More stories

  • in

    Trees outside forests are an underestimated resource in a country with low forest cover

    1.Turner, W. R., Nakamura, T. & Dinetti, M. Global urbanization and the separation of humans from nature. Bioscience 54, 585–590 (2004).Article 

    Google Scholar 
    2.Schnell, S., Kleinn, C. & Ståhl, G. Monitoring trees outside forests: a review. Environ. Monitor. Assess. 187, 600 (2015).Article 

    Google Scholar 
    3.Ahmed, P. Trees outside forests (TOF): a case study of wood production and consumption in Haryana. Int. For. Rev. 10, 165–172 (2008).
    Google Scholar 
    4.Krishnankutty, C. N., Thampi, K. B. & Chundamannil, M. Trees outside forests (TOF): a case study of the wood productionconsumption situation in Kerala. Int. For. Rev. 10, 156–164 (2008).
    Google Scholar 
    5.Smeets, E. M. W. & Faaij, A. P. C. Bioenergy potentials from forestry in 2050: an assessment of the drivers that determine the potentials. Clim. Change 81, 353–390 (2007).CAS 
    Article 
    ADS 

    Google Scholar 
    6.Schnell, S., Altrell, D., Ståhl, G. & Kleinn, C. The contribution of trees outside forests to national tree biomass and carbon stocks-a comparative study across three continents. Environ. Monitor. Assess. 187, 4197 (2015).Article 

    Google Scholar 
    7.Zomer, R. J. et al. Trees on farms: an update and reanalysis of agroforestry’s global extent and socio-ecological characteristics. World Agroforestry Center Working Paper 179 (2014).8.Ghosh, M. & Sinha, B. Policy analysis for realizing the potential of timber production from trees outside forests (TOF) in India. Int. For. Rev. 20, 89–103 (2018).
    Google Scholar 
    9.Pain-Orcet, M. & Bellefontaine, R. Trees outside the forest: a new perspective on the management of forest resources in the tropics. Beyond tropical deforestation: from tropical deforestation to forest cover dynamics and forest development, 423–430 (2004)10.Bellefontaine, R., Petit, S., Pain Orcet, M., Deleporte, P. & Bertault, J.G. Trees outside forests: towards better awareness. Food and Agriculture Organization, 216 (Rome, 2002)11.Kleinn, C. On large-area inventory and assessment of trees outside forests. UNASYLVA-FAO- 3–10 (2000).12.FAO. Global forest resources assessment 2005: Progress towards sustainable forest management. Food and Agriculture Organization of the United Nations (2006).13.Eggleston, S., Buendia, L., Miwa, K., Ngara, T. & Tanabe, K. 2006 IPCC guidelines for national greenhouse gas inventories Vol. 5 (Institute for Global Environmental Strategies Hayama, Japan, 2006).14.FAO. World Urbanization Prospects the Revision 2012 (Technical Report, 2011).15.Tewari, V. P., Sukumar, R., Kumar, R. & Gadow, K. Forest observational studies in India: past developments and considerations for the future. For. Ecol. Manag. 316, 32–46 (2014).Article 

    Google Scholar 
    16.Nath, T. K. & Inoue, M. Impacts of participatory forestry on livelihoods of ethnic people: experience from Bangladesh. Soc. Nat. Resour. 23, 1093–1107 (2010).Article 

    Google Scholar 
    17.Islam, S.S. Stratified Two-Stage Sampling (Self-Weighted) for assessment of village forest resources. J. Trop. For. Sci., 9–16 (2004)18.Zashimuddin, M. Community forestry for poverty reduction in Bangladesh. For. Poverty Reduct. Commun. For. Make Money, 81–94 (2004).19.FAO. Global Forest Resources Assessment 2015. Technical Report, Rome (2015).20.Muhammed, N., Koike, M. & Haque, F. Forest policy and sustainable forest management in Bangladesh: an analysis from national and international perspectives. New For. 36, 201–216 (2008).Article 

    Google Scholar 
    21.Manning, A. D., Fischer, J. & Lindenmayer, D. B. Scattered trees are keystone structures-implications for conservation. Biol. Conserv. 132, 311–321 (2006).Article 

    Google Scholar 
    22.Potapov, P. et al. Comprehensive monitoring of Bangladesh tree cover inside and outside of forests, 2000–2014. Environ. Res. Lett. 12, 104015 (2017).Article 
    ADS 

    Google Scholar 
    23.Schumacher, J. & Nord-Larsen, T. Wall-to-wall tree type classification using airborne lidar data and CIR images. Int. J. Remote Sens. 35, 3057–3073 (2014).Article 
    ADS 

    Google Scholar 
    24.Ouma, Y. O. & Tateishi, R. Urban-trees extraction from Quickbird imagery using multiscale spectex-filtering and non-parametric classification. ISPRS J. Photogramm. Remote Sens. 63, 333–351 (2008).Article 
    ADS 

    Google Scholar 
    25.Levin, N. et al. Mapping forest patches and scattered trees from SPOT images and testing their ecological importance for woodland birds in a fragmented agricultural landscape. Int. J. Remote Sens. 30, 3147–3169 (2009).Article 
    ADS 

    Google Scholar 
    26.Sandberg, G., Ulander, L. M. H., Wallerman, J. & Fransson, J. E. S. Measurements of forest biomass change using P-band synthetic aperture radar backscatter. IEEE Trans. Geosci. Remote Sens. 52, 6047–6061 (2014).Article 
    ADS 

    Google Scholar 
    27.Mitchard, E. T. A. et al. Mapping tropical forest biomass with radar and spaceborne LiDAR in Lopé National Park, Gabon: overcoming problems of high biomass and persistent cloud. Biogeosciences 9, 179–191 (2012).Article 
    ADS 

    Google Scholar 
    28.Minh, D. H. T. et al. Relating P-band synthetic aperture radar tomography to tropical forest biomass. IEEE Trans. Geosci. Remote Sens. 52, 967–979 (2013).Article 

    Google Scholar 
    29.Stovall, A. E. L., Shugart, H. & Yang, X. Tree height explains mortality risk during an intense drought. Nat. Commun. 10, 1–6 (2019).Article 

    Google Scholar 
    30.Swatantran, A., Tang, H., Barrett, T., DeCola, P. & Dubayah, R. Rapid, high-resolution forest structure and terrain mapping over large areas using single photon lidar. Sci. Rep. 6, 28277 (2016).CAS 
    Article 
    ADS 

    Google Scholar 
    31.Stovall, A. E. L. & Shugart, H. H. Improved biomass calibration and validation with terrestrial LiDAR: implications for future LiDAR and SAR missions. IEEE J. Sel. Top. Appl. Earth Observ. Remote Sens. 11, 3527–3537 (2018).Article 
    ADS 

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

    Google Scholar 
    33.Martone, M. et al. The global forest/non-forest map from TanDEM-X interferometric SAR data. Remote Sens. Environ. 205, 352–373 (2018).Article 
    ADS 

    Google Scholar 
    34.Lang, N., Schindler, K. & Wegner, J. D. Country-wide high-resolution vegetation height mapping with Sentinel-2. Remote Sens. Environ. 233, 111347 (2019).Article 
    ADS 

    Google Scholar 
    35.UNFAO. The State of World fisheries and Aquaculture 2014, vol. 24 (2014).36.Reddy, C. S., Pasha, S. V., Jha, C. S., Diwakar, P. G. & Dadhwal, V. K. Development of national database on long-term deforestation (1930–2014) in Bangladesh. Glob. Planet. Change 139, 173–182 (2016).Article 
    ADS 

    Google Scholar 
    37.Long, A. J. & Nair, P. K. R. Trees outside forests: agro-, community, and urban forestry. In Planted Forests: Contributions to the Quest for Sustainable Societies, 145–174 (Springer, 1999).38.Muhammed, N., Masum, M. F. H., Hossain, M. M., Chakma, S. & Oesten, G. Economic dependence of rural people on homestead forestry in Mymensingh, Bangladesh. J. For. Res. 24, 591–597 (2013).Article 

    Google Scholar 
    39.Motiur, R. M., Furukawa, Y., Kawata, I., Rahman, M. M. & Alam, M. Role of homestead forests in household economy and factors affecting forest production: a case study in southwest Bangladesh. J. For. Res. 11, 89–97 (2006).Article 

    Google Scholar 
    40.Salam, M. A., Noguchi, T. & Koike, M. Understanding why farmers plant trees in the homestead agroforestry in Bangladesh. Agrofor. Syst. 50, 77–93 (2000).Article 

    Google Scholar 
    41.Rossi, J.-P. & Rousselet, J. The spatial distribution of trees outside forests in a large open-field region and its potential impact on habitat connectivity for forest insects. Türkiye Ormancılık Dergisi 17, 62–64 (2016).Article 

    Google Scholar 
    42.Kabir, M. E. & Webb, E. L. Can homegardens conserve biodiversity in Bangladesh?. Biotropica 40, 95–103 (2008).
    Google Scholar 
    43.Gibbons, P. et al. The future of scattered trees in agricultural landscapes. Conserv. Biol. 22, 1309–1319 (2008).CAS 
    Article 

    Google Scholar 
    44.World Bank. No Title (2018).45.Olson, D. M. et al. Terrestrial ecoregions of the world: a new map of life on eartha new global map of terrestrial ecoregions provides an innovative tool for conserving biodiversity. BioScience 51, 933–938 (2001).Article 

    Google Scholar 
    46.Shimada, M. et al. New global forest/non-forest maps from ALOS PALSAR data (2007–2010). Remote Sens. Environ. 155, 13–31. https://doi.org/10.1016/j.rse.2014.04.014 (2014).Article 
    ADS 

    Google Scholar 
    47.GDAL/OGR contributors (2019). GDAL/OGR Geospatial Data Abstraction software Library. Open Source Geospatial Foundation. https://gdal.org.48.Montesano, P. M., Sun, G., Dubayah, R. & Ranson, K. J. The uncertainty of plot-scale forest height estimates from complementary spaceborne observations in the taiga-tundra ecotone. Remote Sens. 6, 10070–10088 (2014).Article 
    ADS 

    Google Scholar 
    49.Montesano, P. M., Sun, G., Dubayah, R. O. & Ranson, K. J. Spaceborne potential for examining taiga-tundra ecotone form and vulnerability. Biogeosciences 13, 3847–3861 (2016).Article 
    ADS 

    Google Scholar 
    50.Montesano, P. M. et al. The use of sun elevation angle for stereogrammetric boreal forest height in open canopies. Remote Sens. Environ. 196, 76–88 (2017).Article 
    ADS 

    Google Scholar 
    51.Olofsson, P. et al. Good practices for estimating area and assessing accuracy of land change. Remote Sens. Environ. 148, 42–57 (2014).Article 
    ADS 

    Google Scholar  More

  • in

    On species delimitation, hybridization and population structure of cassava whitefly in Africa

    1.Hebert, P. D., Penton, E. H., Burns, J. M., Janzen, D. H. & Hallwachs, W. T. species in one: DNA barcoding reveals cryptic species in the neotropical skipper butterfly Astraptes fulgerator. Proc. Natl. Acad. Sci. U.S.A. 101, 14812–14817. https://doi.org/10.1073/pnas.0406166101 (2004).CAS 
    Article 
    PubMed 
    PubMed Central 
    ADS 

    Google Scholar 
    2.Saez, A. G. & Lozano, E. Body doubles. Nature 433, 111. https://doi.org/10.1038/433111a (2005).CAS 
    Article 
    PubMed 
    ADS 

    Google Scholar 
    3.Vyskočilová, S., Tay, W. T., van Brunschot, S., Seal, S. & Colvin, J. An integrative approach to discovering cryptic species within the Bemisia tabaci whitefly species complex. Sci. Rep. 8, 10886. https://doi.org/10.1038/s41598-018-29305-w (2018).CAS 
    Article 
    PubMed 
    PubMed Central 
    ADS 

    Google Scholar 
    4.Liu, S. S. et al. Asymmetric mating interactions drive widespread invasion and displacement in a whitefly. Science 318, 1769–1772. https://doi.org/10.1126/science.1149887 (2007).CAS 
    Article 
    PubMed 
    ADS 

    Google Scholar 
    5.Vyskocilova, S., Seal, S. & Colvin, J. Relative polyphagy of “Mediterranean” cryptic Bemisia tabaci whitefly species and global pest status implications. J. Pest Sci. 92, 1071–1088. https://doi.org/10.1007/s10340-019-01113-9 (2019).Article 

    Google Scholar 
    6.Behere, G. T., Tay, W. T., Russell, D. A. & Batterham, P. Molecular markers to discriminate among four pest species of Helicoverpa (Lepidoptera: Noctuidae). Bull. Entomol. Res. 98, 599–603. https://doi.org/10.1017/S0007485308005956 (2008).CAS 
    Article 
    PubMed 

    Google Scholar 
    7.Elfekih, S., Tay, W. T., Gordon, K., Court, L. N. & De Barro, P. J. Standardized molecular diagnostic tool for the identification of cryptic species within the Bemisia tabaci complex. Pest Manag. Sci. 74, 170–173. https://doi.org/10.1002/ps.4676 (2018).CAS 
    Article 
    PubMed 

    Google Scholar 
    8.Walsh, T. K. et al. Mitochondrial DNA genomes of five major Helicoverpa pest species from the Old and New Worlds (Lepidoptera: Noctuidae). Ecol. Evol. 9, 2933–2944. https://doi.org/10.1002/ece3.4971 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    9.Anderson, C. J., Tay, W. T., McGaughran, A., Gordon, K. & Walsh, T. K. Population structure and gene flow in the global pest, Helicoverpa armigera. Mol. Ecol. 25, 5296–5311. https://doi.org/10.1111/mec.13841 (2016).CAS 
    Article 
    PubMed 

    Google Scholar 
    10.Elfekih, S. et al. Genome-wide analyses of the Bemisia tabaci species complex reveal contrasting patterns of admixture and complex demographic histories. PLoS ONE 13, e0190555. https://doi.org/10.1371/journal.pone.0190555 (2018).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    11.Anderson, C. J. et al. Hybridization and gene flow in the mega-pest lineage of moth, Helicoverpa. Proc. Natl. Acad. Sci. U.S.A. 115, 5034–5039. https://doi.org/10.1073/pnas.1718831115 (2018).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    12.FAOSTAT. http://www.fao.org/faostat/en/#data/QC/visualize (2017).13.Legg, J. P. et al. Spatio-temporal patterns of genetic change amongst populations of cassava Bemisia tabaci whiteflies driving virus pandemics in East and Central Africa. Virus Res. 186, 61–75. https://doi.org/10.1016/j.virusres.2013.11.018 (2014).CAS 
    Article 
    PubMed 

    Google Scholar 
    14.Patil, B. L. & Fauquet, C. M. Cassava mosaic geminiviruses: Actual knowledge and perspectives. Mol. Plant Pathol. 10, 685–701. https://doi.org/10.1111/j.1364-3703.2009.00559.x (2009).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    15.Macfadyen, S. et al. Cassava whitefly, Bemisia tabaci (Gennadius) (Hemiptera: Aleyrodidae) in East African farming landscapes: A review of the factors determining abundance. Bull. Entomol. Res. 108, 565–582. https://doi.org/10.1017/S0007485318000032 (2018).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    16.Minato, N. et al. Surveillance for Sri Lankan cassava mosaic virus (SLCMV) in Cambodia and Vietnam one year after its initial detection in a single plantation in 2015. PLoS ONE 14, e0212780. https://doi.org/10.1371/journal.pone.0212780 (2019).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    17.Wang, H. L. et al. First Report of Sri Lankan cassava mosaic virus Infecting Cassava in Cambodia. Plant Dis. 100, 1029–1029. https://doi.org/10.1094/Pdis-10-15-1228-Pdn (2016).Article 

    Google Scholar 
    18.De Barro, P. J., Liu, S. S., Boykin, L. M. & Dinsdale, A. B. Bemisia tabaci: A statement of species status. Annu. Rev. Entomol. 56, 1–19. https://doi.org/10.1146/annurev-ento-112408-085504 (2011).CAS 
    Article 
    PubMed 

    Google Scholar 
    19.Hopkinson, J. et al. Insecticide resistance status of Bemisia tabaci MEAM1 (Hemiptera: Aleyrodidae) in Australian cotton production valleys. Austral Entomol. 59, 202–214 (2020).Article 

    Google Scholar 
    20.Hadjistylli, M., Roderick, G. K. & Gauthier, N. First report of the Sub-Saharan Africa 2 species of the Bemisia tabaci complex in the Southern France. Phytoparasitica 43, 679–687. https://doi.org/10.1007/s12600-015-0480-3 (2015).Article 

    Google Scholar 
    21.Lee, W., Park, J., Lee, G. S., Lee, S. & Akimoto, S. Taxonomic status of the Bemisia tabaci complex (Hemiptera: Aleyrodidae) and reassessment of the number of its constituent species. PLoS ONE 8, e63817. https://doi.org/10.1371/journal.pone.0063817 (2013).CAS 
    Article 
    PubMed 
    PubMed Central 
    ADS 

    Google Scholar 
    22.Mugerwa, H. et al. African ancestry of New World, Bemisia tabaci-whitefly species. Sci. Rep. 8, 2734. https://doi.org/10.1038/s41598-018-20956-3 (2018).CAS 
    Article 
    PubMed 
    PubMed Central 
    ADS 

    Google Scholar 
    23.Martin, J. H. An identification guide to common whitefly pest species of the world (Homopt Aleyrodidae). Int. J. Pest Manag. 334, 298–322. https://doi.org/10.1080/09670878709371174 (1987).Article 

    Google Scholar 
    24.Martin, J. H. & Mound, L. A. An annotated check list of the world’s whiteflies (Insecta: Hemiptera: Aleyrodidae). Zootaxa 1492, 1–84 (2007).Article 

    Google Scholar 
    25.Mound, L. A. Host-correlated variation in Bemisia tabaci (Gennadius). Proc. R. Entomol. Soc. Lond. A38, 171–180 (1963).ADS 

    Google Scholar 
    26.Tay, W. T. et al. Novel molecular approach to define pest species status and tritrophic interactions from historical Bemisia specimens. Sci. Rep. https://doi.org/10.1038/s41598-017-00528-7 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    27.Tay, W. T., Evans, G. A., Boykin, L. M. & De Barro, P. J. Will the real Bemisia tabaciplease stand up?. PLoS ONE https://doi.org/10.1371/journal.pone.0050550 (2012).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    28.Dinsdale, A., Cook, L., Riginos, C., Buckley, Y. M. & De Barro, P. Refined global analysis of Bemisia tabaci (Hemiptera: Sternorrhyncha: Aleyrodoidea: Aleyrodidae) mitochondrial cytochrome oxidase 1 to identify species level genetic boundaries. Ann. Entomol. Soc. Am. 103, 196–208. https://doi.org/10.1603/An09061 (2010).Article 

    Google Scholar 
    29.Kunz, D., Tay, W. T., Elfekih, S., Gordon, K. H. J. & De Barro, P. J. Take out the rubbish – Removing NUMTs and pseudogenes from the Bemisia tabacicryptic species mtCOI database. bioRxiv. https://doi.org/10.1101/724765 (2019).Article 

    Google Scholar 
    30.Wongnikong, W., van Brunschot, S. L., Hereward, J. P., De Barro, P. J. & Walter, G. H. Testing mate recognition through reciprocal crosses of two native populations of the whitefly Bemisia tabaci (Gennadius) in Australia. Bull. Entomol. Res. 110, 328–339. https://doi.org/10.1017/S0007485319000683 (2020).CAS 
    Article 
    PubMed 

    Google Scholar 
    31.Mugerwa, H., Wang, H.-L., Sseruwagi, P., Seal, S. & Colvin, J. Whole-genome single nucleotide polymorphism and mating compatibility studies reveal the presence of distinct species in sub-Saharan Africa Bemisia tabaci whiteflies. Insect Sci. https://doi.org/10.1111/1744-7917.12881 (2020).Article 
    PubMed 

    Google Scholar 
    32.Delatte, H. et al. A new silverleaf-inducing biotype Ms of Bemisia tabaci (Hemiptera: Aleyrodidae) indigenous to the islands of the south-west Indian Ocean. Bull. Entomol. Res. 95, 29–35. https://doi.org/10.1079/Ber2004337 (2005).CAS 
    Article 
    PubMed 

    Google Scholar 
    33.Boykin, L. M., Savill, A. & De Barro, P. Updated mtCOI reference dataset for the Bemisia tabaci species complex. F1000Research 6, 1835. https://doi.org/10.12688/f1000research.12858.1 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    34.Liu, S. S., Colvin, J. & De Barro, P. J. Species concepts as applied to the whitefly Bemisia tabaci systematics: How many species are there?. J Integr Agr 11, 176–186. https://doi.org/10.1016/S2095-3119(12)60002-1 (2012).Article 

    Google Scholar 
    35.Tay, W. T. et al. The trouble with MEAM2: Implications of pseudogenes on species delimitation in the globally invasive Bemisia tabaci (Hemiptera: Aleyrodidae) cryptic species complex. Genome Biol. Evol. 9, 2732–2738. https://doi.org/10.1093/gbe/evx173 (2017).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    36.Kunz, D. et al. Draft mitochondrial DNA genome of a 1920 Barbados cryptic Bemisia tabaci “New World” species (Hemiptera: Aleyrodidae). Mitochondrial DNA B 4, 1183–1184. https://doi.org/10.1080/23802359.2019.1591197 (2019).Article 

    Google Scholar 
    37.Paini, D. R. et al. Global threat to agriculture from invasive species. Proc. Natl. Acad. Sci. U.S.A. 113, 7575–7579. https://doi.org/10.1073/pnas.1602205113 (2016).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    38.Wosula, E. N., Chen, W. B., Fei, Z. J. & Legg, J. P. Unravelling the genetic diversity among cassava Bemisia tabaci whiteflies using NextRAD sequencing. Genome Biol. Evol. 9, 2958–2973. https://doi.org/10.1093/gbe/evx219 (2017).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    39.Thresh, J. M., Fargette, D. & Otim-Nape, G. W. Effects of African cassava mosaic geminivirus on the yield of cassava. Trop. Sci. 34, 26–42 (1994).
    Google Scholar 
    40.Legg, J. et al. A global alliance declaring war on cassava viruses in Africa. Food Secur. 6, 231–248. https://doi.org/10.1007/s12571-014-0340-x (2014).Article 

    Google Scholar 
    41.Legg, J. P. et al. Biology and management of Bemisia whitefly vectors of cassava virus pandemics in Africa. Pest Manag. Sci. 70, 1446–1453. https://doi.org/10.1002/ps.3793 (2014).CAS 
    Article 
    PubMed 

    Google Scholar 
    42.Berry, S. D. et al. Molecular evidence for five distinct Bemisia tabaci (Homoptera: Aleyrodidae) geographic haplotypes associated with cassava plants in sub-Saharan Africa. Ann. Entomol. Soc. Am. 97, 852–859. https://doi.org/10.1603/0013-8746(2004)097[0852:Meffdb]2.0.Co;2 (2004).CAS 
    Article 

    Google Scholar 
    43.Mugerwa, H., Rey, M. E. C., Tairo, F., Ndunguru, J. & Sseruwagi, P. Two sub-Saharan Africa 1 populations of Bemisia tabaci exhibit distinct biological differences in fecundity and survivorship on cassava. Crop Prot. 117, 7–14. https://doi.org/10.1016/j.cropro.2018.11.011 (2019).Article 

    Google Scholar 
    44.Ghosh, S., Bouvaine, S. & Maruthi, M. N. Prevalence and genetic diversity of endosymbiotic bacteria infecting cassava whiteflies in Africa. BMC Microbiol. https://doi.org/10.1186/s12866-015-0425-5 (2015).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    45.Weir, B. S. & Cockerham, C. C. Estimating F-statistics for the analysis of population-structure. Evolution 38, 1358–1370. https://doi.org/10.2307/2408641 (1984).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    46.Ghosh, S., Bouvaine, S., Richardson, S. C. W., Ghanim, M. & Maruthi, M. N. Fitness costs associated with infections of secondary endosymbionts in the cassava whitefly species Bemisia tabaci. J. Pest Sci. 91, 17–28. https://doi.org/10.1007/s10340-017-0910-8 (2018).Article 

    Google Scholar 
    47.Elfekih, S. et al. Evolutionary genomics of Bemisia tabaci and characterization of its endosymbiont metacommunities using nextRAD sequencing. International Plant and Animal Genome Asia, Singapore 23–25 July 2015 (2015).48.Elfekih, S. et al. Genome-wide SNPs Decipher Global Incursion pathways in the Bemisia tabaci species complex. International Plant and Animal Genome Conferences San Diego, 9–13 January 2016 (2016).49.Elfekih, S. et al. Genome-wide scans unravel fine-scale invasion routes in the Bemisia tabaci species complex. 2nd International Whitefly Symposium, Arusha, Tanzania. p38, 14–19 February 2016 (2016).50.Boykin, L. M., Bell, C. D., Evans, G., Small, I. & De Barro, P. J. Is agriculture driving the diversification of the Bemisia tabaci species complex (Hemiptera: Sternorrhyncha: Aleyrodidae)? Dating, diversification and biogeographic evidence revealed. BMC Evol. Biol. 13, 228. https://doi.org/10.1186/1471-2148-13-228 (2013).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    51.Boykin, L. M. et al. Review and guide to a future naming system of African Bemisia tabaci species. Syst. Entomol. 43, 427–433. https://doi.org/10.1111/syen.12294 (2018).Article 

    Google Scholar 
    52.Hebert, P. D. N., Cywinska, A., Ball, S. L. & DeWaard, J. R. Biological identifications through DNA barcodes. Proc. R. Soc. B Biol. Sci. 270, 313–321. https://doi.org/10.1098/rspb.2002.2218 (2003).CAS 
    Article 

    Google Scholar 
    53.Hanemaaijer, M. J. et al. Mitochondrial genomes of Anophelesarabiensis, An. gambiae and An. coluzzii show no clear species division [version 2; peer review: 2 approved]. F1000Research 7, 347. https://doi.org/10.12688/f1000research.13807.2 (2019).Article 
    PubMed Central 

    Google Scholar 
    54.Tabachnick, W. J. Culicoides variipennis and bluetongue-virus epidemiology in the United States. Annu. Rev. Entomol. 41, 23–43. https://doi.org/10.1146/annurev.en.41.010196.000323 (1996).CAS 
    Article 
    PubMed 

    Google Scholar 
    55.Legg, J. P., French, R., Rogan, D., Okao-Okuja, G. & Brown, J. K. A distinct Bemisia tabaci (Gennadius) (Hemiptera: Sternorrhyncha: Aleyrodidae) genotype cluster is associated with the epidemic of severe cassava mosaic virus disease in Uganda. Mol. Ecol. 11, 1219–1229. https://doi.org/10.1046/j.1365-294X.2002.01514.x (2002).CAS 
    Article 
    PubMed 

    Google Scholar 
    56.Colvin, J., Omongo, C. A., Maruthi, M. N., Otim-Nape, G. W. & Thresh, J. M. Dual begomovirus infections and high Bemisia tabaci populations: Two factors driving the spread of a cassava mosaic disease pandemic. Plant Pathol. 53, 577–584. https://doi.org/10.1111/j.1365-3059.2004.01062.x (2004).Article 

    Google Scholar 
    57.Polston, J. E., De Barro, P. & Boykin, L. M. Transmission specificities of plant viruses with the newly identified species of the Bemisia tabaci species complex. Pest Manag. Sci. 70, 1547–1552. https://doi.org/10.1002/ps.3738 (2014).CAS 
    Article 
    PubMed 

    Google Scholar 
    58.Ally, H. M. et al. What has changed in the outbreaking populations of the severe crop pest whitefly species in cassava in two decades?. Sci. Rep. https://doi.org/10.1038/s41598-019-50259-0 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    59.Kalyebi, A. et al. Within-season changes in land use impact pest abundance in smallholder African cassava production systems. Insects (2021) (Revised Submitted).60.Kalyebi, A. et al. African cassava whitefly, Bemisia tabaci, cassava colonization preferences and control implications. PLoS ONE 13, e0204862. https://doi.org/10.1371/journal.pone.0204862 (2018).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    61.Macfadyen, S. et al. Landscape factors and how they influence whitefly pests in cassava fields across East Africa. Landsc. Ecol. 36, 45–67. https://doi.org/10.1007/s10980-020-01099-1 (2021).Article 

    Google Scholar 
    62.Tay, W. T. et al. A high-throughput amplicon sequencing approach for population-wide species diversity and composition survey. bioRxiv https://doi.org/10.1101/2020.10.12.336545 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    63.Manani, D. M., Ateka, E. M., Nyanjom, S. R. G. & Boykin, L. M. Phylogenetic relationships among whiteflies in the Bemisia tabaci(Gennadius) species complex from major cassava growing areas in Kenya. Insects https://doi.org/10.3390/insects8010025 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    64.Gottelli, D., Marino, J., Sillero-Zubiri, C. & Funk, S. M. The effect of the last glacial age on speciation and population genetic structure of the endangered Ethiopian wolf (Canis simensis). Mol. Ecol. 13, 2275–2286. https://doi.org/10.1111/j.1365-294X.2004.02226.x (2004).CAS 
    Article 
    PubMed 

    Google Scholar 
    65.Sezonlin, M. et al. Phylogeography and population genetics of the maize stalk borer Busseola fusca (Lepidoptera, Noctuidae) in sub-Saharan Africa. Mol. Ecol. 15, 407–420. https://doi.org/10.1111/j.1365-294X.2005.02761.x (2006).CAS 
    Article 
    PubMed 

    Google Scholar 
    66.Lehmann, T. et al. The rift valley complex as a barrier to gene flow for Anopheles gambiae in Kenya. J. Hered. 90, 613–621. https://doi.org/10.1093/jhered/90.6.613 (1999).CAS 
    Article 
    PubMed 

    Google Scholar 
    67.Schmidt, H. et al. Transcontinental dispersal of Anopheles gambiae occurred from West African origin via serial founder events. Commun. Biol. 2, 473. https://doi.org/10.1038/s42003-019-0717-7 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    68.Mairal, M. et al. Geographic barriers and Pleistocene climate change shaped patterns of genetic variation in the Eastern Afromontane biodiversity hotspot. Sci. Rep. 7, 45749. https://doi.org/10.1038/srep45749 (2017).CAS 
    Article 
    PubMed 
    PubMed Central 
    ADS 

    Google Scholar 
    69.Freilich, X. et al. Comparative Phylogeography of Ethiopian anurans: Impact of the Great Rift Valley and Pleistocene climate change. BMC Evol. Biol. 16, 206. https://doi.org/10.1186/s12862-016-0774-1 (2016).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    70.Huhndorf, M. H., Peterhans, J. C. K. & Loew, S. S. Comparative phylogeography of three endemic rodents from the Albertine Rift, east central Africa. Mol. Ecol. 16, 663–674. https://doi.org/10.1111/j.1365-294X.2007.03153.x (2007).Article 
    PubMed 

    Google Scholar 
    71.Matsubayashi, K. W., Ohshima, I. & Nosil, P. Ecological speciation in phytophagous insects. Entomol. Exp. Appl. 134, 1–27. https://doi.org/10.1111/j.1570-7458.2009.00916.x (2010).Article 

    Google Scholar 
    72.Malka, O. et al. Species-complex diversification and host-plant associations in Bemisia tabaci: A plant-defence, detoxification perspective revealed by RNA-Seq analyses. Mol. Ecol. 27, 4241–4256. https://doi.org/10.1111/mec.14865 (2018).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    73.Chen, W. B. et al. The draft genome of whitefly Bemisia tabaciMEAM1, a global crop pest, provides novel insights into virus transmission, host adaptation, and insecticide resistance. BMC Biol. https://doi.org/10.1186/s12915-016-0321-y (2016).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    74.Xie, W. et al. The invasive MED/Q Bemisia tabaci genome: A tale of gene loss and gene gain. BMC Genomics 19, 68. https://doi.org/10.1186/s12864-018-4448-9 (2018).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    75.Katoh, K. & Standley, D. M. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol. Biol. Evol. 30, 772–780. https://doi.org/10.1093/molbev/mst010 (2013).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    76.Bolger, A. M., Lohse, M. & Usadel, B. Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120. https://doi.org/10.1093/bioinformatics/btu170 (2014).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    77.Li, H. & Durbin, R. Fast and accurate long-read alignment with Burrows–Wheeler transform. Bioinformatics 26, 589–595. https://doi.org/10.1093/bioinformatics/btp698 (2010).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    78.Li, H. et al. The sequence alignment/map format and SAMtools. Bioinformatics 25, 2078–2079. https://doi.org/10.1093/bioinformatics/btp352 (2009).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    79.Eaton, D. A. R. PyRAD: Assembly of de novo RADseq loci for phylogenetic analyses. Bioinformatics 30, 1844–1849. https://doi.org/10.1093/bioinformatics/btu121 (2014).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    80.Zheng, X. W. et al. A high-performance computing toolset for relatedness and principal component analysis of SNP data. Bioinformatics 28, 3326–3328. https://doi.org/10.1093/bioinformatics/bts606 (2012).CAS 
    Article 
    PubMed 
    PubMed Central 

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

    Google Scholar 
    82.Stamatakis, A. RAxML-VI-HPC: Maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 22, 2688–2690. https://doi.org/10.1093/bioinformatics/btl446 (2006).CAS 
    Article 
    PubMed 

    Google Scholar 
    83.Pickrell, J. K. & Pritchard, J. K. Inference of population splits and mixtures from genome-wide allele frequency data. PLoS Genet. https://doi.org/10.1371/journal.pgen.1002967 (2012).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    84.Decker, J. E. et al. Worldwide patterns of ancestry, divergence, and admixture in domesticated cattle. PLoS Genet. 10, e1004254. https://doi.org/10.1371/journal.pgen.1004254 (2014).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    85.Gompert, Z. et al. Admixture and the organization of genetic diversity in a butterfly species complex revealed through common and rare genetic variants. Mol. Ecol. 23, 4555–4573. https://doi.org/10.1111/mec.12811 (2014).Article 
    PubMed 

    Google Scholar  More

  • in

    Late Pleistocene South American megafaunal extinctions associated with rise of Fishtail points and human population

    1.Barnosky, A. D. & Lindsey, E. L. Timing of Quaternary megafaunal extinction in South America in relation to human arrival and climate change. Quat. Int. 217, 10–29 (2010).Article 

    Google Scholar 
    2.Broughton, J. M. & Weitzel, E. M. Population reconstructions for humans and megafauna suggest mixed causes for North American Pleistocene extinctions. Nat. Commun. 9, 1–12 (2018).CAS 
    Article 

    Google Scholar 
    3.Haynes, G. in The Encyclopedia of the Anthropocene 1 (eds. DellaSala, D. A., & Goldstein, M. I.) 219–226 (Springer, 2018).4.Wolfe, A. L. & Broughton, J. M. A foraging theory perspective on the associational critique of North American Pleistocene overkill. J. Archaeol. Sci. 119, 105162 (2020).Article 

    Google Scholar 
    5.Rothschild, B. M. & Laub, R. Hyperdisease in the late Pleistocene: validation of an early 20th century hypothesis. Naturwissenschaften 93, 557–564 (2006).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    6.Borrero, L. A. in American megafaunal extinctions at the end of the Pleistocene (ed. Haynes, G.) 145–168 (Springer, 2009).7.Cione, A. L., Tonni, E. P., & Soibelzon, L. in American megafaunal extinctions at the end of the Pleistocene (ed. Haynes, G.) 125–144 (Springer, 2009).8.Lima-Ribeiro, M. S., Nogués-Bravo, D., Terribile, L. C., Batra, P. & Diniz-Filho, J. A. F. Climate and humans set the place and time of Proboscidean extinction in late Quaternary of South America. Palaeogeogr. Palaeoclimatol. Palaeoecol. 392, 546–556 (2013).Article 

    Google Scholar 
    9.Grayson, D. K. & Meltzer, D. J. A requiem for North American overkill. J. Archaeol. Sci. 30, 585–593 (2003).Article 

    Google Scholar 
    10.Firestone, R. B. et al. Evidence for an extraterrestrial impact 12,900 years ago that contributed to the megafaunal extinctions and the Younger Dryas cooling. PNAS 104, 16016–16021 (2007).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    11.Waters, M. R. & Stafford, T. W. in Paleoamerican Odyssey (eds. Graf, K. E., Ketron, C. V. & Waters, M.) 543–562 (Texas A&M University Press, 2013).12.Politis, G., Prado, J. L., & Beukens, R. P. in Ancient Peoples and Landscapes (ed. Johnson, E.) 187–205 (Tech University Press, 1995).13.Martin, P. S. The Discovery of America: The first Americans may have swept the Western Hemisphere and decimated its fauna within 1000 years. Science 179, 969–974 (1973).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    14.Fiedel, S. J. in Paleoamerican origins: beyond Clovis (eds. Bonnichsen, R., Lepper, B. T., Stanford, D. & Waters M. R.) 97–102 (Texas A&M University Press, 2005).15.Surovell, T. A., Pelton, S. R., Anderson-Sprecher, R. & Myers, A. D. Test of Martin’s overkill hypothesis using radiocarbon dates on extinct megafauna. PNAS 113, 886–891 (2016).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    16.Pedro, J. B. et al. Southern Ocean deep convection as a driver of Antarctic warming events. Geophys. Res. Lett. 43, 2192–2199 (2016).ADS 
    Article 

    Google Scholar 
    17.Politis, G. in Clovis. Origins and Adaptations (eds Bonnichsen, R. & Turnmire, K.) 287–301 (Texas A&M University Press, 1991).18.Nami, H. G. Fishtailed projectile points in the Americas: Remarks and hypotheses on the peopling of northern South America and beyond. Quat. Int., in press (2020).19.Waters, M. R., Amorosi, T. & Stafford, T. W. Redating Fell’s cave, Chile and the chronological placement of the Fishtail projectile point. Am. Antiq. 80, 376–386 (2015).Article 

    Google Scholar 
    20.Loponte, D., Carbonera, M. & Silvestre, R. Fishtail projectile points from South America: the Brazilian record. Archaeol. Discov. 3, 85 (2015).Article 

    Google Scholar 
    21.Nami, H. G. & Yataco Capcha, J. Further Data on Fell Points from the Southern Cone of South America. PaleoAmerica 6, 379–386 (2020).Article 

    Google Scholar 
    22.Weitzel, C., Mazzia, N. & Flegenheimer, N. Assessing Fishtail points distribution in the southern Cone. Quat. Int. 473, 161–172 (2018).Article 

    Google Scholar 
    23.Martínez, G., Gutiérrez, M. A. & Tonni, E. P. Paleoenvironments and faunal extinctions: analysis of the archaeological assemblages at the Paso Otero locality (Argentina) during the Late Pleistocene–Early Holocene. Quat. Int 299, 53–63 (2013).Article 

    Google Scholar 
    24.Prates, L., Politis, G. G. & Perez, S. I. Rapid radiation of humans in South America after the last glacial maximum: a radiocarbon-based study. PLoS ONE 15, e0236023 (2020).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    25.Barnosky, A. D., Koch, P. L., Feranec, R. S., Wing, S. L. & Shabel, A. B. Assessing the causes of late Pleistocene extinctions on the continents. Science 306, 70–75 (2004).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    26.Buchanan, B., Collard, M., Hamilton, M. J. & O’Brien, M. J. Points and prey: a quantitative test of the hypothesis that prey size influences early Paleoindian projectile point form. J. Archaeol. Sci. 38, 852–864 (2011).Article 

    Google Scholar 
    27.Timpson, A. et al. Reconstructing regional population fluctuations in the European Neolithic using radiocarbon dates: a new case-study using an improved method. J. Archaeol. Sci. 52, 549–557 (2014).Article 

    Google Scholar 
    28.Crema, E. R., Habu, J., Kobayashi, K. & Madella, M. Summed probability distribution of 14C dates suggests regional divergences in the population dynamics of the Jomon period in eastern Japan. PLoS ONE 11, e0154809 (2016).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    29.Phillips, S. J., Anderson, R. P. & Schapire, R. E. Maximum entropy modeling of species geographic distributions. Ecol. Model. 190, 231–259 (2006).Article 

    Google Scholar 
    30.Elith, J. & Leathwick, J. R. Species distribution models: ecological explanation and prediction across space and time. Ann. Rev. Ecol. Evol. Syst. 40, 677–697 (2009).Article 

    Google Scholar 
    31.d’Amen, M. et al. Using species richness and functional traits predictions to constrain assemblage predictions from stacked species distribution models. J. Biogeog. 42, 1255–1266 (2015).Article 

    Google Scholar 
    32.McCulloch, R. D. et al. Climatic inferences from glacial and palaeoecological evidence at the last glacial termination, southern South America. J. Quat. Sci.: Published Quat. Res. Assoc. 15, 409–417 (2000).Article 

    Google Scholar 
    33.Moreno, P. I., Villa-Martínez, R., Cárdenas, M. L. & Sagredo, E. A. Deglacial changes of the southern margin of the southern westerly winds revealed by terrestrial records from SW Patagonia (52 S). Quat. Sci. Rev. 41, 1–21 (2012).ADS 
    Article 

    Google Scholar 
    34.Barnosky, A. D. et al. Variable impact of late-Quaternary megafaunal extinction in causing ecological state shifts in North and South America. PNAS 113, 856–861 (2016).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    35.Politis, G. & Messineo, P. The Campo Laborde site: New evidence for the Holocene survival of Pleistocene megafauna in the Argentine Pampas. Quat. Int. 191, 98–114 (2008).Article 

    Google Scholar 
    36.Politis, G. G., Messineo, P. G., Stafford, T. W. & Lindsey, E. L. Campo Laborde: a Late Pleistocene giant ground sloth kill and butchering site in the Pampas. Sci. Adv. 5, eaau4546 (2019).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    37.Phillips, S. J., Anderson, R. P., Dudík, M., Schapire, R. E. & Blair, M. E. Opening the black box: an open‐source release of Maxent. Ecography 40, 887–893 (2017).Article 

    Google Scholar 
    38.Fordham, D. A. et al. PaleoView: a tool for generating continuous climate projections spanning the last 21 000 years at regional and global scales. Ecography 40, 1348–1358 (2017).Article 

    Google Scholar 
    39.Brown, J. L., Hill, D. J., Dolan, A. M., Carnaval, A. C. & Haywood, A. M. PaleoClim, high spatial resolution paleoclimate surfaces for global land areas. Sci. Data 5, 1–9 (2018).ADS 
    Article 
    CAS 

    Google Scholar 
    40.QGIS Development Team. QGIS Geographic Information System. Open Source Geospatial Foundation Project. http://qgis.osgeo.org (2020).41.Warren, D. L., Glor, R. E. & Turelli, M. Environmental niche equivalency versus conservatism: quantitative approaches to niche evolution. Evolution 62, 2868–2883 (2008).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    42.Tonni, E. P., Carlini, A. A., Yané, G. J. S. & Figini, A. J. Cronología radiocarbónica y condiciones climáticas en la “Cueva del Milodón” (sur de Chile) durante el Pleistoceno Tardío. Ameghiniana 40, 609–615 (2003).
    Google Scholar 
    43.Tonni, E. P. & Carlini, A. A. Neogene vertebrates from Argentine Patagonia: their relationship with the most significant climatic changes. Dev. Quat. Sci. 11, 269–283 (2008).
    Google Scholar 
    44.Schmitt, S., Pouteau, R., Justeau, D., de Boissieu, F. & Birnbaum, P. ssdm: an r package to predict distribution of species richness and composition based on stacked species distribution models. Methods Ecol. Evol. 8, 1795–1803 (2017).Article 

    Google Scholar 
    45.Mothé, D. et al. An artifact embedded in an extinct proboscidean sheds new light on human-megafaunal interactions in the quaternary of South America. Quat. Sci. Rev. 229, 106125 (2020).Article 

    Google Scholar 
    46.Jaimes, A. Condiciones tafonómicas, huesos modificados y comportamiento humano en los sitios de matanza de El Vano (Tradición El Jobo) y Lange/Ferguson (Tradición Clovis). Bol. Antropol. Am. 41, 159–184 (2005).
    Google Scholar 
    47.Moreno, P. I. et al. Renewed glacial activity during the Antarctic Cold reversal and persistence of cold conditions until 11.5 ka in SW Patagonia. Geology 37, 375–378 (2009).ADS 
    CAS 
    Article 

    Google Scholar 
    48.Obase, T. & Abe‐Ouchi, A. Abrupt Bølling‐Allerød warming simulated under gradual forcing of the last deglaciation. Geophys. Res. Lett. 46, 11397–11405 (2019).ADS 
    Article 

    Google Scholar 
    49.de Porras, M. E. et al. Environmental and climatic changes in central Chilean Patagonia since the late glacial (Mallín El Embudo, 44 S). Clim 10, 1063–1078 (2014).ADS 

    Google Scholar 
    50.Mendelová, M., Hein, A. S., Rodes, A., Smedley, R. K. & Xu, S. Glacier expansion in central Patagonia during the Antarctic Cold Reversal followed by retreat and stabilisation during the Younger Dryas. Quat. Sci. Rev. 227, 106047 (2020).Article 

    Google Scholar 
    51.Villavicencio, N. A. et al. Combination of humans, climate, and vegetation change triggered Late Quaternary megafauna extinction in the Última Esperanza region, southern Patagonia, Chile. Ecography 39, 125–140 (2016).Article 

    Google Scholar 
    52.Prieto, A. R. Vegetational History of the Late glacial-holocene transition in the grasslands of Eastern Argentina. Palaeogr. Palaeoclimatol. Palaeoecol. 157, 167–188 (2000).ADS 
    Article 

    Google Scholar 
    53.Miotti, L., Tonni, E. & Marchionni, L. What happened when the Pleistocene megafauna became extinct? Quat. Int. 473, 173–189 (2018).Article 

    Google Scholar 
    54.Méndez, C. et al. J. L. Human effects in Holocene fire dynamics of central Western Patagonia (~ 44° S, Chile). Front. Ecol. Evol. 4, 100 (2016).Article 

    Google Scholar 
    55.Pires, M. M. et al. Pleistocene megafaunal interaction networks became more vulnerable after human arrival. Proc. R. Soc. Lon. [Biol.] 282, 20151367 (2015).
    Google Scholar 
    56.Pires, M. et al. Before, during and after megafaunal extinctions: human impact on Pleistocene-Holocene trophic networks in South Patagonia. Quat. Sci. Rev. 250, 106296 (2020).Article 

    Google Scholar 
    57.Haynes, G. & Klimowicz, J. Recent elephant-carcass utilization as a basis for interpreting mammoth exploitation. Quat. Int. 359, 19–37 (2015).Article 

    Google Scholar 
    58.Surovell, T. A. & Grund, B. S. The associational critique of quaternary overkill and why it is largely irrelevant to the extinction debate. Am. Antiq. 77, 672–687 (2012).Article 

    Google Scholar 
    59.Rindel, D. D., Moscardi, B. F., & Perez, S. I. The distribution of the guanaco (Lama guanicoe) in Patagonia during Late Pleistocene–Holocene and its importance for prehistoric human diet. Holocene, https://doi.org/10.1177/0959683620981689 (2020).60.Metcalf, J. L. et al. Synergistic roles of climate warming and human occupation in Patagonian megafaunal extinctions during the Last Deglaciation. Sci. Adv. 2, e1501682 (2016).ADS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    61.Villavicencio, N. A., Corcoran, D. & Marquet, P. A. Assessing the causes behind the late quaternary extinction of Hhorses in South america using species distribution models. Fron. Ecol. Evol. 7, 226 (2019).Article 

    Google Scholar 
    62.Prado, J. L., & Alberdi, M. T. Fossil Horses of South America. Phylogeny, Systemics and Ecology. (Springer, 2017).63.Varela, L. & Fariña, R. A. Co-occurrence of mylodontid sloths and insights on their potential distributions during the late Pleistocene. Quat. Res. 85, 66–74 (2016).Article 

    Google Scholar 
    64.R-Development Core Team R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. http://www.R-project.org/ (2020).65.Menegaz, A. N., Goin, F. J. & Jaureguizar, E. O. Análisis morfológico y morfométrico multivariado de los representantes fósiles y vivientes del género Lama (Artiodactyla, Camelidae). Sus implicancias sistemáticas, biogeográficas, ecológicas y biocronológicas. Ameghiniana 26, 153–172 (1989).
    Google Scholar 
    66.Scherer, C. S. Os Camelidae Lamini (Mammalia, Artiodactyla) do plesistoceno da América do Sul: aspectos taxonômicos e filogenéticos. Unpublished PhD Thesis, Universidade Federal do Rio Grande do Sul. (2009).67.Weinstock, J. et al. The Late Pleistocene distribution of vicuñas (Vicugna vicugna) and the “extinction” of the gracile llama (“Lama gracilis”): new molecular data. Quat. Sci. Rev. 28, 1369–1373 (2009).ADS 
    Article 

    Google Scholar 
    68.Mothé, D., Avilla, L. S. & Cozzuol, M. A. The south American gomphotheres (Mammalia, Proboscidea, Gomphotheriidae): taxonomy, phylogeny, and biogeography. J. Mamm. Evol. 20, 23–32 (2013).Article 

    Google Scholar 
    69.Prado, J. L. & Alberdi, M. T. Global evolution of equidae and gomphotheriidae from South America. Integ. Zool. 9, 434–443 (2014).Article 

    Google Scholar  More

  • in

    Estimation of tuna population by the improved analytical pipeline of unique molecular identifier-assisted HaCeD-Seq (haplotype count from eDNA)

    1.Garlapati, D., Charankumar, B., Ramu, K., Madeswaran, P. & Ramana Murthy, M. V. A review on the applications and recent advances in environmental DNA (eDNA) metagenomics. Rev. Environ. Sci. Biotechnol. 18, 389–411 (2019).CAS 
    Article 

    Google Scholar 
    2.Rees, H. C., Maddison, B. C., Middleditch, D. J., Patmore, J. R. M. & Gough, K. C. REVIEW: The detection of aquatic animal species using environmental DNA—a review of eDNA as a survey tool in ecology. J. Appl. Ecol. 51, 1450–1459 (2014).CAS 
    Article 

    Google Scholar 
    3.Ruppert, K. M., Kline, R. J. & Rahman, M. S. Past, present, and future perspectives of environmental DNA (eDNA) metabarcoding: a systematic review in methods, monitoring, and applications of global eDNA. Glob. Ecol. Conserv. 17, e00547 (2019).Article 

    Google Scholar 
    4.Thomsen, P. F. & Willerslev, E. Environmental DNA—an emerging tool in conservation for monitoring past and present biodiversity. Biol. Conserv. 183, 4–18 (2015).Article 

    Google Scholar 
    5.Knudsen, S. W. et al. Species-specific detection and quantification of environmental DNA from marine fishes in the Baltic Sea. J. Exp. Mar. Biol. Ecol. 510, 31–45 (2019).CAS 
    Article 

    Google Scholar 
    6.Salter, I., Joensen, M., Kristiansen, R., Steingrund, P. & Vestergaard, P. Environmental DNA concentrations are correlated with regional biomass of Atlantic cod in oceanic waters. Commun. Biol. 2, 1–9 (2019).Article 
    CAS 

    Google Scholar 
    7.Stoeckle, M. Y. et al. Trawl and eDNA assessment of marine fish diversity, seasonality, and relative abundance in coastal New Jersey, USA. ICES J. Mar. Sci. https://doi.org/10.1093/icesjms/fsaa225 (2020).Article 

    Google Scholar 
    8.Maruyama, A., Nakamura, K., Yamanaka, H., Kondoh, M. & Minamoto, T. The release rate of environmental DNA from juvenile and adult fish. PLoS ONE 9, e114639 (2014).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    9.Jane, S. F. et al. Distance, flow and PCR inhibition: eDNA dynamics in two headwater streams. Mol. Ecol. Resour. 15, 216–227 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    10.Yoshitake, K. et al. HaCeD-Seq: a novel method for reliable and easy estimation about the fish population using haplotype count from eDNA. Mar. Biotechnol. N. Y. 21, 813–820 (2019).CAS 
    Article 

    Google Scholar 
    11.Kinde, I., Wu, J., Papadopoulos, N., Kinzler, K. W. & Vogelstein, B. Detection and quantification of rare mutations with massively parallel sequencing. Proc. Natl. Acad. Sci. 108, 9530–9535 (2011).ADS 
    PubMed 
    Article 

    Google Scholar 
    12.Casbon, J. A., Osborne, R. J., Brenner, S. & Lichtenstein, C. P. A method for counting PCR template molecules with application to next-generation sequencing. Nucleic Acids Res. 39, e81–e81 (2011).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    13.Kivioja, T. et al. Counting absolute numbers of molecules using unique molecular identifiers. Nat. Methods 9, 72–74 (2012).CAS 
    Article 

    Google Scholar 
    14.Vander Heiden, J. A. et al. pRESTO: a toolkit for processing high-throughput sequencing raw reads of lymphocyte receptor repertoires. Bioinformatics 30, 1930–1932 (2014).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    15.Shugay, M. et al. MAGERI: Computational pipeline for molecular-barcoded targeted resequencing. PLOS Comput. Biol. 13, e1005480 (2017).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    16.Clement, K., Farouni, R., Bauer, D. E. & Pinello, L. AmpUMI: design and analysis of unique molecular identifiers for deep amplicon sequencing. Bioinformatics 34, i202–i210 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    17.Kumar, G., Kocour, M. & Kunal, S. P. Mitochondrial DNA variation and phylogenetic relationships among five tuna species based on sequencing of D-loop region. Mitochondrial DNA Part A 27, 1976–1980 (2016).CAS 
    Article 

    Google Scholar 
    18.Nomura, S. et al. Genetic population structure of the Pacific bluefin tuna Thunnus orientalis and the yellowfin tuna Thunnus albacares in the North Pacific Ocean. Fish. Sci. 80, 1193–1204 (2014).CAS 
    Article 

    Google Scholar 
    19.Gleiss, A. C., Schallert, R. J., Dale, J. J., Wilson, S. G. & Block, B. A. Direct measurement of swimming and diving kinematics of giant Atlantic bluefin tuna (Thunnus thynnus). R. Soc. Open Sci. 6, 190203 (2019).ADS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    20.Juan-Jordá, M. J., Mosqueira, I., Freire, J. & Dulvy, N. K. The conservation and management of tunas and their relatives: setting life history research priorities. PLoS ONE 8, e70405 (2013).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    21.Shibata, M. et al. Transcriptomic features associated with energy production in the muscles of Pacific bluefin tuna and Pacific cod. Biosci. Biotechnol. Biochem. 80, 1114–1124 (2016).CAS 
    PubMed 
    Article 

    Google Scholar 
    22.Swanson, D., Block, R. & Mousa, S. A. Omega-3 fatty acids EPA and DHA: health benefits throughout life. Adv. Nutr. Bethesda Md 3, 1–7 (2012).CAS 
    Article 

    Google Scholar 
    23.Collette, B. B. et al. Conservation. High value and long life–double jeopardy for tunas and billfishes. Science 333, 291–292 (2011).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    24.Kumai, H. & Miyashita, S. Life cycle of the Pacific bluefin tuna is completed under reared condition. Nippon Suisan Gakkaishi Jpn. 69, 124–127 (2003).Article 

    Google Scholar 
    25.Miyashita, S. et al. Maturation and spawning of cultured bluefin tuna, Thunnus thynnus. Suisanzoushoku Jpn. 48, 475–488 (2000).
    Google Scholar 
    26.Cho, J. et al. Production performance of Pacific bluefin tuna Thunnus orientalis larvae and juveniles fed commercial diets and effects of switching diets. Aquac. Sci. 64, 359–370 (2016).CAS 

    Google Scholar 
    27.Tsuji, S. et al. Environmental DNA analysis shows high potential as a tool for estimating intraspecific genetic diversity in a wild fish population. Mol. Ecol. Resour. https://doi.org/10.1111/1755-0998.13165 (2020).Article 
    PubMed 

    Google Scholar 
    28.Tsuji, S. et al. Evaluating intraspecific genetic diversity using environmental DNA and denoising approach: a case study using tank water. Environ. DNA 2, 42–52 (2020).Article 

    Google Scholar 
    29.Ppyun, H. et al. Improved PCR performance and fidelity of double mutant Neq A523R/N540R DNA polymerase. Enzym. Microb. Technol. 82, 197–204 (2016).CAS 
    Article 

    Google Scholar 
    30.Hsieh, T. C., Ma, K. H. & Chao, A. iNEXT: an R package for rarefaction and extrapolation of species diversity (Hill numbers). Methods Ecol. Evol. 7, 1451–1456 (2016).Article 

    Google Scholar 
    31.Myers, R. A. & Worm, B. Rapid worldwide depletion of predatory fish communities. Nature 423, 280–283 (2003).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    32.Worm, B. et al. Rebuilding global fisheries. Science 325, 578–585 (2009).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    33.Bravington, M., Grewe, P. & Davies, C. Fishery-independent estimate of spawning biomass of southern bluefin tuna through identification of close-kin using genetic markers. FRDC Report2007034CSIRO Aust. (2014).34.Bravington, M. V., Grewe, P. M. & Davies, C. R. Absolute abundance of southern bluefin tuna estimated by close-kin mark-recapture. Nat. Commun. 7, 13162 (2016).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    35.Song, N., Jia, N., Yanagimoto, T., Lin, L. & Gao, T. Genetic differentiation of Trachurus japonicus from the Northwestern Pacific based on the mitochondrial DNA control region. Mitochondrial DNA 24, 705–712 (2013).CAS 
    PubMed 
    Article 

    Google Scholar 
    36.Zhu, Y., Cheng, Q. & Rogers, S. M. Genetic structure of Scomber japonicus (Perciformes: Scombridae) along the coast of China revealed by complete mitochondrial cytochrome b sequences. Mitochondrial DNA Part DNA Mapp. Seq. Anal. 27, 3828–3836 (2016).CAS 
    Article 

    Google Scholar 
    37.Tzeng, T.-D. Population structure and historical demography of the spotted mackerel (Scomber australasicus) off Taiwan inferred from mitochondrial control region sequencing. Zool. Stud. 8, 656–663 (2007).
    Google Scholar 
    38.Ichinokawa, M., Okamura, H. & Kurota, H. The status of Japanese fisheries relative to fisheries around the world. ICES J. Mar. Sci. 74, 1277–1287 (2017).Article 

    Google Scholar 
    39.Pikitch, E. K. et al. Ecosystem-based fishery management. Science 305, 346–347 (2004).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    40.Yates, M. C., Fraser, D. J. & Derry, A. M. Meta-analysis supports further refinement of eDNA for monitoring aquatic species-specific abundance in nature. Environ. DNA 1, 5–13 (2019).Article 

    Google Scholar 
    41.Lacoursière Roussel, A., Rosabal, M. & Bernatchez, L. Estimating fish abundance and biomass from eDNA concentrations: variability among capture methods and environmental conditions. Mol. Ecol. Resour. 16, 1401–1414 (2016).PubMed 
    Article 
    CAS 

    Google Scholar 
    42.Yamamoto, S. et al. Environmental DNA as a ‘Snapshot’ of fish distribution: a case study of Japanese jack mackerel in Maizuru bay, Sea of Japan. PLoS ONE 11, e0149786 (2016).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    43.Untergasser, A. et al. Primer3–new capabilities and interfaces. Nucleic Acids Res. 40, e115 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    44.Sekino, M. & Yamashita, H. Mitochondrial DNA barcoding for Okinawan oysters: a cryptic population of the Portuguese oyster Crassostrea angulata in Japanese waters. Fish. Sci. 79, 61–76 (2013).CAS 
    Article 

    Google Scholar 
    45.Tamura, K., Stecher, G., Peterson, D., Filipski, A. & Kumar, S. MEGA6: molecular evolutionary genetics analysis version 6.0. Mol. Biol. Evol. 30, 2725–2729 (2013).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    46.Minamoto, T., Naka, T., Moji, K. & Maruyama, A. Techniques for the practical collection of environmental DNA: filter selection, preservation, and extraction. Limnology 17, 23–32 (2016).CAS 
    Article 

    Google Scholar 
    47.Miya, M. et al. MiFish, a set of universal PCR primers for metabarcoding environmental DNA from fishes: detection of more than 230 subtropical marine species. R. Soc. Open Sci. 2, 150088 (2015).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    48.Magoč, T. & Salzberg, S. L. FLASH: fast length adjustment of short reads to improve genome assemblies. Bioinformatics 27, 2957–2963 (2011).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar  More

  • in

    Effects of temperature and phytoplankton community composition on subitaneous and resting egg production rates of Acartia omorii in Tokyo Bay

    Population dynamicsThe abundance of A. omorii peaked in April 2016, March 2017, and April–May 2018, ranging from 2.99 × 104 to 6.73 × 104 individuals m−3 (Fig. 3, Table 1). Anakubo and Murano28 reported that the abundance of A. omorii, including individuals from the nauplius to adult stage, peaked at 5.34 × 104 m−3 in April 1981 in Tokyo Bay. Itoh and Aoki18 surveyed A. omorii in Tokyo Bay from March 1990 to November 1992 and found that abundance peaked at 2.13 × 104 individuals m−3 in March. Tachibana et al.40 reported that abundance peaked at 2.60 × 104 individuals m−3 in April in 2007. The peak abundance in Tokyo Bay apparently occurs between March and May; this pattern was also observed in the Seto Inland Sea (Table 1). According to Kasahara et al.15, the peak period of abundance of adults and later-stage copepodites in the Seto Inland Sea occurred in mid-May at  > 3.5 × 104 individuals m−3. The abundance of adults and later-stage copepodites significantly decreased in July. The planktonic population disappeared from the water column in mid-August and recovered in November, when the water temperature was  50 g wet weight m−3). Hence summer jellyfish bloom probably has a strong negative impact on the abundance of A. omorii. Unfortunately, we do not have enough information on jellyfish abundance in Onagawa Bay, effect of jellyfish should be considered for the better understanding of A. omorii dynamics in future.A. omorii abundance became zero only from September to November 2016 in the 3 years of the present study. However, in other years, A. omorii appeared at low abundance in summer, as reported in previous studies conducted in Tokyo Bay 16,18,28,40. The period of complete disappearance was short, and low densities were recorded throughout the year in most studies (Table 1), indicating that the population is maintained in the planktonic stages even under unfavorable conditions in Tokyo Bay.EPR and egg type dynamicsSEPR peaked in winter (January or February) and May in both years at both stations (Fig. 8a). The second peaks were caused by increase in unhatched egg production, except for station CB in 2017 (Figs. 7b,c and 8b,c). Increase in unhatched egg production occurred when the surface water temperature exceeded 18 °C (Fig. 2a) and day length increased to  > 14 h, similar to that reported by Uye5. From January to May 1982, 80% of eggs that were newly spawned by A. omorii females collected in Fukuyama Harbor of the Seto Inland Sea were subitaneous; resting eggs appeared in June when the surface water temperature exceeded 17.5 °C. In the present study, on June 9, 32% of the total eggs produced were resting and did not hatch within 14 days but hatched after being reincubated at 15 °C for 2 weeks. Uye5 also demonstrated the effect of photoperiod on egg production by A. omorii via laboratory experiments under two temperature conditions (15 °C and 20 °C ± 1 °C). Approximately half were resting eggs under l4L–10D photoperiodic conditions at both water temperatures, indicating that photoperiod is important for shift to resting eggs. Many copepods have dormancy strategies at the thermal limits of species distribution46. In Tokyo Bay, the photoperiod exceeds 14 h in mid-May, when the water temperature usually exceeds 18 °C (Fig. 2a). Therefore, higher water temperatures and increased day length periods are synchronized in early summer in Tokyo Bay and may serve as cues for diapause egg production.Unhatched PEPR exceeded subitaneous PEPR from May to June (Fig. 9), when abundance drastically decreased to  45%) (Fig. 6) in the late period of A. omorii appearance (in May at station CB and in early June at station F3). However, in early June at station CB and late June at station F3, when the abundance sharply decreased, the proportion of females producing subitaneous eggs increased, whereas those producing quiescent eggs decreased from the previous month (Fig. 11). Uye5 reported a similar result; the proportion of diapause eggs to total eggs produced by A. omorii in the Seto Inland Sea peaked on June 9 and then reduced by half by June 30. Quiescent eggs have been defined as subitaneous eggs with arrested development that remain in a quiescent stage in unsuitable environmental conditions49. Uye5 defined diapause eggs as eggs that did not hatch during 2 weeks at in situ water temperatures but hatched within 2 weeks when the incubation temperature decreased to 15 °C. These eggs may be classified as quiescent eggs in the present study. The results of Uye5 and the present study suggest that not all produced eggs of this species shift from subitaneous to quiescent eggs at higher water temperatures.As mentioned in the previous subsection, a planktonic population of A. omorii has been found in mid-summer at low abundance in Tokyo Bay16,18,28,40 (Fig. 3). Similar results were reported in Maizuru Bay50. Itoh et al.45 investigated the vertical distribution of copepods at 1-m depth intervals at station F3 in Tokyo Bay in mid-summer, when hypoxia develops near the bottom, and showed that A. omorii population had a sharp peak, with densities exceeding 1.5 × 103 individuals m−3, at 8 m at a water temperature of 18 °C and just above the hypoxic zone45. This suggests that A. omorii maintains a planktonic stage even at low density in mid-summer, whereas most of the population estivates by forming resting eggs in bottom sediments. These mid-summer populations are presumably hatched from subitaneous eggs spawned in mid-July (Figs. 7, 10). Uye5 also reported that more than half of the eggs were still subitaneous in late July in the Seto Inland Sea. Therefore, we tentatively think that this phenomenon is a bet-hedging strategy of A. omorii in an unfavorable and uncertain environment. In contrast, Ueda50 stated that the increase in subitaneous EPR in summer was due to immature development of this species. Thus, these remaining populations may not contribute to the autumnal development of the population. To understand how A. omorii survive in mid-summer, more detailed field investigations are warranted, including egg and nauplii dynamics in the water column, egg hatching process from sediments, and differences in endogenous factors in individual females producing subitaneous and diapause eggs in summer.Information on A. omorii’s delayed-hatching eggs is strictly limited. Delayed-hatching eggs are eggs hatching over a wide time span regardless of environmental conditions4,11. Takayama and Toda4 defined the unhatched eggs of A. japonica hatching during 72 h–50 days as “delayed-hatching eggs.” Thus, delayed-hatching eggs may have been included in the quiescent and diapause eggs defined in this study. Our results showed that no eggs hatched between 48–96 h and 7 days in the experiment at in situ water temperature; many quiescent eggs hatched within a few days after reincubation at lower water temperatures (Figs. 10, 11). Therefore, delayed-hatching eggs may not have been produced in the present study.Effects of water temperature on the production of subitaneous and diapause eggsMultiple regression analysis revealed that subitaneous SEPR negatively correlated with bottom water temperature, inconsistent with the results of Uye3. EPR and copepod growth generally increase with increased water temperature51. Uye3 reported that EPR of A. omorii also increased with water temperature; they developed a simple model equation describing the fecundity of A. omorii in Onagawa Bay via a laboratory experiment:$${text{F}} = 0.000{331 }left( {{text{T}} + {12}.0} right)^{{{3}.{25}}} {text{SW}}_{{text{f}}} /left( {0.{47}0 + {text{S}}} right),$$
    where F is daily fecundity (eggs female−1 day−1), T is water temperature (°C), S is chlorophyll a concentration (µg L−1), and Wf is female carbon content (µg). The fecundity predicted by the above described model was similar to the observed EPR of this species in Onagawa Bay3.Many studies have used Uye’s equation to estimate A. omorii egg production. Kang et al.52 reported A. omorii’s EPR in Ilkwang Bay to be 22–57 eggs female−1 day−1, which was higher than that in the present study (1.6–18.7 eggs female−1 day−1) (Fig. 7a). Liang and Uye17 estimated A. omorii’s EPR in the Seto Inland Sea by two methods: the above described model (estimated incubation fecundity)3 and based on the number of eggs remaining in the water column and the adult female population (egg-ratio fecundity)53. In the Seto Inland Sea, the estimated incubation fecundity was 26–60 eggs female−1 day−1 and the egg-ratio fecundity was 0.5–25 eggs female−1 day−1; the estimated incubation fecundity was always greater than the egg-ratio fecundity17.Suspension-feeding copepods may ingest their own eggs and nauplii. In the Seto Inland Sea, possible egg predators were the dominant copepods Centropages abdominalis and A. omorii54. Based on their abundance (0.2–39 predators L−1) and assuming a predator clearance rate of 50 mL d−1, C. abdominalis and A. omorii could remove 1–86% of eggs in the water column per day. Liang and Uye17 noted that their predators were abundant when the abundance of surviving eggs in the water column was low; therefore, they tentatively concluded that the difference between the two estimates was due to egg loss by predation, including cannibalism. However, it is unlikely that fecundity reached its highest value ( > 50 eggs female−1 day−1) in mid-July when the population disappeared from the water column17. At that time, the water temperature was 25 °C, which also does not support the increase in fecundity observed by Liang and Uye17.The model equation of Uye3 was derived from Onagawa Bay, where the average water temperature is 7.7–21.9°C55. In laboratory experiments using A. omorii from Onagawa Bay, EPR decreased when the water temperature exceeded 22.5°C3. In Uye’s equation3, the decrease in egg production above 22.5 °C was not foreseen, whereas water temperature exceeded 22.5 °C in the Seto Inland Sea17, Ikkwang Bay52, and Tokyo Bay (Fig. 2). Thus, Uye’s model equation3 is not applicable to these warm environments.Based on the temperature regime, seasonal population dynamics and egg types produced are divided into two types: no resting egg production in the colder Onagawa Bay and resting egg production in the warmer Tokyo Bay and Seto Inland Sea. As mentioned above, A. omorii in Onagawa Bay exists throughout the year, even in summer14 and hardly produces diapause eggs5,7. However, the population almost disappears in late summer in Tokyo Bay16,18,28,40 and the Seto Inland Sea15,17. Furthermore, in these warm coastal waters, A. omorii produced diapause eggs just before copepodite disappearance from the water column. Therefore, a separate equation for estimating egg production should be developed, depending on the temperature regime of the habitat.Recent climate change, particularly global warming, may affect A. omorii’s egg production. In Tokyo Bay, between 1955 and 2015, the water temperature increased by 1.0 °C and 0.94 °C at the surface and bottom layers, respectively, in winter and autumn56,57. In summer, the water temperature at both the surface and bottom layers decreased, probably due to strengthened estuary circulation56,57. Considering the response of A. omorii to water temperature, the increase in winter temperature might reduce subitaneous egg production, resulting in delayed population increase. In contrast, the decrease in summer temperature might lead to reduced diapause egg production per amount of body carbon. The long-term trends of water temperature might have different effects on each egg type production and alter the dynamics of A. omorii egg production.Effects of phytoplankton composition on the production of subitaneous and diapause eggsThe EPR of A. omorii may be saturated at low (1–2 µg L−1) chlorophyll a concentrations3,19. However, multiple regression analysis revealed that small diatoms stimulate subitaneous SEPR (Figs. 8, 12). The EPR at station CB was quite high ( > 14 eggs female−1 day−1) in January and February 2018, when the diatoms comprised Skeletonema and Chaetoceros. In contrast, at station F3, EPR drastically decreased from 18.7 ± 6.3 eggs female−1 day−1 in January to 8.4 ± 4.6 eggs female−1 day−1 in February 2018. The EPR at station F3 in February was significantly lower than that at station CB (Tukey’s post hoc test, p  95%) at station CB in January and February 2018, suggesting that small diatoms ingestion enhances A. omorii’s egg production.It is also likely that small nanoflagellates have an adverse effect on egg production. At station F3, the proportion of nanoflagellates to total phytoplankton carbon biomass was high ( > 93%) in February and March (Fig. 12), when EPR was quite low ( More

  • in

    Reproductive performance in houbara bustard is affected by the combined effects of age, inbreeding and number of generations in captivity

    1.Conde, D. A., Flesness, N., Colchero, F., Jones, O. R. & Scheuerlein, A. An emerging role of zoos to conserve biodiversity. Science 331, 1390–1391 (2011).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    2.Ballou, J. D. et al. Demographic and genetic management of captive populations. in Wild Mammals in Captivity: Principles and Techniques for Zoo Management (eds. Kleiman, D. G., Thompson, K. V. & Kirk Baer, C.) 219–252 (The University of Chicago Press, 2010).3.Ralls, K. & Ballou, J. D. Captive breeding and reintroduction. in Encyclopedia of Biodiversity (ed. Levin, S. A.) 662–667 (Elsevier Academic Press, 2013). https://doi.org/10.1016/B978-0-12-384719-5.00268-9.4.IUCN. Guidelines on the Use of Ex Situ Management for Species Conservation (2nd ed.). www.iucn.org/about/work/programmes/species/publications/iucn_guidelines_and__policy__statements/ (2014).5.Lacy, R. C. Loss of genetic diversity from managed populations: interacting effects of drift, mutation, immigration, selection, and population subdivision. Conserv. Biol. 1, 143–158 (1987).Article 

    Google Scholar 
    6.Lockyear, K. M., MacDonald, S. E., Waddell, W. T. & Goodrowe, K. L. Investigation of captive red wolf ejaculate characteristics in relation to age and inbreeding. Theriogenology 86, 1369–1375 (2016).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    7.Frankham, R. Genetic adaptation to captivity in species conservation programs. Mol. Ecol. 17, 325–333 (2008).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    8.Keller, L. F. & Waller, D. M. Inbreeding effects in wild populations. Trends Ecol. Evol. 17, 230–241 (2002).Article 

    Google Scholar 
    9.Robert, A., Couvet, D. & Sarrazin, F. Integration of demography and genetics in population restorations. Écoscience 14, 463–471 (2007).Article 

    Google Scholar 
    10.Charlesworth, D. & Charlesworth, B. Inbreeding depression and its evolutionary consequences. Annu. Rev. Ecol. Syst. 18, 237–268 (1987).Article 

    Google Scholar 
    11.McPhee, M. E. & McPhee, N. F. Relaxed selection and environmental change decrease reintroduction success in simulated populations: altered selection in captive populations. Anim. Conserv. 15, 274–282 (2012).Article 

    Google Scholar 
    12.Ford, M. J. Selection in captivity during supportive breeding may reduce fitness in the wild. Conserv. Biol. 16, 815–825 (2002).Article 

    Google Scholar 
    13.Stockwell, C. A., Hendry, A. P. & Kinnison, M. T. Contemporary evolution meets conservation biology. Trends Ecol. Evol. 18, 94–101 (2003).Article 

    Google Scholar 
    14.Robert, A. Captive breeding genetics and reintroduction success. Biol. Conserv. 142, 2915–2922 (2009).Article 

    Google Scholar 
    15.Araki, H., Cooper, B. & Blouin, M. S. Genetic effects of captive breeding cause a rapid, cumulative fitness decline in the wild. Science 318, 100–103 (2007).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    16.Christie, M. R., Marine, M. L., French, R. A. & Blouin, M. S. Genetic adaptation to captivity can occur in a single generation. Proc. Natl. Acad. Sci. 109, 238–242 (2012).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    17.West-Eberhard, M. J. Phenotypic plasticity and the origins of diversity. Annu. Rev. Ecol. Syst. 20, 249–278 (1989).Article 

    Google Scholar 
    18.Gordon, S. P., Hendry, A. P. & Reznick, D. N. Predator-induced contemporary evolution, phenotypic plasticity, and the evolution of reaction norms in guppies. Copeia 105, 514–522 (2017).Article 

    Google Scholar 
    19.Forslund, P. & Pärt, T. Age and reproduction in birds—hypotheses and tests. Trends Ecol. Evol. 10, 374–378 (1995).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    20.Smith, J. M. Review lectures on senescence—I. The causes of ageing. Proc. R. Soc. Lond. B Biol. Sci. 157, 115–127 (1962).ADS 
    Article 

    Google Scholar 
    21.Partridge, L. & Barton, N. H. Optimally, mutation and the evolution of ageing. Nature 362, 305–311 (1993).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    22.Jones, O. R. et al. Diversity of ageing across the tree of life. Nature 505, 169–173 (2014).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    23.Langen, K., Bakker, T. C. M., Baldauf, S. A., Shrestha, J. & Thünken, T. Effects of ageing and inbreeding on the reproductive traits in a cichlid fish I: the male perspective. Biol. J. Linn. Soc. 120, 752–761 (2017).Article 

    Google Scholar 
    24.Kirkwood, T. B. L. Evolution of ageing. Nature 270, 301 (1977).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    25.Benton, C. H. et al. Inbreeding intensifies sex- and age-dependent disease in a wild mammal. J. Anim. Ecol. 87, 1500–1511 (2018).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    26.de Boer, R. A., Eens, M. & Müller, W. Sex-specific effects of inbreeding on reproductive senescence. Proc. R. Soc. B Biol. Sci. 285, 20180231 (2018).Article 

    Google Scholar 
    27.Promislow, D. E. L. & Tatar, M. Mutation and senescence: where genetics and demography meet. Genetica 102, 299–314 (1998).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    28.Charlesworth, B. & Hughes, K. A. Age-specific inbreeding depression and components of genetic variance in relation to the evolution of senescence. Proc. Natl. Acad. Sci. 93, 6140–6145 (1996).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    29.Snoke, M. S. & Promislow, D. E. L. Quantitative genetic tests of recent senescence theory: age-specific mortality and male fertility in Drosophila melanogaster. Heredity 91, 546–556 (2003).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    30.Robert, A., Toupance, B., Tremblay, M. & Heyer, E. Impact of inbreeding on fertility in a pre-industrial population. Eur. J. Hum. Genet. 17, 673–681 (2009).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    31.Lesobre, L. et al. Conservation genetics of Houbara Bustard (Chlamydotis undulata undulata): population structure and its implications for the reinforcement of wild populations. Conserv. Genet. 11, 1489–1497 (2010).Article 

    Google Scholar 
    32.Rabier, R., Robert, A., Lacroix, F. & Lesobre, L. Genetic assessment of a conservation breeding program of the houbara bustard (Chlamydotis undulata undulata) in Morocco, based on pedigree and molecular analyses. Zoo Biol. 39, 365–447 (2020).Article 

    Google Scholar 
    33.Hardouin, L. A., Legagneux, P., Hingrat, Y. & Robert, A. Sex-specific dispersal responses to inbreeding and kinship. Anim. Behav. https://doi.org/10.1016/j.anbehav.2015.04.002 (2015).Article 

    Google Scholar 
    34.Cornec, C., Robert, A., Rybak, F. & Hingrat, Y. Male vocalizations convey information on kinship and inbreeding in a lekking bird. Ecol. Evol. 9, 4421–4430 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    35.Vuarin, P. et al. No evidence for prezygotic postcopulatory avoidance of kin despite high inbreeding depression. Mol. Ecol. 27, 5252–5262 (2018).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    36.Bacon, L., Hingrat, Y. & Robert, A. Evidence of reproductive senescence of released individuals in a reinforced bird population. Biol. Conserv. 215, 288–295 (2017).Article 

    Google Scholar 
    37.Chantepie, S. et al. Quantitative genetics of the aging of reproductive traits in the houbara bustard. PLoS ONE 10, e0133140 (2015).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    38.Preston, B. T., Saint Jalme, M., Hingrat, Y., Lacroix, F. & Sorci, G. Sexually extravagant males age more rapidly. Ecol. Lett. 14, 1017–1024 (2011).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    39.Preston, B. T., Saint Jalme, M., Hingrat, Y., Lacroix, F. & Sorci, G. The sperm of aging male bustards retards their offspring’s development. Nat. Commun. 6, 1–9 (2015).Article 
    CAS 

    Google Scholar 
    40.Vuarin, P. et al. Post-copulatory sexual selection allows females to alleviate the fitness costs incurred when mating with senescing males. Proc. R. Soc. B Biol. Sci. 286, 20191675 (2019).Article 

    Google Scholar 
    41.Chargé, R. et al. Quantitative genetics of sexual display, ejaculate quality and size in a lekking species. J. Anim. Ecol. 82, 399–407 (2013).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    42.Chargé, R. et al. Does recognized genetic management in supportive breeding prevent genetic changes in life-history traits?. Evol. Appl. 7, 521–532 (2014).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    43.Gaucher, P. et al. Taxonomy of the Houbara Bustard Chlamydotis undulata subspecies considered on the basis of sexual display and genetic divergence. Ibis 138, 273–282 (1996).Article 

    Google Scholar 
    44.Hingrat, Y., Saint Jalme, M., Chalah, T., Orhant, N. & Lacroix, F. Environmental and social constraints on breeding site selection. Does the exploded-lek and hotspot model apply to the Houbara bustard Chlamydotis undulata undulata?. J. Avian Biol. 39, 393–404 (2008).Article 

    Google Scholar 
    45.Duursma, D. E., Gallagher, R. V., Price, J. J. & Griffith, S. C. Variation in avian egg shape and nest structure is explained by climatic conditions. Sci. Rep. 8, 1–10 (2018).
    Google Scholar 
    46.Cucco, M., Grenna, M. & Malacarne, G. Female condition, egg shape and hatchability: a study on the grey partridge. J. Zool. 287, 186–194 (2012).Article 

    Google Scholar 
    47.Adamou, A.-E. et al. Egg size and shape variation in Rufous Bush Chats Cercotrichas galactotes breeding in date palm plantations: hatching success increases with egg elongation. Avian Biol. Res. 11, 100–107 (2018).Article 

    Google Scholar 
    48.Goriup, P. D. The world status of the Houbara Bustard Chlamydotis undulata. Bird Conserv. Int. 7, 373–397 (1997).Article 

    Google Scholar 
    49.BirdLife International. Chlamydotis undulata. The IUCN Red List of Threatened Species 2016: e.T22728245A90341807. (2016) https://doi.org/10.2305/IUCN.UK.2016-3.RLTS.T22728245A90341807.en.50.Lacroix, F., Seabury, J., Al Bowardi, M. & Renaud, J. The Emirates Center for Wildlife Propagation: developing a comprehensive strategy to secure a self-sustaining population of houbara bustard (Chlamydotis undulata undulata) in Eastern Morocco. Houbara News 5, (2003).
    51.Conway, W. Wild and zoo animal interactive management and habitat conservation. Biodivers. Conserv. 4, 573–594 (1995).Article 

    Google Scholar 
    52.Saint Jalme, M., Gaucher, P. & Paillat, P. Artificial insemination in Houbara bustards (Chlamydotis undulata): influence of the number of spermatozoa and insemination frequency on fertility and ability to hatch. Reproduction 100, 93–103 (1994).CAS 
    Article 

    Google Scholar 
    53.Allendorf, F. W. Delay of adaptation to captive breeding by equalizing family size. Conserv. Biol. 7, 416–419 (1993).Article 

    Google Scholar 
    54.Percie du Sert, N. et al. The ARRIVE guidelines 2.0: updated guidelines for reporting animal research. PLOS Biol. 18, e3000410 (2020).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    55.Vuarin, P. et al. Sperm competition accentuates selection on ejaculate attributes. Biol. Lett. 15, 20180889 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    56.Chalah, T., Seigneurin, F., Blesbois, E. & Brillard, J. P. In vitro comparison of fowl sperm viability in ejaculates frozen by three different techniques and relationship with subsequent fertility in vivo. Cryobiology 39, 185–191 (1999).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    57.Hoyt, D. F. Practical methods of estimating volume and fresh weight of bird eggs. Auk 96, 73–77 (1979).
    Google Scholar 
    58.Wellmann, R. optiSel: Optimum Contribution Selection and Population Genetics. R package version 2.0.2. https://CRAN.R-project.org/package=optiSel (2018).59.R Core Team. R: a language and environment for statistical computing. R Foundation for Statistical Computing. https://www.R-project.org (2019).60.Princée, F. P. G. Exploring Studbooks for Wildlife Management and Conservation (Springer, Berlin, 2016).
    Google Scholar 
    61.Brooks, M. E. et al. glmmTMB balances speed and flexibility among packages for zero-inflated generalized linear mixed modeling. The R Journal. 9, 378–400 (2017).Article 

    Google Scholar 
    62.Ludecke, D., Makowski, D. & Waggoner, P. performance: Assessment of Regression Models Performance. R package version 0.3.0. https://CRAN.R-project.org/package=performance (2019).63.Ludecke, D. ggeffects: tidy data frames of marginal effects from regression models. J. Open Source Softw. 3, 772. https://doi.org/10.21105/joss.00772 (2018).ADS 
    Article 

    Google Scholar 
    64.Wickham, H. ggplot2: elegant graphics for data analysis (Springer, Berlin, 2009).
    Google Scholar 
    65.Newton, I. & Rothery, P. Senescence and reproductive value in sparrowhawks. Ecology 78, 1000–1008 (1997).Article 

    Google Scholar 
    66.Bouwhuis, S., Sheldon, B. C., Verhulst, S. & Charmantier, A. Great tits growing old: selective disappearance and the partitioning of senescence to stages within the breeding cycle. Proc. R. Soc. B Biol. Sci. 276, 2769–2777 (2009).CAS 
    Article 

    Google Scholar 
    67.Angelier, F., Shaffer, S. A., Weimerskirch, H. & Chastel, O. Effect of age, breeding experience and senescence on corticosterone and prolactin levels in a long-lived seabird: the wandering albatross. Gen. Comp. Endocrinol. 149, 1–9 (2006).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    68.Angelier, F., Weimerskirch, H., Dano, S. & Chastel, O. Age, experience and reproductive performance in a long-lived bird: a hormonal perspective. Behav. Ecol. Sociobiol. 61, 611–621 (2007).Article 

    Google Scholar 
    69.Ottinger, M. A. et al. The Japanese quail: a model for studying reproductive aging of hypothalamic systems. Exp. Gerontol. 39, 1679–1693 (2004).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    70.Lecomte, V. J. et al. Patterns of aging in the long-lived wandering albatross. Proc. Natl. Acad. Sci. 107, 6370–6375 (2010).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    71.Opatová, P. et al. Inbreeding depression of sperm traits in the zebra finch Taeniopygia guttata. Ecol. Evol. 6, 295–304 (2016).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    72.Croquet, C. et al. Linear and curvilinear effects of inbreeding on production traits for Walloon Holstein cows. J. Dairy Sci. 90, 465–471 (2007).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    73.Leroy, G. Inbreeding depression in livestock species: review and meta-analysis. Anim. Genet. 45, 618–628 (2014).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    74.Ralls, K. et al. Call for a paradigm shift in the genetic management of fragmented populations: genetic management. Conserv. Lett. 11, e12412 (2018).Article 

    Google Scholar 
    75.Huisman, J., Kruuk, L. E. B., Ellis, P. A., Clutton-Brock, T. & Pemberton, J. M. Inbreeding depression across the lifespan in a wild mammal population. Proc. Natl. Acad. Sci. 113, 3585–3590 (2016).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    76.Frankham, R. & Ralls, K. Inbreeding leads to extinction. Nature 392, 441–442 (1998).ADS 
    CAS 
    Article 

    Google Scholar 
    77.Armbruster, P. & Reed, D. H. Inbreeding depression in benign and stressful environments. Heredity 95, 235–242 (2005).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    78.Robert, A. Negative environmental perturbations may improve species persistence. Proc. R. Soc. B Biol. Sci. 273, 2501–2506 (2006).Article 

    Google Scholar 
    79.Crnokrak, P. & Roff, D. A. Inbreeding depression in the wild. Heredity 83, 260–270 (1999).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    80.Araki, H., Berejikian, B. A., Ford, M. J. & Blouin, M. S. Fitness of hatchery-reared salmonids in the wild. Evol. Appl. 1, 342–355 (2008).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    81.Lynch, M. & O’Hely, M. Captive breeding and genetic fitness of natural populations. Conserv. Genet. 2, 363–378 (2001).Article 

    Google Scholar 
    82.Robert, A., Sarrazin, F., Couvet, D. & Legendre, S. Releasing adults versus young in reintroductions: interactions between demography and genetics. Conserv. Biol. 18, 1078–1087 (2004).Article 

    Google Scholar 
    83.Roche, E. A., Cuthbert, F. J. & Arnold, T. W. Relative fitness of wild and captive-reared piping plovers: does egg salvage contribute to recovery of the endangered Great Lakes population?. Biol. Conserv. 141, 3079–3088 (2008).Article 

    Google Scholar 
    84.Ford, N. B. & Seigel, R. A. Phenotypic plasticity in reproductive traits: evidence from a viviparous snake. Ecology 70, 1768–1774 (1989).Article 

    Google Scholar 
    85.Bacon, L. Etude des paramètres de reproduction et de la dynamique d’une population renforcée d’outardes Houbara nord-africaines (Chlamydotis undulata undulata) au Maroc. (Museum National d’Histoire Naturelle, 2017).86.Robert, A. et al. Defining reintroduction success using IUCN criteria for threatened species: a demographic assessment. Anim. Conserv. 18, 397–406 (2015).Article 

    Google Scholar 
    87.Bacon, L., Robert, A. & Hingrat, Y. Long lasting breeding performance differences between wild-born and released females in a reinforced North African Houbara bustard (Chlamydotis undulata undulata) population: a matter of release strategy. Biodivers. Conserv. 28, 553–570 (2019).Article 

    Google Scholar 
    88.Vuarin, P. et al. Paternal age negatively affects sperm production of the progeny. Ecol. Lett. https://doi.org/10.1111/ele.13696 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    89.Keller, L. F., Reid, J. M. & Arcese, P. Testing evolutionary models of senescence in a natural population: age and inbreeding effects on fitness components in song sparrows. Proc. R. Soc. B Biol. Sci. 275, 597–604 (2008).CAS 
    Article 

    Google Scholar 
    90.Reynolds, R. M. et al. Age specificity of inbreeding load in Drosophila melanogaster and implications for the evolution of late-life mortality plateaus. Genetics 177, 587–595 (2007).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    91.Tan, C. K. W., Pizzari, T. & Wigby, S. Parental age, gametic age, and inbreeding interact to modulate offspring viability in Drosophila melanogaster. Evolution 67, 3043–3051 (2013).PubMed 
    PubMed Central 

    Google Scholar 
    92.Deubel, W., Bassukas, I. D., Schlereth, W., Lorenz, R. & Hempel, K. Age dependent selection against HPRT deficient T lymphocytes in the HPRT± heterozygous mouse. Mutat. Res. Mol. Mech. Mutagen. 351, 67–77 (1996).CAS 
    Article 

    Google Scholar 
    93.Réale, D. & Festa-Bianchet, M. Predator-induced natural selection on temperament in bighorn ewes. Anim. Behav. 65, 463–470 (2003).Article 

    Google Scholar 
    94.Coltman, D. W., Pilkington, J. G., Smith, J. A. & Pemberton, J. M. Parasite-mediated selection against Inbred Soay Sheep in a free-living, island population. Evolution 53, 1259 (1999).PubMed 
    PubMed Central 

    Google Scholar 
    95.Wang, J., Hill, W. G., Charlesworth, D. & Charlesworth, B. Dynamics of inbreeding depression due to deleterious mutations in small populations: mutation parameters and inbreeding rate. Genet. Res. 74, 165–178 (1999).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar  More

  • in

    A prevalent and culturable microbiota links ecological balance to clinical stability of the human lung after transplantation

    Combined culture-dependent and culture-independent approach identifies the prevalent and viable bacterial community members of the human lung post-transplantTo characterize the bacterial community composition of the lung microbiota post-transplant, we performed 16S rRNA gene amplicon sequencing of 234 longitudinal BALF samples from 64 lung transplant recipients collected over a 49-month period (Fig. 1a, Supplementary Table 1). A total of 7164 operational taxonomic units (OTUs) were identified, excluding OTUs contributing to reads in 11 negative control samples32 (see “Methods”, Supplementary Fig. 1a, Supplementary Data 1 and 2). In accordance with previous studies on BALF samples from healthy non-transplant individuals4,5,6,26, we found that Bacteroidetes and Firmicutes followed by Proteobacteria and Actinobacteria are the most abundant phyla in the post-transplant lung (Fig. 1b). Prevalence analysis across all BALF samples showed that the community composition is highly variable with only 22 OTUs shared by ≥50% of the samples (Supplementary Fig. 1b, Supplementary Data 3). However, these 22 OTUs constituted 42% of the total number of rarefied reads, indicating that they are predominant members of the post-transplant lung microbiota (Fig. 1c, Supplementary Fig. 1c, Supplementary Table 2, Supplementary Data 3). They belonged to the genera Prevotella 7, Streptococcus, Veillonella, Neisseria, Alloprevotella, Pseudomonas, Gemella, Granulicatella, Campylobacter, Porphyromonas and Rothia, the majority of which are also prevailing community members in the healthy human lung3,5,7,26, suggesting a considerable overlap in the overall composition of the lung microbiota between the healthy and the transplanted lung.Fig. 1: Combining BALF amplicon sequencing and bacterial culturing to deduce the microbial ecology of deep lung microbiota.a Schematic of the sampling of Bronchoalveolar lavage fluid (BALF) from lung transplant recipients over time (months post-transplant). b Relative abundances (%) of most abundant phyla across BALF samples. Box plots show median (middle line), 25th, 75th percentile (box) and 5th and 95th percentile (whiskers) as well as outliers (single points). c Prevalence (% samples) vs contribution to total reads across samples for most abundant phyla. Dot color shows different genera and size show total rarefied reads. Gray dashed horizontal line shows prevalence ≥50%. d Scatter plot shows correlation between number of observed OTUs and bacterial counts per BALF sample obtained by quantifying 16S rRNA gene copies with qPCR. Linear regression is shown by the blue line with gray shaded area showing 95% confidence interval (n = 234, two-sided, F(1, 232) = 91.04, P = 2.2 × 10−16), Coefficient of correlation; R2 = 0.28. e Bar chart shows lung taxa (genera; OTU IDs) that contributed ≥75% of total bacterial biomass across samples (n = 234). Venn diagram inset shows overlap (yellow) between the most prevalent (≥50% incidence, light blue) and the most abundant (≥75% total count, red) taxa in the transplanted lung. Bar colors also show the same.Full size imageDifferences in bacterial loads between samples can skew community analyses when based on relative abundance profiling alone. Therefore, we used qPCR to determine the total copies of the 16S rRNA gene as an estimate for bacterial counts, and normalized the abundances of each OTU across the 234 samples (absolute abundance). We found that the bacterial counts vastly differed between samples, ranging between 101 and 106 gene copies per ml of BALF (Supplementary Fig. 1d). The number of observed OTUs increased with decreasing counts (Fig. 1d) suggesting that a large fraction of the OTUs were detected in samples of low bacterial biomass and hence represent either transient or extremely low-abundant community members, or sequencing artefacts and contaminations. In turn, 19 of the 7164 OTUs constituted >75% of the total bacterial biomass detected across the 234 BALF samples (Fig. 1e). This included 11 of the 22 most prevalent OTUs (see above) plus eight OTUs that were detected in only a few samples but at very high abundance (Staphylococcus; OTU_2, Corynebacterium 1; OTU_16 and OTU_24, Anaerococcus; OTU_49 and OTU_234, Haemophilus; OTU_78, Streptococcus; OTU_6768, Peptoniphilus; OTU_63, Supplementary Table 2). It is important to differentiate these opportunistic colonizers from other community members with low incidence, as they reached very high bacterial counts in some samples with potential implications for lung health.To demonstrate the viability of prevalent lung microbiota members and to establish a reference catalogue of bacterial isolates from the human lung for experimental studies, we complemented the amplicon sequencing with a bacterial culturing approach (Supplementary Fig. 2). We cultivated 21 random BALF samples from 18 individuals, on 15 different semi-solid media (both general and selective) in combination with 3 oxygen concentrations; aerobic, 5% CO2, and anaerobic (See “Methods” and Supplementary Table 3), representing 26 different conditions. We cultured fresh BALF immediately upon extraction (within 2 h), as we observed loss in bacterial diversity upon cultivating frozen samples. This resulted in a total of 300 bacterial isolates, representing 5 phyla, 7 classes, 13 orders, and 17 families from which we built an open-access biobank called the Lung Microbiota culture Collection (LuMiCol, Supplementary Data 4, https://github.com/sudu87/Microbial-ecology-of-the-transplanted-human-lung).To examine the extent of overlap between bacteria in LuMiCol and the diversity obtained by amplicon sequencing, we included 16S rRNA gene sequences from 215 isolates that passed our quality filter into the community analysis, which allowed for the retrieval of OTU-isolate matching pairs32 (Methods). We found that 213 isolates matched to 47 OTUs (Fig. 2a, c, Supplementary Data 5), including 17 of the most prevalent and abundant bacteria (Fig. 1e, Supplementary Table 2). As expected, bacteria with high abundance in the amplicon sequencing-based community analysis were isolated more frequently, with Firmicutes revealing the highest isolate diversity (Fig. 2a–c, Supplementary Data 4, 5) and being recovered under the most diverse culturing conditions.Fig. 2: A lung microbiota culture collection (LuMiCol) reveals extended diversity and phenotypic characteristics of the lower airway bacterial community.a Phylogenetic tree of the 47 OTU-isolate matching pairs inferred with FastTree. Branch bootstrap support values (size of dark gray circles) ≥80% are displayed. b Growth characteristics of each OTU-isolate matching pair in three different oxygen conditions (Anaerobic – light brown, 5% CO2-yellow, aerobic-light blue, n = 3). Column with pie charts shows growth on semi-solid agar. Heatmap shows median change in Optical Density (OD) at 600 nm growth in three different liquid media (THY, RPMI, RPMI without glucose) over 3 days. c Cumulative counts of each OTU-isolate matching pair across all BALF samples (gray). d Number of isolates in Lumicol (black) per OTU-isolate matching pair. Taxa are labeled as genus; OTU ID, with an indication of whether they are prevalent (gray rectangle) or opportunistic (magenta rectangle) in the lower airway community. The names of the closest hit in databases: eHOMD and SILVA are used as species descriptor.Full size imageIn summary, our results from the combined culture-dependent and culture-independent approach show that the lung microbiota post-transplant is highly variable in terms of both bacterial load and community composition with many transient and low-abundant bacterial taxa. However, a few community members display relatively high prevalence and/or abundance suggesting that they represent important colonizers of the human lung.LuMiCol informs on the diversity and metabolic preferences of culturable human lung bacteriaWe characterized the culturable community members of the lower respiratory tract contained in LuMiCol by testing a wide range of growth conditions and phenotypic properties (see “Methods”). The majority of the cultured isolates could taxonomically be assigned at the species level based on genotyping of the 16S rRNA gene V1-V5 region. However, the limited taxonomic resolution offered by this method does not allow to discriminate between closely related strains, which can include both pathogenic and non-pathogenic bacteria. Hence for Streptococcus, we additionally tested for type of hemolysis (alpha, beta, or gamma) and resistance to optochin, which differentiates the pathogenic pneumococcus and the non-pathogenic viridans groups (Fig. 2a, Supplementary Fig. 2b, c). This demonstrated that the 16 Streptococcus OTU-isolate pairs belong to the viridans group of streptococci (VS)33. Interestingly, these isolates exhibited the highest genotypic and phenotypic diversity throughout our collection and belonged to five OTUs among the 22 most prevalent community members, with Streptococcus mitis (OTU_11) present in 93.6% of all samples.BALF from healthy individuals contains amino acids, citrate, urate, fatty acids, and antioxidants such as glutathione but no detectable glucose34, which is associated with increased bacterial load and infection35,36,37. To get insights into basic bacterial metabolism, we assessed the growth of all 47 isolates matching an OTU under different oxygen concentrations. We used undefined rich media (Todd-Hewitt Yeast extract) and defined low-complexity liquid media (RPMI 1640), including a glucose-free version to mimic the deep lung environment (see “Methods”). Despite the presence of oxygen in the human lung, the majority of the isolates were either obligate or facultative anaerobes (Fig. 2a), including some of the most prevalent members (Prevotella melaninogenica (OTU_3), Streptococcus mitis (OTU_11), Veillonella atypica (OTU_6) and Granulicatella adiacens (OTU_17). A similar trend was also observed in liquid media under anaerobic conditions, with the exception of the genera Prevotella, Veillonella and Granulicatella. Most streptococci from the human lung grew best in complex liquid media containing glucose under anaerobic conditions, including the most prevalent species in our cohort, S. mitis (OTU_11) (Fig. 2b). However, noticeable exceptions were S. vestibularis (OTU_34), S. oralis (OTU_3427 and OTU_1567), and S. gordonii (OTU_10031), which grew equally well in the presence of oxygen and in low-complexity liquid medium (Fig. 2b). Most Actinobacteria grew best on rich medium in the presence of 5% CO2, with an exception of Actinomyces odontolyticus (OTU_39), which required anaerobic conditions. Some Actinobacteria grew equally well in anaerobic conditions as in the presence of 5% CO2, i.e., Corynebacterium durum (OTU_501), Actinobacteria sp. oral taxon (OTU_328 and OTU_228).The two most predominant opportunistic pathogens in our lung cohort, P. aeruginosa (OTU_1) and S. aureus (OTU_2), grew best in rich liquid medium in the presence of oxygen (Fig. 2c), although these also grew to lower degree under anaerobic conditions. These results indicate that changes in the physicochemical conditions in the lung may favor the growth of these two opportunistic pathogens. In summary, our observations from the bacterial culture collection provide first insights into the phenotypic properties of human lung bacteria and will serve as a basis for future experimental work.Identification of four compositionally distinct pneumotypes post-transplant using machine learning based on ecological metricsTo detect and characterize differences in bacterial community composition between BALF samples from transplant patients, we clustered the samples using an unsupervised machine learning algorithm based on pairwise Bray–Curtis dissimilarity32 (beta diversity, See “Methods”, Supplementary Data 6). This segregated the samples into four partitions around medoids (PAMs) at both phylum and OTU level (Fig. 3a, b, Supplementary Fig. 3a, b). We refer to these clusters as “pneumotypes” PAM1, PAM2, PAM3, and PAM4 (Supplementary Table 4). PAM1 formed the largest cluster consisting of the majority of samples (n = 115) followed by PAM3 (n = 76), PAM2 (n = 19), and PAM4 (n = 24) (Supplementary Data 7). Examination of various diversity measures (Species occurrence, OTU diversity, OTU richness, Fig. 3c–e), distribution of the dominant community members (Fig. 3f), and bacterial counts (16S rRNA gene copies, Fig. 3g) revealed distinctive characteristics between the four pneumotypes.Fig. 3: Bacterial communities of the lung post-transplant fall into four ‘pneumotypes’ with distinct community characteristics.a, b Principal component analysis shows Partition around medoids (PAMs) at phylum and OTU level respectively generated by k-medoid-based unsupervised machine learning using Bray–Curtis dissimilarity (occurrence and abundance). Pneumotypes are color coded: Balanced (red, n = 115), Staphylococcus (green, n = 19), Microbiota-depleted (MD, blue, n = 76), and Pseudomonas (orange, n = 24). c–g Violin plots show distributions of pairwise species occurrence (Sorenson’s index, PERMANOVA, two-sided, F(3, 229) = 8.49, P = 9.9 × 10−5), OTU diversity (Kruskal–Wallis test, χ2 = 89.2, df = 3, two-sided, P = 2.2 × 10−16), OTU richness (ANOVA, F(3, 229) = 43.9, two-sided, P = 2.2 × 10−16), proportion of most dominant OTUs (Kruskal–Wallis test, χ2 = 94.45, df = 3, two-sided, P = 2.2 × 10−16), and total bacterial counts (ANOVA, F(3, 229) = 43.9, two-sided, P = 2.2 × 10−16), respectively, across the four pneumotypes. h, i Enrichment analysis of prevalence (green dotted line ≥50%) and absolute abundance across all samples of the 30 most dominant taxa (i.e., OTUs) in PneumotypeBalanced and PneumotypeMD respectively, when each was compared to the other three combined pneumotypes (gray boxes). Differential abundances after enrichment analysis was calculated between each PAM and the other three PAMs combined, using ART-ANOVA. j Heatmap shows relative percentage of taxa (right colored panel) cultured from paired samples of Bronchial aspiration (BA) and Bronchoalveolar lavage fluid (BALF) from each pneumotype (left colored panel). Oropharyngeal flora mainly corresponds to PneumotypeBalanced (i.e., Streptococcus, Prevotella, Veillonella). All box plots including insets show median (middle line), 25th, 75th percentile (box) and 5th and 95th percentile (whiskers) as well as outliers (single points). Multiple comparison of beta diversity indices was done by pairwise PERMANOVA (adonis) with False Discovery rate (FDR). Post hoc analyses (95% Confidence Interval) were done by using Tukey’s test (ANOVA) or Dunn’s test (Kruskal test) with False Discovery Rate (FDR) or least-squares means (ART-ANOVA) with False Discovery Rate (FDR). * P  More

  • in

    Reply to: “Results from a biodiversity experiment fail to represent economic performance of semi-natural grasslands”

    1.Schaub, S. et al. Plant diversity effects on forage quality, yield and revenues of semi-natural grasslands. Nat. Commun. 11, 1–11 (2020).Article 

    Google Scholar 
    2.Tonn, B., Komainda, M. & Isselstein, J. Results from a biodiversity experiment fail to represent economic performance of semi-natural grasslands. Nat. Commun. https://doi.org/10.1038/s41467-021-22309-7 (2021).3.Roscher, C. et al. The role of biodiversity for element cycling and trophic interactions: an experimental approach in a grassland community. Basic Appl. Ecol. 5, 107–121 (2004).Article 

    Google Scholar 
    4.Jochum, M. et al. The results of biodiversity–ecosystem functioning experiments are realistic. Nat. Ecol. Evol. 4, 1485–1494 (2020).Article 

    Google Scholar 
    5.Roscher, C., Schumacher, J., Weisser, W. W., Schmid, B. & Schulze, E. D. Detecting the role of individual species for overyielding in experimental grassland communities composed of potentially dominant species. Oecologia 154, 535–549 (2007).ADS 
    Article 

    Google Scholar 
    6.Deak, A., Hall, M., Sanderson, M. & Archibald, D. Production and nutritive value of grazed simple and complex forage mixtures. Agron. J. 99, 814–821 (2007).Article 

    Google Scholar 
    7.Sturludóttir, E. et al. Benefits of mixing grasses and legumes for herbage yield and nutritive value in Northern Europe and Canada. Grass Forage Sci. 69, 229–240 (2014).Article 

    Google Scholar 
    8.Oelmann, Y., Vogel, A., Wegener, F., Weigelt, A. & Scherer-Lorenzen, M. Management intensity modifies plant diversity effects on N yield and mineral N in soil. Soil Sci. Soc. Am. J. 79, 559–568 (2015).ADS 
    CAS 
    Article 

    Google Scholar 
    9.Schaub, S., Buchmann, N., Lüscher, A. & Finger, R. Economic benefits from plant species diversity in intensively managed grasslands. Ecol. Econ. 168, 106488 (2020b).Article 

    Google Scholar 
    10.Trenbath, B. R. Biomass productivity of mixtures. Adv. Agron. 26, 177–210 (1974).Article 

    Google Scholar 
    11.Binder, S., Isbell, F., Polasky, S., Catford, J. A. & Tilman, D. Grassland biodiversity can pay. Proc. Natl Acad. Sci. USA 115, 3876–3881 (2018).CAS 
    Article 

    Google Scholar 
    12.Weigelt, A., Weisser, W., Buchmann, N. & Scherer‐Lorenzen, M. Biodiversity for multifunctional grasslands: equal productivity in high‐diversity low‐input and low‐diversity high‐input systems. Biogeosciences 6, 1695–1706 (2009).ADS 
    CAS 
    Article 

    Google Scholar 
    13.Vogel, A., Scherer-Lorenzen, M. & Weigelt, A. Grassland resistance and resilience after drought depends on management intensity and species richness. PLoS ONE 7, e36992 (2012).ADS 
    CAS 
    Article 

    Google Scholar 
    14.Finn, J. A. et al. Ecosystem function enhanced by combining four functional types of plant species in intensively managed grassland mixtures: a 3‐year continental‐scale field experiment. J. Appl. Ecol. 50, 365–375 (2013).Article 

    Google Scholar 
    15.Jans, F., Kessler, J., Münger, A. & Schlegel, P. in Fütterungsempfehlungen für Wiederkäuer (Grünes Buch) Ch. 7 (Agroscope, 2015).16.FAO (Food and Agriculture Organization of the United Nations), IDF (International Dairy Federation), and IFCN (IFCN Dairy Research Network). World Mapping of Animal Feeding Systems in the Dairy Sector. (FAO, 2014).17.Delaby, L., Peyraud, J. L., Foucher, N. & Michel, G. The effect of two contrasting grazing managements and level of concentrate supplementation on the performance of grazing dairy cows. Anim. Res. 52, 437–460 (2003).Article 

    Google Scholar 
    18.Leiber, F., Wettstein, H. R. & Kreuzer, M. Is the intrinsic potassium content of forages an important factor in intake regulation of dairy cows? J. Anim. Physiol. Anim. Nutr. 93, 391–399 (2009).CAS 
    Article 

    Google Scholar 
    19.Schaub, S. et al. Data: forage quality and biomass yield of the Management Experiment set up within the Jena Experiment. ETH Zur. Res. Collect. https://doi.org/10.3929/ethz-b-000374100 (2019). More