: DMEM supplemented with 15% SSR, 0.22 mM 2-mercaptoethanol (21438–82, Nacalai Tesque), 1% antibiotic–antimycotic mixed solution, 1% nonessential amino acids5,22. As a supplement to the iPSC medium, we used 1000 × human LIF (125–05603; Wako Pure Chemical Industries), 4.0 ng/ml basic FGF (064–04541; Wako Pure Chemical Industries), 0.75 μM CHIR99021 (034–23103; Wako Pure Chemical Industries), 0.25 μM PD0325901 (163–24001; Wako Pure Chemical Industries) for the media used to culture Okinawa rail, Japanese ptarmigan, Blakiston’s fish owl, Japanese golden eagle, and chicken-derived iPSCs. The supplements for media used to culture Okinawa rail and Japanese ptarmigan-derived iPSCs included 2.5 μM Gö6983 (074–06443, Wako Pure Chemical Industries). To analyze the cellular characteristics, we focused on the Janus kinase (JAK), FGF, ROCK, and glycolytic pathways, since the dependency of these pathways can indicate differences in cellular characteristics. We used 1–10 μM JAK inhibitor I (4200099; MERCK, Darmstadt, Germany), 0.5–4 μM of PD173074, which inhibits FGF receptor (FGFR) inhibitor (160–26831; Wako Pure Chemical Industries), 10 μM of Y27632, which inhibits ROCK (036–24023; Wako Pure Chemical Industries), and 2 or 4 mM 2-deoxyglucose (2DG, D0051; Tokyo Chemical Industry, Tokyo, Japan).AP and immunological staining of fibroblasts and iPSCsA red-color AP staining kit (AP100 R-1; System Bioscience, Palo Alto, CA, USA) was used to detect AP activity of iPSCs. iPSCs were stained for SSEA-1, SSEA-3, and SSEA-4 antibodies (Supplementary Table 2). To stain the iPSCs with the SSEA antibodies, the cells were fixed in 4% paraformaldehyde in phosphate buffered saline (PBS) for 3 min. Cells were permeabilized by 0.5% Triton X-100 (35501-15; Nacalai Tesque, Kyoto, Japan) for 60 min. After three washes with PBS, the iPSCs were blocked with 1% bovine serum albumin (BSA, 01863-06; Nacalai Tesque) for 45 min. iPSCs were incubated with a primary antibody overnight and then exposed to the corresponding fluorescent-labeled secondary antibodies for 60 min. Counterstaining was performed with a 4′,6-diamidino-2-phenylindole (DAPI) solution (Cellstain-DAPI solution, DOJINDO, Kumamoto, Japan).Japanese golden eagle and chicken-derived fibroblasts were seeded in 12-well cell culture plates for immunological staining. After 48 h of incubation, F-actin staining was performed using Alexa Fluor 568 phalloidin (A12380; Thermo Fisher Scientific) according to the manufacturer’s protocol. Double staining was performed with an anti-vimentin antibody (MA5-11883; Thermo Fisher Scientific) and Alexa Fluor 488-labeled secondary antibody (A-11001; Thermo Fisher Scientific) (Supplementary Table 2). The samples were counterstained with Cellstain-DAPI solution (DOJINDO) as described above.Detection of reprogramming vectors and internal control genes from iPSCsDNA was isolated using the EZ1 DNA Tissue Kit (953034; QIAGEN, Hilden, Germany). PCR was performed with 100 ng of template DNA. Primer sequences are listed in Supplementary Tables 3 and 4. We performed PCR assays using KOD FX Neo (KFX-201; TOYOBO, Osaka, Japan). PCR was conducted by predenaturation at 94 °C for 2 min, denaturation at 98 °C for 10 s, and extension at 68 °C for 30 s, with 40 cycles of denaturation and extension. PCR products were analyzed by electrophoresis on 2.0% agarose/Tris-acetate–ethylenediaminetetraacetic acid (EDTA) gels.Sequential passagingMouse, Okinawa rail, and Japanese ptarmigan-derived primary cells and iPSCs were seeded in six-well plates with feeder cells for analysis. When cell growth became confluent, all cells and the number of cells per dish was enumerated using a Countess cell counter (Thermo Fisher Scientific). The harvested and seeded cell numbers were used to calculate the PD time as an indicator of the speed of cell growth, using the formula PD = log2 (A/B), where A is the number of harvested cells at the end of each passage, and B is the number of seeded cells at the start25.Detection of mRNA expressionTotal RNA was isolated from iPSCs using an EZ1 RNA Tissue Mini Kit (959034; QIAGEN). cDNA was synthesized from total RNA using the PrimeScript reverse transcription (RT) reagent kit (Perfect Real Time, RR047A; TaKaRa Bio, Ohtsu, Japan). Real-time PCR was performed in a 12.5 μl volume containing 2 × KOD SYBR qPCR Mix (QKD-201; Toyobo), 10 ng of cDNA solution, and 0.3 μM of each primer. The primer sequences are listed in Supplementary Tables 5–10. The reaction was performed in duplicate. The cycling program was as follows: 98 °C for 120 s (initial denaturation), 98 °C for 10 s (denaturation), 58 °C for 10 s (annealing), and 68 °C for 32 s (extension) for 40 cycles. We normalized the expression levels of the target genes to that of glyceraldehyde-3-phosphate dehydrogenase (GAPDH).Mitochondria stainingMitochondria were stained by incubation with 50 nM MitoTracker Orange (M7510; Thermo Fisher Scientific) or 20 nM tetramethyl rhodamine ethyl ester perchlorate (TMRE, T669; Thermo Fisher Scientific) for 10 min. After staining, the solution was removed, and fresh medium was added for observation.EB formation and in vitro differentiationIn vitro differentiation of Okinawa rail, Japanese ptarmigan, Blakiston’s fish owl, and Japanese golden eagle iPSCs was performed. To generate EBs, iPSCs were seeded in low-binding dishes in KAv-1 medium. After 7–14 days, floating EBs were selected and seeded in 0.1% gelatin-coated 6-well plates with KAv-1 medium. To induce differentiation into neural cells, the floating EBs were cultured in 0.1% gelatin-coated plates containing KAv-1 supplemented with 10 μM ATRA and 4.0 ng/ml FGF for 7 days.Cells were immunochemically stained after in vitro differentiation using antibody to TUJ1, alpha-smooth muscle, or Gata4 (Supplementary Table 2). Differentiated cells were stained based on the immunological staining procedure of iPSCs described above.Teratoma formation and tissue sectioningThe Animal Committee of Iwate University approved the experimental protocol for teratoma formation (approval numbers A201734, A201737). For teratoma formation, 1 × 106 iPSCs were injected into the testes of SCID mice (C.B-17/Icr-scid/scidJcl; CLEA Japan, Tokyo, Japan). After 4–34 weeks post-injection, tumor tissues were excised from the mice. Each tumor tissue was fixed with 10% formaldehyde in PBS. Fixed tissue sections were stained with hematoxylin-eosin (HE) and observed by microscopy.Immunological staining was performed in addition to HE staining. For immunological staining, antibody to TUJ1, alpha-smooth muscle, or Gata4 was used (Supplementary Table 2). The paraffin block of each teratoma was sliced to produce a section 5 μm thick. After deparaffinization, the antigen was activated with citric acid buffer (SignalStain Citrate Unmasking Solution (10×), 14746; Cell Signaling Technology, Beverly, MA, USA) by microwaving for 10 min. To block endogenous peroxidase, tissue sections were incubated with 3% hydrogen peroxide (081–04215; Wako Pure Chemical). After washing with purified water, the tissue sections were incubated with 5% goat serum (555–76251; Wako Pure Chemical) in PBS. Next, the section were incubated in a solution containing a 1:100 dilution of primary antibody overnight at 4 °C. After washing with PBS, the tissue sections were incubated with horseradish peroxidase (HRP) conjugated secondary antibody (anti-IgG (H+L chain), mouse, pAb-HRP, code no. 330; MBL Co., Ltd., Nagoya, Japan) or anti-IgG (H+L chain, rabbit, pAb-HRP, code no. 458; MBL) for 1 h (Supplementary Table 2). After washing with PBS, the tissue sections were incubated with 3,3′-diaminobenzidine substrate solution (Histostar, code no. 8469; MBL) for 5–20 min. After washing with purified water, tissue sections were counterstained with hematoxylin for 1–2 min.DNA component analysisCultured cells fixed with 70% ethanol at least 4 h under −20 °C condition. The fixed cells stained with the Muse Cell Cycle Assay Kit (Merck Millipore Corporation, Darmstadt, Germany). The stained cells analyzed with Muse Cell Analyzer (Merck Millipore Corporation) were used for DNA content analysis.Karyotype analysisOur iPSCs were treated with 0.02 mg/ml colcemid. Those iPSCs exposed to a hypotonic solution and fixed with Carnoy’s fluid. We counted the chromosomal number in 50 cells and performed a G-banding analysis in 20 cells22.Production of interspecific chimeras and their immunological stainingTo evaluate whether iPSCs derived from Japanese ptarmigan could contribute to the generation of interspecific chimeras in chick embryos, iPSCs were stained with 10 μM CellTracker Green CMFDA (5-chloromethylfluorescein diacetate, C7025; Thermo Fisher Scientific) for 30 min. Eggs of white leghorn chicken were purchased from a local farm (Goto-furanjyo, Gifu, Japan). We injected the labeled Japanese ptarmigan iPSCs into stage X chick blastoderms and cultured the embryos26. To confirm the contribution of chimera, fluorescence was observed after 72 h. To analyze the tissue-level contribution of chimera, embryos on day 5. The embryos were embedded in optimal cutting temperature compound (Sakura Finetek Japan, Tokyo, Japan), frozen in liquid nitrogen, and stored at −80 °C until use. Cryosections 20 μm in thickness were prepared using a cryostat, air-dried for 30 min at room temperature, and fixed with 4% paraformaldehyde for 2 min at room temperature. After washing three times with PBS, sections were incubated with PBS containing 5% FBS for 1 h. After blocking with FBS, the sections were incubated with an anti-hygromicin resistance gene antibody (anti-HPT2; Supplementary Table 2) overnight. After washing three times with PBS, the sections were incubated with secondary antibody (goat anti-mouse IgG, Alexa Fluor 568; Supplementary Table 2) and Cellstain- DAPI solution (DOJINDO) for 1 h.Detection of contribution of chimera from genomeWe injected Japanese ptarmigan iPSCs (without CellTracker Green CMFDA label) into a stage X chicken blastoderms. On day 5, the entire chicken embryos were collected. The genome of each embryo was collected using NucleoSpin Tissue (U0952S; MACHEREY-NAGEL, Düren, Germany). After collecting the chimeric genome, we detected the reprogramming vector cassette using genomic PCR analysis using 50 ng of template genome. To extend the target sequence, we used the KOD FX Neo (KFX-201; TOYOBO). Primer information is provided in Supplementary Table 11. This analysis was performed according to the manufacturer’s protocol. The cycling program comprised 45 cycles of 94 °C for 120 s (initial denaturation), 98 °C for 10 s (denaturation), and 68 °C for 50 s (annealing and extension). After PCR, 2% agarose gel electrophoresis was performed. Gels were stained with GelGreen (517–53333; Biotium, Inc., Fremont, CA, USA).Real-time PCR was also performed to detect the contribution of chimera. The fluorescence probe and primers designed to detect chimeric contributions are summarized in Supplementary Table 12. The template was a 30 ng genome. The analysis was performed using 1 × THUNDERBIRD Probe qPCR Mix (QPS-101; TOYOBO), 0.3 μM of each primer, 0.2 μM of probe, and 1 × Rox. Fifty cycle of 95 °C for 60 s (initial denaturation), 95 °C for 15 s (denaturation), and 60 °C for 60 s (annealing and extension) were used. The expression levels of the target genes were normalized to that of chicken Tsc-2.RNA preparation and sequencing for RNA-seq analysisTotal RNA from iPSCs, fibroblasts, and chicken embryo stage X was collected using NucleoSpin Tissue (740952.50; MACHEREY-NAGEL). Triplicate samples of all iPSCs, fibroblasts, and chicken embryo stage X were prepared. To prepare the library, we used the TruSeq Stranded mRNA LT Sample Prep Kit (RS-122-2101; Illumina, San Diego, CA, USA). The quality of the library was evaluated using the Qubit DNA Assay (Thermo Fisher Scientific) on a TapeStation with a D1000 screen tape (Agilent Technologies, Santa Clara, CA, USA). The cDNA samples were used for the sequencing reaction on an Illumina HiSeq X sequencing machine, resulting in more than 40 M reads with 150 bp ends for each sample, except chicken fibroblast No. 3, which displayed more than 40 M reads with 75 bp ends. To analyze the RNA-seq data, we used the CLC Genomic Workbench (CLC Bio, Aarhus, Denmark). In the trim read step, low-quality sequence with the quality score of the CLC workbench, 5′ end, 3′ end, and short sequences (shorter than 15 sequences) were removed. The trimmed sequence data were mapped onto the chicken reference genome. Gene expression data were obtained in this step. PCA was performed and a heat map created with CLC Genomic Workbench using gene expression data. In this step, normalization was automatically performed using TMM methods. To compare chicken cells, RNA-seq data from SRA (SRP115012 (GEO: GSE102353) and SRP087639 (GSE86592) were used. The RNA-seq data has been submitted to the DNA DataBank of Japan under accession number DRA013522 (Submission), PRJDB13093(BioProject), SAMD00444261–SAMD00444287 (BioSample).Statistics and reproducibilityNonparametric multiple comparison analysis used the Steal–Dwass test (Figs. 2e, 3 [Okinawa rail, Japanese ptarmigan, and Blakiston’s fish owl], 4d, 4f, 5b, 5d, 5f, 5h, 10i). For nonparametric independent two-group analysis, we used the Mann–Whitney U test (Fig. 3, for mouse and chicken, and 4b). Statistically significant differences are indicated by *(p  More

  • in

    Protection status, human disturbance, snow cover and trapping drive density of a declining wolverine population in the Canadian Rocky Mountains

    Le Saout, S. et al. Protected areas and effective biodiversity conservation. Science 342, 803–805 (2013).Article 
    ADS 
    PubMed 

    Google Scholar 
    Chape, S., Harrison, J., Spalding, M. & Lysenko, I. Measuring the extent and effectiveness of protected areas as an indicator for meeting global biodiversity targets. Philos. Trans. R. Soc. B Biol. Sci. 360, 443–455 (2005).Article 
    CAS 

    Google Scholar 
    Hansen, A. J. & DeFries, R. Ecological mechanisms linking protected areas to surrounding lands. Ecol. Appl. 17, 974–988 (2007).Article 
    PubMed 

    Google Scholar 
    Balme, G. A., Slotow, R. & Hunter, L. T. B. Edge effects and the impact of non-protected areas in carnivore conservation: Leopards in the Phinda-Mkhuze Complex, South Africa. Anim. Conserv. 13, 315–323 (2010).Article 

    Google Scholar 
    Woodroffe, R. & Ginsberg, J. R. Edge effects and the extinction of populations inside protected areas. Science 280, 2126–2128 (1998).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Watson, J. E. M., Dudley, N., Segan, D. B. & Hockings, M. The performance and potential of protected areas. Nature 515, 67–73 (2014).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Jones, K. R. et al. One-third of global protected land is under intense human pressure. Science 360, 788–791 (2018).Article 
    CAS 
    PubMed 

    Google Scholar 
    Balmford, A. et al. Walk on the wild side: Estimating the global magnitude of visits to protected areas. PLoS Biol. 13, 1–6 (2015).Article 

    Google Scholar 
    Larson, C. L., Reed, S. E., Merenlender, A. M. & Crooks, K. R. Effects of recreation on animals revealed as widespread through a global systematic review. PLoS ONE 11, 1–21 (2016).Article 

    Google Scholar 
    Tablado, Z. & Jenni, L. Determinants of uncertainty in wildlife responses to human disturbance. Biol. Rev. 92, 216–233 (2017).Article 
    PubMed 

    Google Scholar 
    Timko, J. A. & Innes, J. L. Evaluating ecological integrity in national parks: Case studies from Canada and South Africa. Biol. Conserv. 142, 676–688 (2009).Article 

    Google Scholar 
    Nagendra, H. et al. Remote sensing for conservation monitoring: Assessing protected areas, habitat extent, habitat condition, species diversity, and threats. Ecol. Indic. 33, 45–59 (2013).Article 

    Google Scholar 
    Frid, A. & Dill, L. M. Human-caused disturbance stimuli as a form of predation risk. Conserv. Ecol. 6, 11 (2002).
    Google Scholar 
    Lima, S. L. & Dill, L. M. Behavioral decisions made under the risk of predation: A review and prospectus. Can. J. Zool. 68, 619–640 (1990).Article 

    Google Scholar 
    Creel, S. & Christianson, D. Relationships between direct predation and risk effects. Trends Ecol. Evol. 23, 194–201 (2008).Article 
    PubMed 

    Google Scholar 
    Williams, B. K., Nichols, J. D. & Conroy, M. J. Analysis and Management of Animal Populations (Academic Press, 2002).
    Google Scholar 
    Royle, J. A., Chandler, R. B., Sollmann, R. & Gardner, B. Spatial Capture-Recapture (Academic Press, 2014).
    Google Scholar 
    Steenweg, R., Hebblewhite, M., Whittington, J. & McKelvey, K. Species-specific differences in detection and occupancy probabilities help drive ability to detect trends in occupancy. Ecosphere 10, e02639 (2019).Article 

    Google Scholar 
    Chen, C. et al. Global camera trap synthesis highlights the importance of protected areas in maintaining mammal diversity. Conserv. Lett. 15, 1–14 (2022).Article 
    CAS 

    Google Scholar 
    Besbeas, P., Freeman, S., Morgan, B. & Catchpole, E. Integrating mark-recapture-recovery and census data to estimate animal abundance and demographic parameters. Biometrics 58, 540–547 (2002).Article 
    MathSciNet 
    CAS 
    PubMed 
    MATH 

    Google Scholar 
    Sun, C. C., Royle, J. A. & Fuller, A. K. Incorporating citizen science data in spatially explicit integrated population models. Ecology 100, 1–12 (2019).Article 

    Google Scholar 
    Doran-Myers, D. et al. Density estimates for Canada lynx vary among estimation methods. Ecosphere 12, 3774 (2021).Article 

    Google Scholar 
    Ripple, W. J. et al. Status and ecological effects of the world’s largest carnivores. Science 343, 1241484 (2014).Article 
    PubMed 

    Google Scholar 
    Frey, S., Fisher, J. T., Burton, A. C. & Volpe, J. P. Investigating animal activity patterns and temporal niche partitioning using camera-trap data: Challenges and opportunities. Remote Sens. Ecol. Conserv. 3, 123–132 (2017).Article 

    Google Scholar 
    Weaver, J. L., Paquet, P. C. & Ruggiero, L. F. Resilience and conservation of large carnivores in the Rocky Mountains. Conserv. Biol. 10, 964–976 (1996).Article 

    Google Scholar 
    Fisher, J. T. et al. Wolverines (Gulo gulo) in a changing landscape and warming climate: A decadal synthesis of global conservation ecology research. Glob. Ecol. Conserv. 34, e02019 (2022).Article 

    Google Scholar 
    Persson, J., Ericsson, G. & Segerström, P. Human caused mortality in the endangered Scandinavian wolverine population. Biol. Conserv. 142, 325–331 (2009).Article 

    Google Scholar 
    Mowat, G. et al. The sustainability of wolverine trapping mortality in Southern Canada. J. Wildl. Manag. 84, 213–226 (2020).Article 

    Google Scholar 
    Bowman, J., Ray, J. C., Magoun, A. J., Johnson, D. S. & Dawson, F. N. Roads, logging, and the large-mammal community of an eastern Canadian boreal forest. Can. J. Zool. 88, 454–467 (2010).Article 

    Google Scholar 
    Heinemeyer, K. et al. Wolverines in winter: Indirect habitat loss and functional responses to backcountry recreation. Ecosphere 10, 2611 (2019).Article 

    Google Scholar 
    Fisher, J. T. et al. Wolverines (Gulo gulo Luscus) on the Rocky Mountain slopes: Natural heterogeneity and landscape alteration as predictors of distribution. Can. J. Zool. 91, 706–716 (2013).Article 

    Google Scholar 
    Magoun, A. J. et al. Integrating motion-detection cameras and hair snags for wolverine identification. J. Wildl. Manag. 75, 731–739 (2011).Article 

    Google Scholar 
    Bischof, R. et al. Estimating and forecasting spatial population dynamics of apex predators using transnational genetic monitoring. Proc. Natl. Acad. Sci. U.S.A. 117, 30531–30538 (2020).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Aronsson, M. & Persson, J. Mismatch between goals and the scale of actions constrains adaptive carnivore management: The case of the wolverine in Sweden. Anim. Conserv. 20, 261–269 (2017).Article 

    Google Scholar 
    Newmark, W. D. Extinction of mammal populations in Western North American National Parks. Conserv. Biol. 9, 512–526 (1995).Article 

    Google Scholar 
    Barrueto, M., Sawaya, M. A. & Clevenger, A. P. Low wolverine (Gulo gulo) density in a national park complex of the Canadian Rocky Mountains. Can. J. Zool. 98, 287–298 (2020).Article 

    Google Scholar 
    Heim, N., Fisher, J. T., Clevenger, A., Paczkowski, J. & Volpe, J. Cumulative effects of climate and landscape change drive spatial distribution of Rocky Mountain wolverine (Gulo gulo L.). Ecol. Evol. 7, 8903–8914 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Steenweg, R. et al. Camera-based occupancy monitoring at large scales: Power to detect trends in grizzly bears across the Canadian Rockies. Biol. Conserv. 201, 192–200 (2016).Article 

    Google Scholar 
    Tourani, M., Dupont, P., Nawaz, M. A. & Bischof, R. Multiple observation processes in spatial capture–recapture models: How much do we gain? Ecology 101, 1–8 (2020).Article 

    Google Scholar 
    Kukka, P. M., Jung, T. S. & Schmiegelow, F. K. A. Spatiotemporal patterns of wolverine (Gulo gulo) harvest: The potential role of refugia in a quota-free system. Eur. J. Wildl. Res. 68, 1566 (2022).Article 

    Google Scholar 
    Krebs, J. et al. Synthesis of survival rates and causes of mortality in North American wolverines. J. Wildl. Manag. 68, 493–502 (2004).Article 

    Google Scholar 
    Stewart, F. E. C. et al. Wolverine behavior varies spatially with anthropogenic footprint: Implications for conservation and inferences about declines. Ecol. Evol. 6, 1493–1503 (2016).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Sawaya, M. A., Clevenger, A. P. & Schwartz, M. K. Demographic fragmentation of a protected wolverine population bisected by a major transportation corridor. Biol. Conserv. 236, 616–625 (2019).Article 

    Google Scholar 
    Gooliaff, T. The Sustainable Annual Take of Canada lynx in the Okanagan Region of British Columbia (2021).Clevenger, A. P. Mitigating highways for a ghost: Data collection challenges and implications for managing wolverines and transportation corridors. Northwest Sci. 87, 257–264 (2013).Article 

    Google Scholar 
    Kindsvater, H. K. et al. Overcoming the data crisis in biodiversity conservation. Trends Ecol. Evol. 33, 676–688 (2018).Article 
    PubMed 

    Google Scholar 
    Gervasi, V. et al. Compensatory immigration counteracts contrasting conservation strategies of wolverines (Gulo gulo) within Scandinavia. Biol. Conserv. 191, 632–639 (2015).Article 

    Google Scholar 
    Rich, L. N. et al. Assessing global patterns in mammalian carnivore occupancy and richness by integrating local camera trap surveys. Glob. Ecol. Biogeogr. 26, 918–929 (2017).Article 

    Google Scholar 
    Decesare, N. J. et al. The role of translocation in recovery of Woodland Caribou populations. Conserv. Biol. 25, 365–373 (2010).PubMed 

    Google Scholar 
    Morris, W. & Doak, D. Quantitative Conservation Biology (Sinauer Associates, 2002).
    Google Scholar 
    Squires, J. R. et al. Combining resource selection and movement behavior to predict corridors for Canada lynx at their southern range periphery. Biol. Conserv. 157, 187–195 (2013).Article 

    Google Scholar 
    Hebblewhite, M. & Whittington, J. Wolves without borders: Transboundary survival of wolves in Banff National Park over three decades. Glob. Ecol. Conserv. 24, e01293 (2020).Article 

    Google Scholar 
    Ministry of Forests Lands Natural Resource Operations and Rural Development. 2020–2022 Hunting and Trapping Regulations Synopsis, 96 (2020) https://www2.gov.bc.ca/assets/gov/sports-recreation-arts-and-culture/outdoor-recreation/fishing-and-hunting/hunting/regulations/2020-2022/hunting-trapping-synopsis-2020-2022.pdf (Accessed 15 Dec 2021).Persson, J., Landa, A., Andersen, R. & Segerström, P. Reproductive characteristics of female wolverines (Gulo gulo) in Scandinavia. J. Mammal. 87, 75–79 (2006).Article 

    Google Scholar 
    Persson, J. Female wolverine (Gulo gulo) reproduction: Reproductive costs and winter food availability. Can. J. Zool. 83, 1453–1459 (2005).Article 

    Google Scholar 
    Persson, J., Rauset, G. R. & Chapron, G. Paying for an endangered predator leads to population recovery. Conserv. Lett. 8, 345–350 (2015).Article 

    Google Scholar 
    Seip, D. R., Johnson, C. J. & Watts, G. S. Displacement of mountain caribou from winter habitat by snowmobiles. J. Wildl. Manag. 71, 1539–1544 (2007).Article 

    Google Scholar 
    Arlettaz, R. et al. Disturbance of wildlife by outdoor winter recreation: Allostatic stress response and altered activity-energy budgets. Ecol. Appl. 25, 1197–1212 (2015).Article 
    PubMed 

    Google Scholar 
    Olson, L. E., Squires, J. R., Roberts, E. K., Ivan, J. S. & Hebblewhite, M. Sharing the same slope: Behavioral responses of a threatened mesocarnivore to motorized and nonmotorized winter recreation. Ecol. Evol. 8, 8555–8572 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Ciuti, S. et al. Effects of humans on behaviour of wildlife exceed those of natural predators in a landscape of fear. PLoS ONE 7, e50611 (2012).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Phillips, G. E. & Alldredge, A. W. Reproductive success of elk following disturbance by humans during calving season. J. Wildl. Manag. 64, 521 (2000).Article 

    Google Scholar 
    Strasser, E. H. & Heath, J. A. Reproductive failure of a human-tolerant species, the American kestrel, is associated with stress and human disturbance. J. Appl. Ecol. 50, 912–919 (2013).Article 

    Google Scholar 
    Rauset, G. R., Low, M. & Persson, J. Reproductive patterns result from age-related sensitivity to resources and reproductive costs in a mammalian carnivore. Ecology 96, 3153–3164 (2015).Article 
    PubMed 

    Google Scholar 
    Inman, R. M., Magoun, A. J., Persson, J. & Mattisson, J. The wolverine’s niche: Linking reproductive chronology, caching, competition, and climate. J. Mammal. 93, 634–644 (2012).Article 

    Google Scholar 
    Persson, J., Willebrand, T., Landa, A., Andersen, R. & Segerström, P. The role of intraspecific predation in the survival of juvenile wolverines Gulo gulo. Wildl. Biol. 9, 21–28 (2003).Article 

    Google Scholar 
    Krebs, J., Lofroth, E. C. & Parfitt, I. Multiscale habitat use by wolverines in British Columbia, Canada. J. Wildl. Manag. 71, 2180–2192 (2007).Article 

    Google Scholar 
    Corradini, A. et al. Effects of cumulated outdoor activity on wildlife habitat use. Biol. Conserv. 253, 108818 (2021).Article 

    Google Scholar 
    Goodbody, T. R. H. et al. Mapping recreation and tourism use across grizzly bear recovery areas using social network data and maximum entropy modelling. Ecol. Modell. 440, 109377 (2021).Article 

    Google Scholar 
    May, R., Landa, A., Van Dijk, J., Linnell, J. D. C. & Andersen, R. Impact of infrastructure on habitat selection of wolverines Gulo gulo. Wildl. Biol. 12, 285–295 (2006).Article 

    Google Scholar 
    Scrafford, M. A., Avgar, T., Heeres, R. & Boyce, M. S. Roads elicit negative movement and habitat-selection responses by wolverines (Gulo gulo luscus). Behav. Ecol. 29, 534–542 (2018).Article 

    Google Scholar 
    Li, X., Zhou, Y., Zhao, M. & Zhao, X. A harmonized global nighttime light dataset 1992–2018. Sci. Data 7, 1–9 (2020).Article 

    Google Scholar 
    Lofroth, E. C., Krebs, J. A., Harrower, W. L. & Lewis, D. Food habits of wolverine Gulo gulo in montane ecosystems of British Columbia, Canada. Wildl. Biol. 13, 31–37 (2007).Article 

    Google Scholar 
    Hebblewhite, M., White, C. A. & Musiani, M. Revisiting extinction in national parks: Mountain Caribou in Banff. Conserv. Biol. 24, 341–344 (2010).Article 
    CAS 
    PubMed 

    Google Scholar 
    Poole, K. G. Kootenay Region Mountain Goat Population Assessment 2013–2015 (2015).Wasstrom, H. E., Cottell, C., Lofroth, E. C. & Larsen, K. W. Has the porcupine population waned in British Columbia? Trends in questionnaires and road-mortality data. Northwest. Nat. 101, 168–179 (2020).Article 

    Google Scholar 
    Hunt, W. A. Banff National Park State of the Park Report—Resource Conservation Technical Summaries 2008 to 2017 (2018).Kuzyk, G. et al. Moose population dynamics during 20 years of declining harvest in British Columbia. Alces 54, 101–119 (2018).
    Google Scholar 
    Kortello, A., Hausleitner, D. & Mowat, G. Mechanisms influencing the winter distribution of wolverine Gulo gulo Luscus in the southern Columbia Mountains, Canada. Wildl. Biol. 2019, 480 (2019).Article 

    Google Scholar 
    Copeland, J. P. et al. The bioclimatic envelope of the wolverine (Gulo gulo): Do climatic constraints limit its geographic distribution? Can. J. Zool. 88, 233–246 (2010).Article 

    Google Scholar 
    Magoun, A. J., Robards, M. D., Packila, M. L. & Glass, T. W. Detecting snow at the den-site scale in wolverine denning habitat. Wildl. Soc. Bull. 41, 381–387 (2017).Article 

    Google Scholar 
    Webb, S. M. et al. Distribution of female wolverines relative to snow cover, Alberta, Canada. J. Wildl. Manag. 80, 1461–1470 (2016).Article 

    Google Scholar 
    SARA Species at Risk Act. Order amending schedule 1 to the species at risk act. SOR/2018-112. Canada Gazette 152 (12), 18 June 2018. (2018) www.canada.ca/en/environment-%0Aclimate-change/services/species-risk-public-registry/%0Aorders/amend-schedule-1-volume-152-number-12-june-2018.%0Ahtml (Accessed on 15 December 2021).Holland, W. D. & Coen, G. M. Ecological (biophysical) land classification of Banff and Jasper National Parks (1983).open.canada.ca. https://open.canada.ca (Accessed on 15 December 2021).den Hartog, J. & Reijns, R. I3S Pattern+ (2016).Greenberg, S. Timelapse: An Image Classifer for Camera Traps (2020).Tobler, M. Camera Base Version (2007).Whittington, J., Hebblewhite, M. & Chandler, R. B. Generalized spatial mark–resight models with an application to grizzly bears. J. Appl. Ecol. 55, 157–168 (2018).Article 

    Google Scholar 
    Gowan, T. A., Crum, N. J. & Roberts, J. J. An open spatial capture–recapture model for estimating density, movement, and population dynamics from line-transect surveys. Ecol. Evol. 11, 7354–7365 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Royle, J. A., Fuller, A. K. & Sutherland, C. Spatial capture–recapture models allowing Markovian transience or dispersal. Popul. Ecol. 58, 53–62 (2016).Article 

    Google Scholar 
    Milleret, C. et al. Estimating abundance with interruptions in data collection using open population spatial capture–recapture models. Ecosphere 11, 1–14 (2020).Article 

    Google Scholar 
    Royle, J. A., Magoun, A. J., Gardner, B., Valkenburg, P. & Lowell, R. E. Density estimation in a wolverine population using spatial capture-recapture models. J. Wildl. Manag. 75, 604–611 (2011).Article 

    Google Scholar 
    Royle, J. A., Chandler, R. B., Sun, C. C. & Fuller, A. K. Reply to efford on ‘Integrating resource selection information with spatial capture–recapture’. Methods Ecol. Evol. 5, 603–605 (2014).Article 

    Google Scholar 
    Kendall, K. C. et al. Using bear rub data and spatial capture-recapture models to estimate trend in a brown bear population. Sci. Rep. 9, 5 (2019).Article 

    Google Scholar 
    Royle, J. A. & Young, K. V. A hierarchical model for spatial capture-recapture data. Ecology 89, 2281–2289 (2008).Article 
    PubMed 

    Google Scholar 
    Gardner, B., Sollmann, R., Kumar, N. S., Jathanna, D. & Karanth, K. U. State space and movement specification in open population spatial capture–recapture models. Ecol. Evol. 8, 10336–10344 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Watanabe, S. Asymptotic equivalence of Bayes cross validation and widely applicable information criterion in singular learning theory. J. Mach. Learn. Res. 11, 3571–3594 (2010).MathSciNet 
    MATH 

    Google Scholar 
    Royle, J. A. & Kéry, M. A Bayesian state-space formulation of dynamic occupancy models. Ecology 88, 1813–1823 (2007).Article 
    PubMed 

    Google Scholar 
    R Core Team. R: A Language and Environment for Statistical Computing (2022) https://www.r-project.org/ (Accessed on 20 June 2022).de Valpine, P. et al. Programming with models: Writing statistical algorithms for general model structures with NIMBLE. J. Comput. Graph. Stat. 26, 403–413 (2017).Article 
    MathSciNet 

    Google Scholar 
    Barrueto, M., Forshner, A., Whittington, J., Clevenger, A. P. & Musiani, M. Data from: Protection status, human disturbance, snow cover and trapping drive density of a declining wolverine population in the Canadian Rocky Mountains. Dryad. https://doi.org/10.5061/dryad.z34tmpghh (2022).Article 

    Google Scholar  More

  • in

    Contrasting sea ice conditions shape microbial food webs in Hudson Bay (Canadian Arctic)

    Laxon SW, Giles KA, Ridout AL, Wingham DJ, Willatt R, Cullen R, et al. CryoSat-2 estimates of Arctic sea ice thickness and volume. Geophys Res Lett. 2013;40:732–7.
    Google Scholar 
    Stroeve J, Notz D. Changing state of Arctic sea ice across all seasons. Environ Res Lett. 2018;13:103001.
    Google Scholar 
    Macdonald RW, Kuzyk ZZA. The Hudson Bay system: a northern inland sea in transition. J Mar Syst. 2011;88:337–40.
    Google Scholar 
    Serreze MC, Barry RG. Processes and impacts of Arctic amplification: a research synthesis. Glob Planet Change. 2011;77:85–96.
    Google Scholar 
    Hochheim KP, Barber DG. An update on the ice climatology of the Hudson Bay system. Arctic, Antarctic, and Alpine Research. 2014;46:66–83.
    Google Scholar 
    Gagnon AS, Gough WA. Climate change scenarios for the Hudson Bay region: an intermodel comparison. Climatic Change. 2005;69:269–97.
    Google Scholar 
    Bring A, Shiklomanov A, Lammers RB. Pan-Arctic river discharge: prioritizing monitoring of future climate change hot spots. Earths Future. 2017;5:72–92.
    Google Scholar 
    Comeau AM, Li WKW, Tremblay JÉ, Carmack EC, Lovejoy C. Arctic Ocean microbial community structure before and after the 2007 record sea ice minimum. PLoS One. 2011;6:e27492.PubMed 
    PubMed Central 

    Google Scholar 
    Li W, McLaughlin F, Lovejoy C, Carmack E. Smallest algae thrive as the Arctic Ocean freshens. Science. 2009;326:539.PubMed 

    Google Scholar 
    Ji R, Jin M, Varpe Ø. Sea ice phenology and timing of primary production pulses in the Arctic Ocean. Glob Change Biol. 2013;19:734–41.
    Google Scholar 
    Ardyna M, Mundy C, Mills M, Oziel L, Lacour L, Verin G, et al. Environmental drivers of under-ice phytoplankton bloom dynamics in the Arctic Ocean. Elem Sci Anthr. 2020;8:e30.
    Google Scholar 
    Kahru M, Brotas V, Manzano-Sarabia M, Mitchell BG. Are phytoplankton blooms occurring earlier in the Arctic? Glob Change Biol. 2011;17:1733–9.
    Google Scholar 
    Barbedo L, Bélanger S, Tremblay J. Climate control of sea-ice edge phytoplankton blooms in the Hudson Bay system. Elem Sci Anthr. 2020;8:1.
    Google Scholar 
    Matthes LC, Ehn JK, Dalman LA, Babb DG, Peeken I, Harasyn M, et al. Environmental drivers of spring primary production in Hudson Bay. Elem Sci Anthr. 2021;9:00160.
    Google Scholar 
    Harvey M, Therriault JC, Simard N. Late-summer distribution of phytoplankton in relation to water mass characteristics in Hudson Bay and Hudson Strait (Canada). Can J Fish Aquat Sci. 1997;54:1937–52.
    Google Scholar 
    Ferland J, Gosselin M, Starr M. Environmental control of summer primary production in the Hudson Bay system: The role of stratification. J Mar Syst. 2011;88:385–400.
    Google Scholar 
    Raven JA. The twelfth Tansley Lecture. Small is beautiful: the picophytoplankton. Funct Ecol. 1998;12:503–13.
    Google Scholar 
    Tilman D, Kilham SS, Kilham P. Phytoplankton community ecology: the role of limiting nutrients. Annu Rev Ecol Syst. 1982;13:349–72.
    Google Scholar 
    Onda DFL, Medrinal E, Comeau AM, Thaler M, Babin M, Lovejoy C. Seasonal and interannual changes in ciliate and dinoflagellate species assemblages in the Arctic Ocean (Amundsen Gulf, Beaufort Sea, Canada). Front Mar Sci. 2017;4:16.
    Google Scholar 
    Campbell K, Mundy CJ, Belzile C, Delaforge A, Rysgaard S. Seasonal dynamics of algal and bacterial communities in Arctic sea ice under variable snow cover. Polar Biol. 2018;41:41–58.
    Google Scholar 
    Forest A, Tremblay JÉ, Gratton Y, Martin J, Gagnon J, Darnis G, et al. Biogenic carbon flows through the planktonic food web of the Amundsen Gulf (Arctic Ocean): a synthesis of field measurements and inverse modeling analyses. Prog Oceanogr. 2011;91:410–36.
    Google Scholar 
    Buchan A, Lecleir GR, Gulvik CA, González JM. Master recyclers: features and functions of bacteria associated with phytoplankton blooms. Nat Rev Microbiol. 2014;12:686–98.PubMed 

    Google Scholar 
    Cole JJ, Likens GE, Strayer DL. Photosynthetically produced dissolved organic carbon: an important carbon source for planktonic bacteria. Limnol Oceanogr. 1982;27:1080–90.
    Google Scholar 
    Horňák K, Kasalický V, Šimek K, Grossart HP. Strain-specific consumption and transformation of alga-derived dissolved organic matter by members of the Limnohabitans-C and Polynucleobacter-B clusters of Betaproteobacteria. Environ Microbiol. 2017;19:4519–35.PubMed 

    Google Scholar 
    Šimek K, Kasalický V, Zapomělová E, Horňák K. Alga-derived substrates select for distinct betaproteobacterial lineages and contribute to niche separation in Limnohabitans strains. Appl Environ Microbiol. 2011;77:7307–15.PubMed 
    PubMed Central 

    Google Scholar 
    Orsi WD, Smith JM, Liu S, Liu Z, Sakamoto CM, Wilken S, et al. Diverse, uncultivated bacteria and archaea underlying the cycling of dissolved protein in the ocean. ISME J. 2016;10:2158–73.PubMed 
    PubMed Central 

    Google Scholar 
    Williams TJ, Wilkins D, Long E, Evans F, Demaere MZ, Raftery MJ, et al. The role of planktonic Flavobacteria in processing algal organic matter in coastal East Antarctica revealed using metagenomics and metaproteomics. Environ Microbiol. 2013;15:1302–17.PubMed 

    Google Scholar 
    Teeling H, Fuchs BM, Becher D, Klockow C, Gardebrechet A, Bennke CM, et al. Substrate-controlled succession of marine bacterioplankton populations induced by a phytoplankton bloom. Science. 2012;336:608–11.PubMed 

    Google Scholar 
    Armengol L, Calbet A, Franchy G, Rodríguez-Santos A, Hernández-León S. Planktonic food web structure and trophic transfer efficiency along a productivity gradient in the tropical and subtropical Atlantic Ocean. Sci Rep. 2019;9:1–19.
    Google Scholar 
    Joly S, Senneville S, Caya D, Saucier FJ. Sensitivity of Hudson Bay Sea ice and ocean climate to atmospheric temperature forcing. Clim Dyn. 2011;36:1835–49.
    Google Scholar 
    Kirillov S, Babb D, Dmitrenko I, Landy D, Lukovich JV, Ehn J, et al. Atmospheric forcing drives the winter sea ice thickness asymmetry of Hudson Bay. Geophys Res Oceans. 2020;125:1–12.
    Google Scholar 
    Tivy A, Howell SEL, Alt B, McCourt S, Chagnon R, Crocker G, et al. Trends and variability in summer sea ice cover in the Canadian Arctic based on the Canadian Ice Service Digital Archive, 1960-2008 and 1968-2008. J Geophys Res Oceans. 2011;116:C03007.
    Google Scholar 
    Barber D, Landry D, Babb D, Kirillov S, Aubry C, Schembri S, et al. Bay-Wide Survey Program Cruise Report – CCGS Amundsen (LEG-1). In: Hudson Bay System Study (BaySys) Phase 1 Report: Hudson Bay Field Program and Data Collection. Landry, DL & Candlish, LM. (Eds). 2019. p. 131–222.Jacquemot L, Kalenitchenko D, Matthes LC, Vigneron A, Mundy CJ, Tremblay JE, et al. Protist communities along freshwater-marine transition zones in Hudson Bay (Canada). Elem Sci Anthr. 2021;9:00111.
    Google Scholar 
    Grasshoff K, Kremling K, Ehrhardt M. Determination of nutrients. In: Methods of Seawater Analysis: Third, Completely Revised and Extended Edition. 1999. p. 159–228.Caporaso JG, Lauber CL, Walters WA, Berg-Lyons D, Lozupone CA, Turnbaugh PJ, et al. Global patterns of 16S rRNA diversity at a depth of millions of sequences per sample. Proc Natl Acad Sci. 2011;108:4516–22.PubMed 

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

    Google Scholar 
    Caporaso JG, Kuczynski J, Stombaugh J, Bittinger K, Bushman FD, Costello EK, et al. QIIME allows analysis of high- throughput community sequencing data. Nat Publ Group. 2010;7:335–6.
    Google Scholar 
    Guillou L, Bachar D, Audic S, Bass D, Berney C, Bittner L, et al. The Protist Ribosomal Reference database (PR2): a catalog of unicellular eukaryote Small Sub-Unit rRNA sequences with curated taxonomy. Nucleic Acids Res. 2013;41:597–604.
    Google Scholar 
    Quast C, Pruesse E, Yilmaz P, Gerken J, Schweer T, Yarza P, et al. The SILVA ribosomal RNA gene database project: improved data processing and web-based tools. Nucleic Acids Res. 2013;41:590–6.
    Google Scholar 
    McMurdie PJ, Holmes S. Phyloseq: an R package for reproducible interactive analysis and graphics of microbiome census data. PLoS One. 2013;8:e61217.PubMed 
    PubMed Central 

    Google Scholar 
    Stamatakis A. RAxML version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics. 2014;30:1312–3.PubMed 
    PubMed Central 

    Google Scholar 
    Oksanen J, Blanchet FG, Kindt R, Legendre P, Minchin PR, O’hara RB, et al. Package ‘vegan’. Community ecology package, version 2(9). 2013;1-295.Faust K, Raes J. CoNet app: inference of biological association networks using Cytoscape. F1000Research. 2016;5:1–14.
    Google Scholar 
    Shannon P, Markiel A, Ozier O, Baliga NS, Wang JT, Ramage D, et al. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Res. 2003;13:2498–504.PubMed 
    PubMed Central 

    Google Scholar 
    Volterra V. Fluctuations in the abundance of a species considered mathematically. Nature. 1926;118:558–60.
    Google Scholar 
    Joli N, Monier A, Logares R, Lovejoy C. Seasonal patterns in Arctic prasinophytes and inferred ecology of Bathycoccus unveiled in an Arctic winter metagenome. ISME J. 2017;11:1372–85.PubMed 
    PubMed Central 

    Google Scholar 
    Barber DG, Hop H, Mundy CJ, Else B, Dmitrenko IA, Tremblay JE, et al. Selected physical, biological and biogeochemical implications of a rapidly changing Arctic Marginal Ice Zone. Prog Oceanogr. 2015;139:122–50.
    Google Scholar 
    Tremblay JÉ, Anderson LG, Matrai P, Coupel P, Bélanger S, Michel C, et al. Global and regional drivers of nutrient supply, primary production and CO2 drawdown in the changing Arctic Ocean. Prog Oceanogr. 2015;139:171–96.
    Google Scholar 
    Needham DM, Fuhrman JA. Pronounced daily succession of phytoplankton, archaea and bacteria following a spring bloom. Nat Microbiol. 2016;1:1–7.
    Google Scholar 
    Berry D, Widder S. Deciphering microbial interactions and detecting keystone species with co-occurrence networks. Front Microbiol. 2014;5.Faust K, Raes J. Microbial interactions: from networks to models. Nat Rev Microbiol. 2012;10:538–50.PubMed 

    Google Scholar 
    Shi S, Nuccio EE, Shi ZJ, He Z, Zhou J, Firestone MK. The interconnected rhizosphere: high network complexity dominates rhizosphere assemblages. Ecol Lett. 2016;19:926–36.PubMed 

    Google Scholar 
    Guillou L, Viprey M, Chambouvet A, Welsh RM, Kirkham AR, Massana R, et al. Widespread occurrence and genetic diversity of marine parasitoids belonging to Syndiniales (Alveolata). Environ Microbiol. 2008;10:3349–65.PubMed 

    Google Scholar 
    Lima-Mendez G, Faust K, Henry N, Decelle J, Colin S, Carcillo F, et al. Determinants of community structure in the global plankton interactome. Science. 2015;10:1–10.
    Google Scholar 
    Chaffron S, Delage E, Budinich M, Vintache D, Henry N, Nef C, et al. Environmental vulnerability of the global ocean epipelagic plankton community interactome. Sci Adv. 2021;7:eabg1921.PubMed 
    PubMed Central 

    Google Scholar 
    Clarke LJ, Bestley S, Bissett A, Deagle BE. A globally distributed Syndiniales parasite dominates the Southern Ocean micro-eukaryote community near the sea-ice edge. ISME J. 2019;13:734–7.PubMed 

    Google Scholar 
    Not F, del Campo J, Balagué V, de Vargas C, Massana R. New insights into the diversity of marine picoeukaryotes. PLoS One. 2009;4:e7143.PubMed 
    PubMed Central 

    Google Scholar 
    Bass D, Stentiford GD, Littlewood DTJ, Hartikainen H. Diverse Applications of Environmental DNA Methods in Parasitology. Trends Parasitol. 2015;31:499–513.PubMed 

    Google Scholar 
    Kellogg CTE, Mcclelland JW, Dunton KH, Crump BC. Strong seasonality in arctic estuarine microbial food webs. Front Microbiol. 2019;10:2628.PubMed 
    PubMed Central 

    Google Scholar 
    Nitsche F, Weittere M, Scheckenbach F, Hausmann K, Wylezich C, Ardnt H. Deep sea records of choanoflagellates with a description of two new species. Acta Protozool. 2007;46:99–106.
    Google Scholar 
    Thaler M, Lovejoy C. Biogeography of heterotrophic flagellate populations indicates the presence of generalist and specialist taxa in the Arctic Ocean. Appl Environ Microbiol. 2015;81:2137–48.PubMed 
    PubMed Central 

    Google Scholar 
    Thomsen HA, Østergaard JB, Hansen LE. Loricate choanoflagellates from West Greenland (August 1988) including the description of Spinoeca buckii gen. et sp. nov. Eur J Protistol. 1995;31:38–44.
    Google Scholar 
    Thomsen HA, Østergaard JB. Acanthoecid choanoflagellates from the Atlantic Arctic Region − a baseline study. Heliyon. 2017;3:e00345.PubMed 
    PubMed Central 

    Google Scholar 
    Buck KR, Garrison DL, Cruz S. Distribution and abundance of choanoflagellates (Acanthoecidae) across the ice-edge zone in the Weddell Sea, Antarctica. Mar Biol. 1988;98:263–9.
    Google Scholar 
    Escalera L, Mangoni O, Bolinesi F, Saggiomo M. Austral summer bloom of loricate choanoflagellates in the central Ross Sea polynya. J Eukaryot Microbiol. 2019;66:849–52.PubMed 

    Google Scholar 
    Delmont TO, Hammar KM, Ducklow HW, Yager PL, Post AF. Phaeocystis antarctica blooms strongly influence bacterial community structures in the Amundsen Sea polynya. Front Microbiol. 2014;5:1–13.
    Google Scholar 
    Dupont CL, Rusch DB, Yooseph S, Lombardo MJ, Alexander Richter R, Valas R, et al. Genomic insights to SAR86, an abundant and uncultivated marine bacterial lineage. ISME J. 2012;6:1186–99.PubMed 

    Google Scholar 
    Abell GCJ, Bowman JP. Ecological and biogeographic relationships of class Flavobacteria in the Southern Ocean. FEMS Microbiol Ecol. 2005;51:265–77.PubMed 

    Google Scholar 
    Delmont TO, Murat Eren A, Vineis JH, Post AF. Genome reconstructions indicate the partitioning of ecological functions inside a phytoplankton bloom in the Amundsen Sea, Antarctica. Front Microbiol. 2015;6:e1090.
    Google Scholar 
    Stingl U, Desiderio RA, Cho JC, Vergin KL, Giovannoni SJ. The SAR92 clade: an abundant coastal clade of culturable marine bacteria possessing proteorhodopsin. Appl Environ Microbiol. 2007;73:2290–6.PubMed 
    PubMed Central 

    Google Scholar 
    de Sousa AGG, Tomasino MP, Duarte P, Fernández-Méndez M, Assmy P, Ribeiro H, et al. Diversity and composition of pelagic prokaryotic and protist communities in a thin Arctic sea-ice regime. Microb Ecol. 2019;78:388–408.PubMed 

    Google Scholar 
    Rapp JZ, Fernández-Méndez M, Bienhold C, Boetius A. Effects of ice-algal aggregate export on the connectivity of bacterial communities in the central Arctic Ocean. Front Microbiol. 2018;9:1035.PubMed 
    PubMed Central 

    Google Scholar 
    Wemheuer B, Güllert S, Billerbeck S, Giebel HA, Voget S, Simon M, et al. Impact of a phytoplankton bloom on the diversity of the active bacterial community in the southern North Sea as revealed by metatranscriptomic approaches. FEMS Microbiol Ecol. 2014;87:378–89.PubMed 

    Google Scholar 
    Jain A, Krishnan KP. Differences in free-living and particle-associated bacterial communities and their spatial variation in Kongsfjorden, Arctic. J Basic Microbiol. 2017;57:827–38.PubMed 

    Google Scholar 
    Granskog MA, Kuzyk ZZA, Azetsu-Scott K, Macdonald RW. Distributions of runoff, sea-ice melt and brine using δ18o and salinity data – a new view on freshwater cycling in Hudson Bay. J Mar Syst. 2011;88:362–74.
    Google Scholar 
    Stegen JC, Lin X, Konopka AE, Fredrickson JK. Stochastic and deterministic assembly processes in subsurface microbial communities. ISME J. 2012;6:1653–64.PubMed 
    PubMed Central 

    Google Scholar 
    Boeuf D, Edwards BR, Eppley JM, Hu SK, Poff KE, Romano AE, et al. Biological composition and microbial dynamics of sinking particulate organic matter at abyssal depths in the oligotrophic open ocean. Proc Natl Acad Sci USA. 2019;116:11824–32.PubMed 
    PubMed Central 

    Google Scholar 
    Gutierrez-Rodriguez A, Stukel MR, Lopes dos Santos A, Biard T, Scharek R, Vaulot D, et al. High contribution of Rhizaria (Radiolaria) to vertical export in the California Current Ecosystem revealed by DNA metabarcoding. ISME J. 2019;13:964–76.PubMed 

    Google Scholar 
    Bråte J, Krabberød AK, Dolven JK, Ose RF, Kristensen T, Bjørklund KR, et al. Radiolaria associated with large diversity of marine alveolates. Protist. 2012;163:767–77.PubMed 

    Google Scholar 
    Dolven JK, Lindqvist C, Albert VA, Bjørklund KR, Yuasa T, Takahashi O, et al. Molecular diversity of alveolates associated with neritic North Atlantic radiolarians. Protist. 2007;158:65–76.PubMed 

    Google Scholar 
    Decelle J, Martin P, Paborstava K, Pond DW, Tarling G, Mahé F, et al. Diversity, ecology and biogeochemistry of cyst-forming Acantharia (Radiolaria) in the oceans. PLoS One. 2013;8:e53598.PubMed 
    PubMed Central 

    Google Scholar 
    Fernandes GL, Shenoy BD, Damare SR. Diversity of bacterial community in the oxygen minimum zones of Arabian Sea and Bay of Bengal as deduced by illumina sequencing. Front Microbiol. 2020;10:e3153.
    Google Scholar 
    Vigneron A, Cruaud P, Culley A, Couture RM, Lovejoy C, Vincent W. Genomic evidence for sulfur intermediates as new biogeochemical hubs in a model aquatic microbial ecosystem. Microbiome. 2020;9:e46.
    Google Scholar 
    Wright JJ, Konwar KM, Hallam SJ. Microbial ecology of expanding oxygen minimum zones. Nat Publ Group. 2012;10:381–94.
    Google Scholar 
    Bianchi D, Weber TS, Kiko R, Deutsch C. Global niche of marine anaerobic metabolisms expanded by particle microenvironments. Nat Geosci. 2018;11:263–68.
    Google Scholar 
    Michel C, Legendre L, Therriault JC, Demers S, Vandevelde T. Springtime coupling between ice algal and phytoplankton assemblages in southeastern Hudson Bay, Canadian Arctic. Polar Biol. 1993;13:441–9.
    Google Scholar 
    Boetius A, Albrecht S, Bakker K, Bienhold C, Felden J, Fernández-méndez M, et al. Export of algal biomass from the metling Arctic sea ice. Science. 2013;339:1430–2.PubMed 

    Google Scholar 
    Vigneron A, Lovejoy C, Cruaud P, Kalenitchenko D, Culley A, Vincent WF. Contrasting winter versus summer microbial communities and metabolic functions in a permafrost thaw lake. Front Microbiol. 2019;10:e1656.
    Google Scholar 
    Tremblay JÉ, Lee J-H, Gosselin M, Belanger S. Nutrient dynamic and marine biological productivity in the greater Hudson Bay marine region. In: An integrated regional impact study (IRIS) Arcticnet University of Manitoba and ArcticNet. 2019. p. 225–44.Wassmann P, Reigstad M. Future Arctic Ocean seasonal ice zones and implications for pelagic-benthic coupling. Oceanography. 2011;24:220–31.
    Google Scholar  More