More stories

  • in

    Repellent, oviposition-deterrent, and insecticidal activity of the fungal pathogen Colletotrichum fioriniae on Drosophila suzukii (Diptera: Drosophilidae) in highbush blueberries

    1.
    Walsh, D. B. et al. Drosophila suzukii (Diptera: Drosophilidae): Invasive pest of ripening soft fruit expanding its geographic range and damage potential. J. Integr. Pest Manag 2, G1–G7 (2011).
    Google Scholar 
    2.
    Hauser, M. A historic account of the invasion of Drosophila suzukii (Matsumura) (Diptera: Drosophilidae) in the continental United States, with remarks on their identification. Pest Manag. Sci. 67, 1352–1357 (2011).
    CAS  PubMed  Google Scholar 

    3.
    Asplen, M. K. et al. Invasion biology of spotted wing drosophila (Drosophila suzukii): a global perspective and future priorities. J. Pest. Sci. 88, 469–494 (2015).
    Google Scholar 

    4.
    Arnó, J., Solà, M., Riudavets, J. & Gabarra, R. Population dynamics, non-crop hosts, and fruit susceptibility of Drosophila suzukii in Northeast Spain. J. Pest. Sci. 89, 713–723 (2016).
    Google Scholar 

    5.
    Keesey, I. W., Knaden, M. & Hansson, B. S. Olfactory specialization in Drosophila suzukii supports an ecological shift in host preference from rotten to fresh fruit. J. Chem. Ecol. 41, 121–128 (2015).
    CAS  PubMed  PubMed Central  Google Scholar 

    6.
    Karageorgi, M. et al. Evolution of multiple sensory systems drives novel egg-laying behavior in the fruit pest Drosophila suzukii. Curr. Biol. 27, 847–853 (2017).
    CAS  PubMed  PubMed Central  Google Scholar 

    7.
    Lee, J. C. et al. The susceptibility of small fruits and cherries to the spotted-wing drosophila Drosophila suzukii. Pest Manag. Sci. 67, 1358–1367 (2011).
    CAS  PubMed  Google Scholar 

    8.
    Raffa, K. F., Bonello, P. & Orrock, J. L. Why do entomologists and plant pathologists approach trophic relationships so differently? Identifying biological distinctions to foster synthesis. New Phytol. 225, 609–620 (2020).
    PubMed  Google Scholar 

    9.
    Scheidler, N. H., Liu, C., Hamby, K. A., Zalom, F. G. & Syed, Z. Volatile codes: Correlation of olfactory signals and reception in Drosophila-yeast chemical communication. Sci. Rep. 5, 1–13 (2015).
    Google Scholar 

    10.
    Hamm, C. A. et al. Wolbachia do not live by reproductive manipulation alone: Infection polymorphism in Drosophila suzukii and D Subpulchrella. Mol. Ecol. 23, 4871–4885 (2014).
    PubMed  PubMed Central  Google Scholar 

    11.
    Cha, D. H. et al. Behavioral evidence for contextual olfactory-mediated avoidance of the ubiquitous phytopathogen Botrytis cinerea by Drosophila suzukii. Insect Sci. 27, 771–779 (2019).
    PubMed  Google Scholar 

    12.
    Bellutti, N. et al. Dietary yeast affects preference and performance in Drosophila suzukii. J. Pest. Sci. 91, 651–660 (2018).
    Google Scholar 

    13.
    Hamby, K. A., Hernández, A., Boundy-Mills, K. & Zalom, F. G. Associations of yeasts with spotted-wing drosophila (Drosophila suzukii; Diptera: Drosophilidae) in cherries and raspberries. Appl. Environ. Microbiol. 78, 4869–4873 (2012).
    CAS  PubMed  PubMed Central  Google Scholar 

    14.
    Mori, B. A. et al. Enhanced yeast feeding following mating facilitates control of the invasive fruit pest Drosophila suzukii. J. Appl. Ecol. 54, 170–177 (2017).
    Google Scholar 

    15.
    Goodhue, R. E., Bolda, M., Farnsworth, D., Williams, J. C. & Zalom, F. G. Spotted wing drosophila infestation of California strawberries and raspberries: Economic analysis of potential revenue losses and control costs. Pest Manag. Sci. 67, 1396–1402 (2011).
    CAS  PubMed  Google Scholar 

    16.
    Barata, A., Malfeito-Ferreira, M. & Loureiro, V. The microbial ecology of wine grape berries. Int. J. Food Microbiol. 153, 243–259 (2012).
    CAS  PubMed  Google Scholar 

    17.
    Cloonan, K. R., Abraham, J., Angeli, S., Syed, Z. & Rodriguez-Saona, C. Advances in the chemical ecology of the spotted wing drosophila (Drosophila suzukii) and its Applications. J. Chem. Ecol. 44, 922–939 (2018).
    CAS  PubMed  Google Scholar 

    18.
    Cloonan, K. R. et al. Laboratory and field evaluation of host-related foraging odor-cue combinations to attract Drosophila suzukii (Diptera: Drosophilidae). J. Econ. Entomol. 112, 2850–2860 (2019).
    PubMed  Google Scholar 

    19.
    Waller, T. J., Vaiciunas, J., Constantelos, C. & Oudemans, P. V. Evidence that blueberry floral extracts influence secondary conidiation and appressorial formation of Colletotrichum fioriniae. Phytopathology 108, 561–567 (2018).
    PubMed  Google Scholar 

    20.
    Pszczółkowska, A. & Okorski, A. First report of anthracnose disease caused by Colletotrichum fioriniae on blueberry in western Poland. Plant. Dis. 100, 21–67 (2016).
    Google Scholar 

    21.
    Wharton, P. & Diéguez-Uribeondo, J. The biology of Colletotrichum acutatum. An del Jardín Botánico Madrid 61, 3–22 (2004).
    Google Scholar 

    22.
    Peres, N. A., Timmer, L. W., Adaskaveg, J. E. & Correll, J. C. Lifestyles of Colletotrichum acutatum. Plant Dis. 89, 784–796 (2005).
    CAS  PubMed  Google Scholar 

    23.
    Polashock, J. J., Caruso, F. L., Averill, A. L. & Schilder, A. C. Compendium of Bluberry, Cranberry, and Lingonberry Diseases and Pests (APS Publications, St. Paul, MN, 2017).
    Google Scholar 

    24.
    Wharton, P. S. & Schilder, A. C. Novel infection strategies of Colletotrichum acutatum on ripe blueberry fruit. Plant Pathol. 57, 122–134 (2008).
    Google Scholar 

    25.
    Verma, N., MacDonald, L. & Punja, Z. K. Inoculum prevalence, host infection and biological control of Colletotrichum acutatum: causal agent of blueberry anthracnose in British Columbia. Plant Pathol. 55, 442–450 (2006).
    Google Scholar 

    26.
    Verma, N., MacDonald, L. & Punja, Z. K. Environmental and host requirements for field infection of blueberry fruits by Colletotrichum acutatum in British Columbia. Plant Pathol. 56, 107–113 (2007).
    Google Scholar 

    27.
    Miles, T. D. & Schilder, A. C. Host defenses associated with fruit infection by Colletotrichum species with an emphasis on anthracnose of blueberries. Plant Health Prog. 14, 30 (2013).
    Google Scholar 

    28.
    Miles, T. D., Hancock, J. F., Callow, P. & Schilder, A. M. C. Evaluation of screening methods and fruit composition in relation to anthracnose fruit rot resistance in blueberries. Plant Pathol. 61, 555–566 (2012).
    Google Scholar 

    29.
    Janzen, D. H. Why fruits rot, seeds mold, and meat spoils. Am. Nat. 111, 691–713 (1977).
    CAS  Google Scholar 

    30.
    Cipollini, M. L. & Stiles, E. W. Fruit rot, antifungal defense, and palatability of fleshy fruits for frugivorous birds. Ecology 74, 751–762 (1993).
    Google Scholar 

    31.
    Peris, J. E., Rodríguez, A., Penã, L. & Fedriani, J. M. Fungal infestation boosts fruit aroma and fruit removal by mammals and birds. Sci. Rep. 7, 1–9 (2017).
    CAS  Google Scholar 

    32.
    Lee, J. C. et al. Characterization and manipulation of fruit susceptibility to Drosophila suzukii. J. Pest. Sci. 89, 771–780 (2016).
    Google Scholar 

    33.
    Choi, M. Y. et al. Effect of non-nutritive sugars to decrease the survivorship of spotted wing drosophila Drosophila suzukii. J Insect Physiol 99, 86–94 (2017).
    CAS  PubMed  Google Scholar 

    34.
    Tochen, S., Walton, V. M. & Lee, J. C. Impact of floral feeding on adult Drosophila suzukii survival and nutrient status. J. Pest Sci. 89, 793–802 (2016).
    Google Scholar 

    35.
    Young, Y., Buckiewicz, N. & Long, T. A. F. Nutritional geometry and fitness consequences in Drosophila suzukii, the spotted-wing drosophila. Ecol. Evol. 8, 2842–2851 (2018).
    PubMed  PubMed Central  Google Scholar 

    36.
    Graziosi, I. & Rieske, L. K. A plant pathogen causes extensive mortality in an invasive insect herbivore. Agric. For. Entomol. 17, 366–374 (2015).
    Google Scholar 

    37.
    Wallingford, A. K., Hesler, S. P., Cha, D. H. & Loeb, G. M. Behavioral response of spotted-wing drosophila, Drosophila suzukii Matsumura, to aversive odors and a potential oviposition deterrent in the field. Pest Manag. Sci. 72, 701–706 (2016).
    CAS  PubMed  Google Scholar 

    38.
    Wallingford, A. K., Cha, D. H., Linn, C. E., Wolfin, M. S. & Loeb, G. M. Robust manipulations of pest insect behavior using repellents and practical application for integrated pest management. Environ. Entomol. 46, 1041–1050 (2017).
    CAS  PubMed  Google Scholar 

    39.
    Göhre, V. & Robatzek, S. Breaking the barriers: microbial effector molecules subvert plant immunity. Annu. Rev. Phytopathol. 46, 189–215 (2008).
    PubMed  Google Scholar 

    40.
    Csorba, T., Kontra, L. & Burgyán, J. Viral silencing suppressors: tools forged to fine-tune host-pathogen coexistence. Virology 479–480, 85–103 (2015).
    PubMed  Google Scholar 

    41.
    Stringlis, I. A., Zhang, H., Pieterse, C. M. J., Bolton, M. D. & De Jonge, R. Microbial small molecules-weapons of plant subversion. Nat. Prod. Rep. 35, 410–433 (2018).
    CAS  PubMed  Google Scholar 

    42.
    McLeod, G. et al. The pathogen causing Dutch elm disease makes host trees attract insect vectors. Proc. Biol. Sci. 272, 2499–2503 (2005).
    PubMed  PubMed Central  Google Scholar 

    43.
    Raguso, R. A. & Roy, B. A. ‘Floral’ scent production by Puccinia rust fungi that mimic flowers. Mol. Ecol. 7, 1127–1136 (1998).
    CAS  PubMed  Google Scholar 

    44.
    Bruce, T. J. A. & Pickett, J. A. Perception of plant volatile blends by herbivorous insects—finding the right mix. Phytochemistry 72, 1605–1611 (2011).
    CAS  PubMed  Google Scholar 

    45.
    Revadi, S. et al. Sexual behavior of Drosophila suzukii. Insects 6, 183–196 (2015).
    PubMed  PubMed Central  Google Scholar 

    46.
    Polashock, J. J., Ehlenfeldt, M. K., Stretch, A. W. & Kramer, M. Anthracnose fruit rot resistance in blueberry cultivars. Plant Dis. 89, 33–38 (2005).
    PubMed  Google Scholar 

    47.
    Hartung, J. S., Burton, C. & Ramsdell, D. C. Epidemiological studies of blueberry anthracnose disease caused by Colletotrichum gloeosporioides. Phytopathology 71, 449 (1981).
    Google Scholar 

    48.
    Cai, P. et al. Potential host fruits for Drosophila suzukii: olfactory and oviposition preferences and suitability for development. Entomol. Exp. Appl. 167, 880–890 (2019).
    Google Scholar 

    49.
    Rodriguez-Saona, C. et al. Differential susceptibility of wild and cultivated blueberries to an invasive frugivorous pest. J. Chem. Ecol. 45, 286–297 (2018).
    PubMed  Google Scholar 

    50.
    Hodge, S. The effect of pH and water content of natural resources on the development of Drosophila melanogaster larvae. Dros. Inf. Serv. 84, 38–43 (2001).
    Google Scholar 

    51.
    Schilder, A. M. C., Gillett, J. M. & Woodworth, J. A. The kaleidoscopic nature of blueberry fruit roots. Acta Hortic. 574, 81–83 (2002).
    Google Scholar 

    52.
    Jaramillo, S. L., Mehlferber, E. & Moore, P. J. Life-history trade-offs under different larval diets in Drosophila suzukii (Diptera: Drosophilidae). Physiol. Entomol. 40, 2–9 (2015).
    Google Scholar 

    53.
    Dalton, D. T. et al. Laboratory survival of Drosophila suzukii under simulated winter conditions of the Pacific Northwest and seasonal field trapping in five primary regions of small and stone fruit production in the United States. Pest Manag. Sci. 67, 1368–1374 (2011).
    CAS  PubMed  Google Scholar 

    54.
    Miller, P. M. V-8 juice agar as a general purpose medium for fungi and bacteria. Phytopathology 45, 461–462 (1955).
    Google Scholar 

    55.
    Feng, Y., Bruton, R., Park, A. & Zhang, A. Identification of attractive blend for spotted wing drosophila, Drosophila suzukii, from apple juice. J. Pest Sci. 91, 1251–1267 (2018).
    Google Scholar 

    56.
    Tochen, S. et al. Temperature-related development and population parameters for Drosophila suzukii (Diptera: Drosophilidae) on cherry and blueberry. Environ. Entomol. 43, 501–510 (2014).
    PubMed  Google Scholar  More

  • in

    Microbial carrying capacity and carbon biomass of plastic marine debris

    1.
    Van Sebille E, Wilcox C, Lebreton L, Maximenko N, Hardesty BD, Van Franeker JA, et al. A global inventory of small floating plastic debris. Environ Res Lett. 2015;10:124006.
    Google Scholar 
    2.
    Reisser J, Shaw J, Hallegraeff G, Proietti M, Barnes DK, Thums M, et al. Millimeter-sized marine plastics: a new pelagic habitat for microorganisms and invertebrates. PLoS ONE. 2014;9:e100289.
    PubMed  PubMed Central  Google Scholar 

    3.
    Mincer TJ, Zettler ER, Amaral-Zettler LA. Biofilms on plastic debris and their influence on marine nutrient cycling, productivity, and hazardous chemical mobility. In: Rei Yamashita KT, Bee Geok Yeo, Hideshige Takada, Jan A. van Franeker, Megan Dalton, Eric Dale, editors. Hazardous chemicals associated with plastics in the marine environment. Springer: Cham; 2016. pp. 221–33.

    4.
    Morét-Ferguson S, Law KL, Proskurowski G, Murphy EK, Peacock EE, Reddy CM. The size, mass, and composition of plastic debris in the western North Atlantic Ocean. Mar Pollut Bull. 2010;60:1873–8.
    PubMed  Google Scholar 

    5.
    Eriksen M, Lebreton LC, Carson HS, Thiel M, Moore CJ, Borerro JC, et al. Plastic pollution in the world’s oceans: more than 5 trillion plastic pieces weighing over 250,000 tons afloat at sea. PLoS ONE. 2014;9:e111913.
    PubMed  PubMed Central  Google Scholar 

    6.
    Zettler ER, Mincer TJ, Amaral-Zettler LA. Life in the “plastisphere”: microbial communities on plastic marine debris. Environ Sci Technol. 2013;47:7137–46.
    CAS  PubMed  Google Scholar 

    7.
    Amaral-Zettler LA, Zettler ER, Slikas B, Boyd GD, Melvin DW, Morrall CE, et al. The biogeography of the plastisphere: implications for policy. Front Ecol Environ. 2015;13:541–6.
    Google Scholar 

    8.
    De Tender CA, Devriese LI, Haegeman A, Maes S, Ruttink T, Dawyndt P. Bacterial community profiling of plastic litter in the Belgian part of the North Sea. Environ Sci Technol. 2015;49:9629–38.
    PubMed  Google Scholar 

    9.
    De Tender CA, Schlundt C, Devriese LI, Mincer TJ, Zettler ER, Amaral-Zettler LA. A review of microscopy and comparative molecular-based methods to characterize “plastisphere” communities. Anal Methods. 2017;9:2132–43.
    Google Scholar 

    10.
    Gong W, Marchetti A. Estimation of 18S gene copy number in marine eukaryotic plankton using a next-generation sequencing approach. Front Mar Sci. 2019;6:219.
    Google Scholar 

    11.
    Bonk F, Popp D, Harms H, Centler F. PCR-based quantification of taxa-specific abundances in microbial communities: quantifying and avoiding common pitfalls. J Microbiol Methods. 2018;153:139–47.
    CAS  PubMed  Google Scholar 

    12.
    Neu TR, Lawrence JR. Innovative techniques, sensors, and approaches for imaging biofilms at different scales. Trends Microbiol. 2015;23:233–42.
    CAS  PubMed  Google Scholar 

    13.
    Bochdansky AB, Clouse MA, Herndl GJ. Eukaryotic microbes, principally fungi and labyrinthulomycetes, dominate biomass on bathypelagic marine snow. ISME J. 2017;11:362–73.
    PubMed  Google Scholar 

    14.
    Schlundt C, Welch JLM, Knochel AM, Zettler ER, Amaral‐Zettler LA. Spatial structure in the “plastisphere”: molecular resources for imaging microscopic communities on plastic marine debris. Mol Ecol Resour. 2020;20:620–634.
    CAS  PubMed  Google Scholar 

    15.
    Bruinsma G, Van der Mei H, Busscher H. Bacterial adhesion to surface hydrophilic and hydrophobic contact lenses. Biomaterials. 2001;22:3217–24.
    CAS  PubMed  Google Scholar 

    16.
    Ogonowski M, Motiei A, Ininbergs K, Hell E, Gerdes Z, Udekwu KI, et al. Evidence for selective bacterial community structuring on microplastics. Environ Microbiol. 2018;20:2796–808.
    CAS  PubMed  Google Scholar 

    17.
    Khachikyan A, Milucka J, Littmann S, Ahmerkamp S, Meador T, Könneke M, et al. Direct cell mass measurements expand the role of small microorganisms in nature. Appl Environ Microbiol. 2019;85:e00493–19.
    CAS  PubMed  PubMed Central  Google Scholar 

    18.
    Romanova N, Sazhin A. Relationships between the cell volume and the carbon content of bacteria. Oceanology. 2010;50:522–30.
    Google Scholar 

    19.
    Menden-Deuer S, Lessard EJ. Carbon to volume relationships for dinoflagellates, diatoms, and other protist plankton. Limnol Oceanogr. 2000;45:569–79.
    CAS  Google Scholar 

    20.
    Massana R, Logares R. Eukaryotic versus prokaryotic marine picoplankton ecology. Environ Microbiol. 2013;15:1254–61.
    PubMed  Google Scholar 

    21.
    Loferer-Krößbacher M, Klima J, Psenner R. Determination of bacterial cell dry mass by transmission electron microscopy and densitometric image analysis. Appl Environ Microbiol. 1998;64:688–94.
    PubMed  PubMed Central  Google Scholar 

    22.
    Lee S, Fuhrman JA. Relationships between biovolume and biomass of naturally derived marine bacterioplankton. Appl Environ Microbiol. 1987;53:1298–303.
    CAS  PubMed  PubMed Central  Google Scholar 

    23.
    Erni-Cassola G, Zadjelovic V, Gibson MI, Christie-Oleza JA. Distribution of plastic polymer types in the marine environment; a meta-analysis. J Hazard Mater. 2019;369:691–8.
    CAS  PubMed  Google Scholar 

    24.
    Dudek KL, Cruz BN, Polidoro B, Neuer S. Microbial colonization of microplastics in the Caribbean Sea. Limnol Oceanogr Lett. 2020;5:5–17.
    Google Scholar 

    25.
    Carpenter EJ, Smith K. Plastics on the Sargasso Sea surface. Science. 1972;175:1240–1.
    CAS  PubMed  Google Scholar 

    26.
    Amaral-Zettler LA, Zettler ER, Mincer TJ. Ecology of the plastisphere. Nat Rev Microbiol. 2020;18:139–51.
    CAS  PubMed  Google Scholar 

    27.
    Patil JS, Anil AC. Biofilm diatom community structure: influence of temporal and substratum variability. Biofouling. 2005;21:189–206.
    CAS  PubMed  Google Scholar 

    28.
    Rummel CD, Jahnke A, Gorokhova E, Kühnel D, Schmitt-Jansen M. Impacts of biofilm formation on the fate and potential effects of microplastic in the aquatic environment. Environ Sci Technol Lett. 2017;4:258–67.
    CAS  Google Scholar 

    29.
    Michels J, Stippkugel A, Lenz M, Wirtz K, Engel A. Rapid aggregation of biofilm-covered microplastics with marine biogenic particles. Proc R Soc B. 2018;285:20181203.
    PubMed  Google Scholar 

    30.
    Lobelle D, Cunliffe M. Early microbial biofilm formation on marine plastic debris. Mar Pollut Bull. 2011;62:197–200.
    CAS  PubMed  Google Scholar 

    31.
    Mueller LN, de Brouwer JF, Almeida JS, Stal LJ, Xavier JB. Analysis of a marine phototrophic biofilm by confocal laser scanning microscopy using the new image quantification software PHLIP. BMC Ecol. 2006;6:1.
    PubMed  PubMed Central  Google Scholar 

    32.
    De Tender CA, Devriese LI, Haegeman A, Maes S, Vangeyte JR, Cattrijsse A, et al. Temporal dynamics of bacterial and fungal colonization on plastic debris in the North Sea. Environ Sci Technol. 2017;51:7350–60.
    PubMed  Google Scholar 

    33.
    Tetu SG, Sarker I, Schrameyer V, Pickford R, Elbourne LD, Moore LR, et al. Plastic leachates impair growth and oxygen production in Prochlorococcus, the ocean’s most abundant photosynthetic bacteria. Commun Biol. 2019;2:1–9.
    Google Scholar 

    34.
    Capolupo M, Sørensen L, Jayasena KDR, Booth AM, Fabbri E. Chemical composition and ecotoxicity of plastic and car tire rubber leachates to aquatic organisms. Water Res. 2020;169:115270.
    CAS  PubMed  Google Scholar 

    35.
    Vosshage AT, Neu TR, Gabel F. Plastic alters biofilm quality as food resource of the freshwater Gastropod Radix balthica. Environ Sci Technol. 2018;52:11387–93.
    CAS  PubMed  Google Scholar 

    36.
    Dussud C, Meistertzheim A, Conan P, Pujo-Pay M, George M, Fabre P, et al. Evidence of niche partitioning among bacteria living on plastics, organic particles and surrounding seawaters. Environ Pollut. 2018;236:807–16.
    CAS  PubMed  Google Scholar 

    37.
    Armitage AR, Gonzalez VL, Fong P. Decoupling of nutrient and grazer impacts on a benthic estuarine diatom assemblage. Estuar Coast Shelf Sci. 2009;84:375–82.
    CAS  PubMed  PubMed Central  Google Scholar 

    38.
    Yokota K, Waterfield H, Hastings C, Davidson E, Kwietniewski E, Wells B. Finding the missing piece of the aquatic plastic pollution puzzle: interaction between primary producers and microplastics. Limnol Oceanogr Lett. 2017;2:91–104.
    Google Scholar 

    39.
    Oberbeckmann S, Kreikemeyer B, Labrenz M. Environmental factors support the formation of specific bacterial assemblages on microplastics. Front Microbiol. 2018;8:2709.
    PubMed  PubMed Central  Google Scholar 

    40.
    Kirstein IV, Wichels A, Krohne G, Gerdts G. Mature biofilm communities on synthetic polymers in seawater-specific or general? Mar Environ Res. 2018;142:147–54.
    CAS  PubMed  Google Scholar 

    41.
    Kettner MT, Rojas‐Jimenez K, Oberbeckmann S, Labrenz M, Grossart HP. Microplastics alter composition of fungal communities in aquatic ecosystems. Environ Microbiol. 2017;19:4447–59.
    CAS  PubMed  Google Scholar 

    42.
    Kettner MT, Oberbeckmann S, Labrenz M, Grossart HP. The eukaryotic life on microplastics in brackish ecosystems. Front Microbiol. 2019;10:538.
    PubMed  PubMed Central  Google Scholar 

    43.
    Bayoudh S, Othmane A, Bettaieb F, Bakhrouf A, Ouada HB, Ponsonnet L. Quantification of the adhesion free energy between bacteria and hydrophobic and hydrophilic substrata. Mater Sci Eng C. 2006;26:300–5.
    CAS  Google Scholar 

    44.
    Bendinger B, Rijnaarts HH, Altendorf K, Zehnder AJ. Physicochemical cell surface and adhesive properties of coryneform bacteria related to the presence and chain length of mycolic acids. Appl Environ Microbiol. 1993;59:3973–7.
    CAS  PubMed  PubMed Central  Google Scholar 

    45.
    Thompson SE, Coates JC. Surface sensing and stress-signalling in Ulva and fouling diatoms–potential targets for antifouling: a review. Biofouling. 2017;33:410–32.
    PubMed  Google Scholar 

    46.
    Araya P, Chamy R, Mota M, Alves M. Biodegradability and toxicity of styrene in the anaerobic digestion process. Biotechnol Lett. 2000;22:1477–81.
    CAS  Google Scholar 

    47.
    Pinto M, Langer TM, Hüffer T, Hofmann T, Herndl GJ. The composition of bacterial communities associated with plastic biofilms differs between different polymers and stages of biofilm succession. PLoS ONE. 2019;14:e0217165.
    CAS  PubMed  PubMed Central  Google Scholar 

    48.
    Datta MS, Sliwerska E, Gore J, Polz MF, Cordero OX. Microbial interactions lead to rapid micro-scale successions on model marine particles. Nat Commun. 2016;7:11965.
    CAS  PubMed  PubMed Central  Google Scholar 

    49.
    Zobell CE. The effect of solid surfaces upon bacterial activity. J Bacteriol. 1943;46:39–56.
    CAS  PubMed  PubMed Central  Google Scholar 

    50.
    Karl DM, Björkman KM, Dore JE, Fujieki L, Hebel DV, Houlihan T, et al. Ecological nitrogen-to-phosphorus stoichiometry at station ALOHA. Deep Sea Res Part II: Topical Stud Oceanogr. 2001;48:1529–66.
    CAS  Google Scholar 

    51.
    Steinberg DK, Carlson CA, Bates NR, Johnson RJ, Michaels AF, Knap AH. Overview of the US JGOFS Bermuda Atlantic Time-series Study (BATS): a decade-scale look at ocean biology and biogeochemistry. Deep Sea Res II. 2001;48:1405–47.
    CAS  Google Scholar 

    52.
    Flemming H-C, Wuertz S. Bacteria and Archaea on Earth and their abundance in biofilms. Nat Rev Microbiol. 2019;17:247.
    CAS  PubMed  Google Scholar 

    53.
    Whitman WB, Coleman DC, Wiebe WJ. Prokaryotes: the unseen majority. Proc Natl Acad Sci. 1998;95:6578–83.
    CAS  PubMed  Google Scholar 

    54.
    Bjørnsen PK. Automatic determination of bacterioplankton biomass by image analysis. Appl Environ Microbiol. 1986;51:1199–204.
    PubMed  PubMed Central  Google Scholar 

    55.
    Bloem J, Veninga M, Shepherd J. Fully automatic determination of soil bacterium numbers, cell volumes, and frequencies of dividing cells by confocal laser scanning microscopy and image analysis. Appl Environ Microbiol. 1995;61:926–36.
    CAS  PubMed  PubMed Central  Google Scholar 

    56.
    Kallmeyer J, Pockalny R, Adhikari RR, Smith DC, D’Hondt S. Global distribution of microbial abundance and biomass in subseafloor sediment. Proc Natl Acad Sci. 2012;109:16213–6.
    CAS  PubMed  Google Scholar 

    57.
    Pernice MC, Forn I, Gomes A, Lara E, Alonso-Sáez L, Arrieta JM, et al. Global abundance of planktonic heterotrophic protists in the deep ocean. ISME J. 2015;9:782–92.
    CAS  PubMed  Google Scholar 

    58.
    Bölter M, Bloem J, Meiners K, Möller R. Enumeration and biovolume determination of microbial cells–a methodological review and recommendations for applications in ecological research. Biol Fertil Soils. 2002;36:249–59.
    Google Scholar  More

  • in

    A new wave of marine fish invasions through the Panama and Suez canals

    1.
    Liu, X. et al. Curr. Biol. 29, 499–505.e4 (2019).
    CAS  Article  Google Scholar 
    2.
    Sardain, A. et al. Nat. Sustain. 2, 274–282 (2019).
    Article  Google Scholar 

    3.
    Review of Maritime Transport 2019 (United Nations, 2019).

    4.
    Leigh, E. G. et al. Biol. Rev. 89, 148–172 (2014).
    Article  Google Scholar 

    5.
    Seebens, H. et al. Proc. Natl Acad. Sci. USA 113, 5646–5651 (2016).
    CAS  Article  Google Scholar 

    6.
    Galil, B. et al. Manag. Biol. Invasion 8, 141–152 (2017).
    Article  Google Scholar 

    7.
    Spanier, E. & Galil, B. S. Endeavour 15, 102–106 (1991).
    Article  Google Scholar 

    8.
    Ruiz, G. M. et al. Smithson. Contrib. Mar. Sci. 38, 73–93 (2009).
    Google Scholar 

    9.
    Muirhead, J. R. et al. Divers. Distrib. 21, 75–87 (2015).
    Article  Google Scholar 

    10.
    Galil, B. S. et al. Biol. Invasion 17, 973–976 (2015).
    Article  Google Scholar 

    11.
    Azzurro, E. et al. Biol. Invasion 18, 2761–2772 (2016).
    Article  Google Scholar 

    12.
    Sharpe, D. et al. Ecology 98, 412–424 (2017).
    CAS  Article  Google Scholar 

    13.
    Salgado, J. et al. Sci. Total Environ. 729, 138444 (2020).
    CAS  Article  Google Scholar 

    14.
    Informe sobre la Aplicación y Eficiencia de Medidas de Mitigación para el Estudio de Impacto Ambiental del Proyecto “Ampliación del Canal de Panamá -Tercer Juego de Esclusas” (Panama Canal Authority, accessed 23 April 2020); https://go.nature.com/2FvcWMF

    15.
    Miller, A. W. & Ruiz, G. M. Nat. Clim. Change 4, 413–416 (2014).
    Article  Google Scholar 

    16.
    Cramer, W. et al. Nat. Clim. Change 8, 972–980 (2018).
    Article  Google Scholar 

    17.
    Ballew, N. G. et al. Sci. Rep. 6, 32169 (2016).
    CAS  Article  Google Scholar 

    18.
    Savva, I. et al. J. Fish. Biol. 97, 148–162 (2020).
    Article  Google Scholar 

    19.
    Shine, C. EPPO Bull. 37, 103–113 (2007).
    Article  Google Scholar 

    20.
    Ballast Water Management (IMO, accessed 20 May 2020); https://go.nature.com/2DUkI2t

    21.
    GloFouling (IMO, accessed 20 May 2020); https://www.glofouling.imo.org/

    22.
    Jouffray, J.-B. et al. One Earth 2, 43–54 (2020).
    Article  Google Scholar 

    23.
    Peleg, O. & Guy-Haim, T. Nature 575, 287 (2019).
    CAS  Article  Google Scholar 

    24.
    Balasingham, K. D. et al. Mol. Ecol. 27, 112–127 (2018).
    CAS  Article  Google Scholar 

    25.
    Martignac, F. et al. Fish. Fish. 16, 486–510 (2014).
    Article  Google Scholar 

    26.
    Putland, R. L. & Mensinger, A. F. Rev. Fish. Biol. Fish. 29, 789–807 (2019).
    Article  Google Scholar 

    27.
    Dennis, C. E. III et al. Biol. Invasion 21, 2837–2855 (2019).
    Article  Google Scholar 

    28.
    Sepulveda, A. J. et al. Trends Ecol. Evol. 35, 668–678 (2020).
    Article  Google Scholar 

    29.
    van Rijn, I. et al. Can. J. Fish. Aquat. Sci. 77, 752–761 (2020).
    Article  Google Scholar  More

  • in

    American mastodon mitochondrial genomes suggest multiple dispersal events in response to Pleistocene climate oscillations

    1.
    Collins, M. et al. In Climate Change 2013—The Physical Science Basis (ed. Intergovernmental Panel on Climate Change) 1029–1136 (Cambridge University Press, Cambridge, 2013).
    2.
    Ackerly, D. D. et al. The geography of climate change: implications for conservation biogeography. Divers. Distrib. 16, 476–487 (2010).
    Google Scholar 

    3.
    Bradshaw, W. E. & Holzapfel, C. M. Evolutionary response to rapid climate change. Science 312, 1477–1478 (2006).
    CAS  PubMed  Google Scholar 

    4.
    Chu, C., Mandrak, N. E. & Minns, C. K. Potential impacts of climate change on the distributions of several common and rare freshwater fishes in Canada. Divers. Distrib. 11, 299–310 (2005).
    Google Scholar 

    5.
    Princé, K. & Zuckerberg, B. Climate change in our backyards: the reshuffling of North America’s winter bird communities. Glob. Change Biol. 21, 572–585 (2015).
    ADS  Google Scholar 

    6.
    Scheffers, B. R. et al. The broad footprint of climate change from genes to biomes to people. Science 354, aaf7671 (2016).
    PubMed  Google Scholar 

    7.
    Lisiecki, L. E. & Raymo, M. E. A Pliocene-Pleistocene stack of 57 globally distributed benthic δ 18 O records. Paleoceanography 20, PA1003 (2005).
    ADS  Google Scholar 

    8.
    Dyke, A. S. An outline of the deglaciation of North America with emphasis on central and northern Canada. Quat. Glaciat. Chronol. Part II 2b, 373–424 (2004).
    Google Scholar 

    9.
    Thompson, L. G. et al. Late glacial stage and Holocene tropical ice core records from Huascaran, Peru. Science 269, 46–50 (1995).
    ADS  CAS  Google Scholar 

    10.
    Johnsen, S. J. et al. Oxygen isotope and palaeotemperature records from six Greenland ice-core stations: camp century, dye-3, GRIP, GISP2, Renland and NorthGRIP. J. Quat. Sci. 16, 299–307 (2001).
    Google Scholar 

    11.
    Kawamura, K. et al. Northern Hemisphere forcing of climatic cycles in Antarctica over the past 360,000 years. Nature 448, 912–916 (2007).
    ADS  CAS  PubMed  Google Scholar 

    12.
    Dyke, A. S. Late quaternary vegetation history of Northern North America based on pollen, macrofossil, and faunal remains. Géogr. Phys. Quat. 59, 211–262 (2005).
    Google Scholar 

    13.
    Froese, D. et al. Fossil and genomic evidence constrains the timing of bison arrival in North America. Proc. Natl Acad. Sci. USA 114, 3457–3462 (2017).
    ADS  CAS  PubMed  Google Scholar 

    14.
    Palkopoulou, E. et al. Holarctic genetic structure and range dynamics in the woolly mammoth. Proc. R. Soc. B Biol. Sci. 280, 20131910 (2013).
    Google Scholar 

    15.
    Debruyne, R. et al. Out of America: ancient DNA evidence for a new world origin of late quaternary woolly mammoths. Curr. Biol. 18, 1320–1326 (2008).
    CAS  PubMed  Google Scholar 

    16.
    Shapiro, B. et al. Rise and fall of the Beringian Steppe Bison. Science 306, 1561–1565 (2004).
    ADS  CAS  PubMed  Google Scholar 

    17.
    Campos, P. F. et al. Ancient DNA analyses exclude humans as the driving force behind late Pleistocene musk ox (Ovibos moschatus) population dynamics. Proc. Natl Acad. Sci. USA 107, 5675–5680 (2010).
    ADS  CAS  PubMed  Google Scholar 

    18.
    Chang, D. et al. The evolutionary and phylogeographic history of woolly mammoths: a comprehensive mitogenomic analysis. Sci. Rep. 7, 44585 (2017).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    19.
    Heintzman, P. D. et al. Bison phylogeography constrains dispersal and viability of the ice free corridor in western Canada. Proc. Natl Acad. Sci. USA 113, 8057–8063 (2016).
    CAS  PubMed  Google Scholar 

    20.
    Zazula, G. D. et al. American mastodon extirpation in the Arctic and Subarctic predates human colonization and terminal Pleistocene climate change. Proc. Natl Acad. Sci. USA 2014, 6–11 (2014).
    Google Scholar 

    21.
    Zazula, G. D. et al. A case of early Wisconsinan “over-chill”: New radiocarbon evidence for early extirpation of western camel (Camelops hesternus) in eastern Beringia. Quat. Sci. Rev. 171, 48–57 (2017).
    ADS  Google Scholar 

    22.
    Saunders, J. J. et al. Paradigms and proboscideans in the southern Great Lakes region, USA. Quat. Int. 217, 175–187 (2010).
    Google Scholar 

    23.
    Oltz, D. F. & Kapp, R. O. Plant remains associated with Mastodon and Mammoth remains in central Michigan. Am. Midl. Nat. 70, 339–346 (1963).
    Google Scholar 

    24.
    Dreimanis, A. Extinction of Mastodons in Eastern North America: testing a new climatic-environmental hypothesis. Ohio J. Sci. 68, 257–272 (1968).
    Google Scholar 

    25.
    Shoshani, J. Understanding proboscidean evolution: a formidable task. Trends Ecol. Evol. 13, 480–487 (1998).
    CAS  PubMed  Google Scholar 

    26.
    Teale, C. L. & Miller, N. G. Mastodon herbivory in mid-latitude late-Pleistocene boreal forests of eastern North America. Quat. Res. 78, 72–81 (2012).
    Google Scholar 

    27.
    Green, J. L., DeSantis, L. R. G. & Smith, G. J. Regional variation in the browsing diet of Pleistocene Mammut americanum (Mammalia, Proboscidea) as recorded by dental microwear textures. Palaeogeogr. Palaeoclimatol. Palaeoecol. 487, 59–70 (2017).
    Google Scholar 

    28.
    Birks, H. H. et al. Evidence for the diet and habitat of two late Pleistocene mastodons from the Midwest, USA. Quat. Res. 91, 792–812 (2019).
    CAS  Google Scholar 

    29.
    Owen-Smith, N. Pleistocene extinctions: the pivotal role of megaherbivores. Paleobiology 13, 351–362 (1987).
    Google Scholar 

    30.
    Barnosky, A. D. et al. Variable impact of late-quaternary megafaunal extinction in causing ecological state shifts in North and South America. Proc. Natl Acad. Sci. USA 113, 856–861 (2016).
    ADS  CAS  PubMed  Google Scholar 

    31.
    Widga, C. et al. Late pleistocene proboscidean population dynamics in the North American midcontinent. Boreas 46, 772–782 (2017).
    Google Scholar 

    32.
    Godfrey-Smith, D., Grist, A. & Stea, R. Dosimetric and radiocarbon chronology of a pre-Wisconsinan mastodon fossil locality at East Milford, Nova Scotia, Canada. Quat. Sci. Rev. 22, 1353–1360 (2003).
    ADS  Google Scholar 

    33.
    Enk, J. et al. Mammuthus population dynamics in late pleistocene North America: divergence, phylogeogrpaphy and introgression. Front. Ecol. Evol. 4, 1–13 (2016).
    Google Scholar 

    34.
    Ishida, Y., Georgiadis, N. J., Hondo, T. & Roca, A. L. Triangulating the provenance of African elephants using mitochondrial DNA. Evol. Appl. 6, 253–265 (2013).
    CAS  PubMed  Google Scholar 

    35.
    Fernando, P., Pfrender, M. E., Encalada, S. E. & Lande, R. Mitochondrial DNA variation, phylogeography and population structure of the Asian elephant. Heredity 84, 362–372 (2000).
    CAS  PubMed  Google Scholar 

    36.
    Fisher, D. In The Proboscidea: Evolution and Paleoecology of Elephants andtheir Relatives (eds. Shoshani, J. & Tassy, P.) 296–315 (Oxford University Press, Oxford, 1996).

    37.
    Fisher, D. C. Paleobiology of pleistocene proboscideans. Annu. Rev. Earth Planet. Sci. https://doi.org/10.1146/annurev-earth-060115-012437 (2018).

    38.
    Rohland, N. et al. Genomic DNA sequences from mastodon and woolly mammoth reveal deep speciation of forest and savanna elephants. PLoS Biol. 8, e1000564 (2010).
    CAS  PubMed  PubMed Central  Google Scholar 

    39.
    Muhs, D. R., Ager, T. A. & Begét, J. E. Vegetation and paleoclimate of the last interglacial period, central Alaska. Quat. Sci. Rev. 20, 41–61 (2001).
    ADS  Google Scholar 

    40.
    Jass, C. N. & Barrón-Ortiz, C. I. A review of quaternary proboscideans from Alberta, Canada. Quat. Int. 443, 88–104 (2017).
    Google Scholar 

    41.
    Shapiro, B. et al. A Bayesian phylogenetic method to estimate unknown sequence ages. Mol. Biol. Evol. 28, 879–887 (2011).
    CAS  PubMed  Google Scholar 

    42.
    Drummond, A. J. & Stadler, T. Bayesian phylogenetic estimation of fossil ages. Philos. Trans. R. Soc. B Biol. Sci. 371, 20150129 (2016).
    Google Scholar 

    43.
    Plint, T., Longstaffe, F. J. & Zazula, G. Giant beaver palaeoecology inferred from stable isotopes. Sci. Rep. 9, 7179 (2019).
    ADS  PubMed  PubMed Central  Google Scholar 

    44.
    Yalden, D. W. The history of British mammals 12–27 (T & A D Poyser Ltd, Berkhamsted, 1999).

    45.
    Schreve, D. C. A new record of Pleistocene hippopotamus from River Severn terrace deposits, Gloucester, UK—palaeoenvironmental setting and stratigraphical significance. Proc. Geol. Assoc. 120, 58–64 (2009).
    Google Scholar 

    46.
    Stoffel, C. et al. Genetic consequences of population expansions and contractions in the common hippopotamus (Hippopotamus amphibius) since the late Pleistocene. Mol. Ecol. 24, 2507–2520 (2015).
    PubMed  Google Scholar 

    47.
    Tape, K. D., Gustine, D. D., Ruess, R. W., Adams, L. G. & Clark, J. A. Range expansion of moose in Arctic Alaska linked to warming and increased shrub habitat. PLoS ONE 11, e0152636 (2016).
    PubMed  PubMed Central  Google Scholar 

    48.
    Tape, K. D., Jones, B. M., Arp, C. D., Nitze, I. & Grosse, G. Tundra be dammed: beaver colonization of the Arctic. Glob. Change Biol. 24, 4478–4488 (2018).
    ADS  Google Scholar 

    49.
    Dabney, J. et al. Complete mitochondrial genome sequence of a Middle Pleistocene cave bear reconstructed from ultrashort DNA fragments. Proc. Natl Acad. Sci. USA 110, 15758–15763 (2013).
    ADS  CAS  PubMed  Google Scholar 

    50.
    Glocke, I. & Meyer, M. Extending the spectrum of DNA sequences retrieved from ancient bones and teeth. Genome Res. 27, 1–8 (2017).
    Google Scholar 

    51.
    Kircher, M., Sawyer, S. & Meyer, M. Double indexing overcomes inaccuracies in multiplex sequencing on the Illumina platform. Nucleic Acids Res. 40, 1–8 (2012).
    Google Scholar 

    52.
    Meyer, M. & Kircher, M. Illumina sequencing library preparation for highly multiplexed target capture and sequencing. Cold Spring Harb. Protoc. 2010, 1–10 (2010).
    Google Scholar 

    53.
    Gansauge, M.-T. & Meyer, M. Single-stranded DNA library preparation for the sequencing of ancient or damaged DNA. Nat. Protoc. 8, 737–748 (2013).
    PubMed  Google Scholar 

    54.
    Gansauge, M.-T. et al. Single-stranded DNA library preparation from highly degraded DNA using T4 DNA ligase. Nucleic Acids Res. 45, 1–10 (2017).
    Google Scholar 

    55.
    Renaud, G., Stenzel, U. & Kelso, J. leeHom: adaptor trimming and merging for Illumina sequencing reads. Nucleic Acids Res. https://doi.org/10.1093/nar/gku699 (2014).

    56.
    Li, H. & Durbin, R. Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics 25, 1754–1760 (2009).
    CAS  PubMed  PubMed Central  Google Scholar 

    57.
    Edgar, R. C. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 32, 1792–1797 (2004).
    CAS  PubMed  PubMed Central  Google Scholar 

    58.
    Darriba, D., Taboada, G. L., Doallo, R. & Posada, D. jModelTest 2: more models, new heuristics and parallel computing. Nat. Methods 9, 772–772 (2012).
    CAS  PubMed  PubMed Central  Google Scholar 

    59.
    Nguyen, L. T., Schmidt, H. A., Von Haeseler, A. & Minh, B. Q. IQ-TREE: a fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol. Biol. Evol. 32, 268–274 (2015).
    CAS  PubMed  PubMed Central  Google Scholar 

    60.
    Drummond, A. J., Suchard, M. A., Xie, D. & Rambaut, A. Bayesian phylogenetics with BEAUti and the BEAST 1.7. Mol. Biol. Evol. 29, 1969–1973 (2012).
    CAS  PubMed  PubMed Central  Google Scholar 

    61.
    Baele, G., Lemey, P. & Suchard, M. A. Genealogical working distributions for Bayesian model testing with phylogenetic uncertainty. Syst. Biol. 65, 250–264 (2016).
    PubMed  Google Scholar 

    62.
    Suchard, M. A. et al. Bayesian phylogenetic and phylodynamic data integration using BEAST 1.10. Virus Evol. 4, vey016 (2018).
    PubMed  PubMed Central  Google Scholar 

    63.
    Stuiver, M. & Reimer, P. J. Extended 14C database and revised CALIB radiocarbon calibration program. Radiocarbon 35, 215–230 (1993).
    Google Scholar 

    64.
    Paradis, E., Claude, J. & Strimmer, K. APE: analyses of phylogenetics and evolution in R language. Bioinformatics 20, 289–290 (2004).
    CAS  Google Scholar 

    65.
    Colleoni, F., Wekerle, C., Näslund, J.-O., Brandefelt, J. & Masina, S. Constraint on the penultimate glacial maximum Northern Hemisphere ice topography (≈140 kyrs BP). Quat. Sci. Rev. 137, 97–112 (2016).
    ADS  Google Scholar  More

  • in

    Host age is not a consistent predictor of microbial diversity in the coral Porites lutea

    1.
    Pootakham, W. et al. Dynamics of coral-associated microbiomes during a thermal bleaching event. Microbiologyopen 7, e00604 (2018).
    PubMed  PubMed Central  Google Scholar 
    2.
    Krediet, C. J., Ritchie, K. B., Paul Valerie, J. & Max, T. Coral-associated micro-organisms and their roles in promoting coral health and thwarting diseases. Proc. R. Soc. B Biol. Sci. 280, 20122328 (2013).
    Google Scholar 

    3.
    Ziegler, M., Seneca, F. O., Yum, L. K., Palumbi, S. R. & Voolstra, C. R. Bacterial community dynamics are linked to patterns of coral heat tolerance. Nat. Commun. 8, 14213 (2017).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    4.
    Rädecker, N., Pogoreutz, C., Voolstra, C. R., Wiedenmann, J. & Wild, C. Nitrogen cycling in corals: The key to understanding holobiont functioning?. Trends Microbiol. 23, 490–497 (2015).
    PubMed  Google Scholar 

    5.
    Ritchie, K. B. & Smith, G. W. Microbial communities of coral surface mucopolysaccharide layers. In Coral Health and Disease (eds Rosenberg, E. & Loya, Y.) 259–264 (Springer, Berlin Heidelberg, 2004).
    Google Scholar 

    6.
    Holm, J. B. & Heidelberg, K. B. Microbiomes of Muricea californica and M. fruticosa: Comparative analyses of two co-occurring eastern pacific octocorals. Front. Microbiol. 7, 917 (2016).
    PubMed  PubMed Central  Google Scholar 

    7.
    Sweet, M. J., Brown, B. E., Dunne, R. P., Singleton, I. & Bulling, M. Evidence for rapid, tide-related shifts in the microbiome of the coral Coelastrea aspera. Coral Reefs 36, 815–828 (2017).
    ADS  Google Scholar 

    8.
    Ziegler, M. et al. Coral microbial community dynamics in response to anthropogenic impacts near a major city in the central Red Sea. Mar. Pollut. Bull. 105, 629–640 (2016).
    CAS  PubMed  Google Scholar 

    9.
    Archer, S. D. J. et al. Air mass source determines airborne microbial diversity at the ocean–atmosphere interface of the Great Barrier Reef marine ecosystem. ISME J. https://doi.org/10.1038/s41396-019-0555-0 (2019).
    Article  PubMed  PubMed Central  Google Scholar 

    10.
    Wainwright, B. J., Afiq-Rosli, L., Zahn, G. L. & Huang, D. Characterisation of coral-associated bacterial communities in an urbanised marine environment shows strong divergence over small geographic scales. Coral Reefs https://doi.org/10.1007/s00338-019-01837-1 (2019).
    Article  Google Scholar 

    11.
    Chu, N. D. & Vollmer, S. V. Caribbean corals house shared and host-specific microbial symbionts over time and space. Environ. Microbiol. Rep. 8, 493–500 (2016).
    CAS  PubMed  Google Scholar 

    12.
    Wainwright, B. J., Bauman, A. G., Zahn, G. L., Todd, P. A. & Huang, D. Characterization of fungal biodiversity and communities associated with the reef macroalga Sargassum ilicifolium reveals fungal community differentiation according to geographic locality and algal structure. Mar. Biodivers. https://doi.org/10.1007/s12526-019-00992-6 (2019).
    Article  Google Scholar 

    13.
    Wainwright, B. J., Zahn, G. L., Arlyza, I. S. & Amend, A. S. Seagrass-associated fungal communities follow Wallace’s line, but host genotype does not structure fungal community. J. Biogeogr. 45, 762–770 (2018).
    Google Scholar 

    14.
    Hernandez-Agreda, A., Leggat, W., Bongaerts, P., Herrera, C. & Ainsworth, T. D. Rethinking the coral microbiome: Simplicity exists within a diverse microbial biosphere. mBio 9, e00812 (2018).
    PubMed  PubMed Central  Google Scholar 

    15.
    Williams, A. D., Brown, B. E., Putchim, L. & Sweet, M. J. Age-related shifts in bacterial diversity in a reef coral. PLoS ONE 10, e0144902 (2015).
    PubMed  PubMed Central  Google Scholar 

    16.
    Pollock, F. J. et al. Coral-associated bacteria demonstrate phylosymbiosis and cophylogeny. Nat. Commun. 9, 4921 (2018).
    ADS  PubMed  PubMed Central  Google Scholar 

    17.
    Epstein, H. E., Torda, G., Munday, P. L. & van Oppen, M. J. H. Parental and early life stage environments drive establishment of bacterial and dinoflagellate communities in a common coral. ISME J. 13, 1635–1638 (2019).
    CAS  PubMed  PubMed Central  Google Scholar 

    18.
    Yatsunenko, T. et al. Human gut microbiome viewed across age and geography. Nature 486, 222–227 (2012).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    19.
    van Dongen, W. F. et al. Age-related differences in the cloacal microbiota of a wild bird species. BMC Ecol. 13, 11 (2013).
    PubMed  PubMed Central  Google Scholar 

    20.
    Huang, D. et al. Extraordinary diversity of reef corals in the South China Sea. Mar. Biodivers. 45, 157–168 (2015).
    Google Scholar 

    21.
    Toda, T. et al. Community structures of coral reefs around Peninsular Malaysia. J. Oceanogr. 63, 113–123 (2007).
    Google Scholar 

    22.
    Tanzil, J. T. I. et al. Regional decline in growth rates of massive Porites corals in Southeast Asia. Glob. Change Biol. 19, 3011–3023 (2013).
    ADS  Google Scholar 

    23.
    Tanzil, J. T. I. et al. Luminescence and density banding patterns in massive Porites corals around the Thai-Malay Peninsula, Southeast Asia. Limnol. Oceanogr. 61, 2003–2026 (2016).
    ADS  Google Scholar 

    24.
    Pootakham, W. et al. High resolution profiling of coral-associated bacterial communities using full-length 16S rRNA sequence data from PacBio SMRT sequencing system. Sci. Rep. 7, 2774 (2017).
    ADS  PubMed  PubMed Central  Google Scholar 

    25.
    Øvreås, L., Daae, F. L., Torsvik, V. & Rodríguez-Valera, F. Characterization of microbial diversity in hypersaline environments by melting profiles and reassociation kinetics in combination with terminal restriction fragment length polymorphism (T-RFLP). Microb. Ecol. 46, 291–301 (2003).
    PubMed  Google Scholar 

    26.
    Baker, B. J. & Banfield, J. F. Microbial communities in acid mine drainage. FEMS Microbiol. Ecol. 44, 139–152 (2003).
    CAS  PubMed  Google Scholar 

    27.
    Li, S.-J. et al. Microbial communities evolve faster in extreme environments. Sci. Rep. 4, 6205 (2014).
    CAS  PubMed  PubMed Central  Google Scholar 

    28.
    Peter, J. et al. A microbial signature of psychological distress in irritable bowel syndrome. Psychosom. Med. 80, 698–709 (2018).
    PubMed  PubMed Central  Google Scholar 

    29.
    Karl, J. P. et al. Effects of psychological, environmental and physical stressors on the gut microbiota. Front. Microbiol. https://doi.org/10.3389/fmicb.2018.02013 (2018).
    Article  PubMed  PubMed Central  Google Scholar 

    30.
    Guest, J. R. et al. 27 years of benthic and coral community dynamics on turbid, highly urbanised reefs off Singapore. Sci. Rep. 6, 36260 (2016).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    31.
    Wong, J. S. Y. et al. Comparing patterns of taxonomic, functional and phylogenetic diversity in reef coral communities. Coral Reefs 37, 737–750 (2018).
    ADS  MathSciNet  Google Scholar 

    32.
    Chow, G. S. E., Chan, Y. K. S., Jain, S. S. & Huang, D. Light limitation selects for depth generalists in urbanised reef coral communities. Mar. Environ. Res. 147, 101–112 (2019).
    CAS  PubMed  Google Scholar 

    33.
    Calvani, R. et al. Of microbes and minds: A narrative review on the second brain aging. Front. Med. (Lausanne) https://doi.org/10.3389/fmed.2018.00053 (2018).
    Article  Google Scholar 

    34.
    Nagpal, R. et al. Gut microbiome and aging: Physiological and mechanistic insights. Nutr Healthy Aging 4, 267–285 (2018).
    PubMed  PubMed Central  Google Scholar 

    35.
    Choi, J., Hur, T.-Y. & Hong, Y. Influence of altered gut microbiota composition on aging and aging-related diseases. J. Lifestyle Med. 8, 1–7 (2018).
    PubMed  PubMed Central  Google Scholar 

    36.
    Soong, K., Chen, C. A. & Chang, J.-C. A very large poritid colony at Green Island, Taiwan. Coral Reefs 18, 42–42 (1999).
    Google Scholar 

    37.
    Goodkin, N. et al. Coral communities of Hong Kong: Long-lived corals in a marginal reef environment. Mar. Ecol. Prog. Ser. 426, 185–196 (2011).
    ADS  Google Scholar 

    38.
    Bythell, J. C., Brown, B. E. & Kirkwood, T. B. L. Do reef corals age?. Biol. Rev. 93, 1192–1202 (2018).
    PubMed  Google Scholar 

    39.
    Lee, N. L. Y., Huang, D., Quek, Z. B. R., Lee, J. N. & Wainwright, B. J. Mangrove-associated fungal communities are differentiated by geographic location and host structure. Front. Microbiol. https://doi.org/10.3389/fmicb.2019.02456 (2019).
    Article  PubMed  PubMed Central  Google Scholar 

    40.
    Wainwright, B. J. et al. Seagrass-associated fungal communities show distance decay of similarity that has implications for seagrass management and restoration. Ecol. Evol. 9, 11288–11297 (2019).
    PubMed  PubMed Central  Google Scholar 

    41.
    Röthig, T., Ochsenkühn, M. A., Roik, A., van der Merwe, R. & Voolstra, C. R. Long-term salinity tolerance is accompanied by major restructuring of the coral bacterial microbiome. Mol. Ecol. 25, 1308–1323 (2016).
    PubMed  PubMed Central  Google Scholar 

    42.
    Sin, T. M. et al. The urban marine environment of Singapore. Region. Stud. Mar. Sci. 8, 331–339 (2016).
    Google Scholar 

    43.
    Chénard, C. et al. Temporal and spatial dynamics of bacteria, Archaea and protists in equatorial coastal waters. Sci. Rep. 9, 1–13 (2019).
    Google Scholar 

    44.
    Ford, A. K. et al. Reefs under Siege—The rise, putative drivers, and consequences of benthic cyanobacterial mats. Front. Mar. Sci. https://doi.org/10.3389/fmars.2018.00018 (2018).
    Article  Google Scholar 

    45.
    Charpy, L., Casareto, B. E., Langlade, M. J. & Suzuki, Y. Cyanobacteria in coral reef ecosystems: A review. J. Mar. Biol. 2012, 1–9 (2012).
    Google Scholar 

    46.
    Huang, D., Tun, K., Chou, L. M. & Todd, P. A. An inventory of zooxanthellate scleractinian corals in Singapore, including 33 new records. Raffles Bull. Zool. Suppl. 22, 69 (2009).
    CAS  Google Scholar 

    47.
    Todd, P. A. et al. Towards an urban marine ecology: Characterizing the drivers, patterns and processes of marine ecosystems in coastal cities. Oikos https://doi.org/10.1111/oik.05946 (2019).
    Article  Google Scholar 

    48.
    Rubin, B. E. R. et al. Investigating the impact of storage conditions on microbial community composition in soil samples. PLoS ONE 8, e70460 (2013).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    49.
    Lauber, C. L., Zhou, N., Gordon, J. I., Knight, R. & Fierer, N. Effect of storage conditions on the assessment of bacterial community structure in soil and human-associated samples: Influence of short-term storage conditions on microbiota. FEMS Microbiol. Lett. 307, 80–86 (2010).
    CAS  PubMed  PubMed Central  Google Scholar 

    50.
    Carruthers, L. V. et al. The impact of storage conditions on human stool 16S rRNA microbiome composition and diversity. PeerJ 7, e8133 (2019).
    PubMed  PubMed Central  Google Scholar 

    51.
    Veron, J. Corals of the World (Australian Institute of Marine Science, Townsville, 2000).
    Google Scholar 

    52.
    Forsman, Z., Wellington, G. M., Fox, G. E. & Toonen, R. J. Clues to unraveling the coral species problem: Distinguishing species from geographic variation in Porites across the Pacific with molecular markers and microskeletal traits. PeerJ 3, e751 (2015).
    PubMed  PubMed Central  Google Scholar 

    53.
    Forsman, Z. H., Barshis, D. J., Hunter, C. L. & Toonen, R. J. Shape-shifting corals: Molecular markers show morphology is evolutionarily plastic in Porites. BMC Evol. Biol. 9, 45 (2009).
    PubMed  PubMed Central  Google Scholar 

    54.
    Terraneo, T. I. et al. Environmental latitudinal gradients and host-specificity shape Symbiodiniaceae distribution in Red Sea Porites corals. J. Biogeogr. https://doi.org/10.1111/jbi.13672 (2019).
    Article  Google Scholar 

    55.
    Caporaso, J. G. et al. Global patterns of 16S rRNA diversity at a depth of millions of sequences per sample. PNAS 108, 4516–4522 (2011).
    ADS  CAS  PubMed  Google Scholar 

    56.
    Lundberg, D. S., Yourstone, S., Mieczkowski, P., Jones, C. D. & Dangl, J. L. Practical innovations for high-throughput amplicon sequencing. Nat. Methods 10, 999–1002 (2013).
    CAS  PubMed  Google Scholar 

    57.
    Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet.journal 17, 10–12 (2011).
    Google Scholar 

    58.
    Callahan, B. J. et al. DADA2: High-resolution sample inference from Illumina amplicon data. Nat. Methods 13, 581–583 (2016).
    CAS  PubMed  PubMed Central  Google Scholar 

    59.
    Davis, N. M., Proctor, D. M., Holmes, S. P., Relman, D. A. & Callahan, B. J. Simple statistical identification and removal of contaminant sequences in marker-gene and metagenomics data. Microbiome 6, 226 (2018).
    PubMed  PubMed Central  Google Scholar 

    60.
    Cole, J. R. et al. The ribosomal database project (RDP-II): Introducing myRDP space and quality controlled public data. Nucleic Acids Res. 35, D169–D172 (2007).
    CAS  PubMed  Google Scholar 

    61.
    Quast, C. et al. The SILVA ribosomal RNA gene database project: Improved data processing and web-based tools. Nucleic Acids Res. 41, D590–D596 (2013).
    CAS  PubMed  PubMed Central  Google Scholar 

    62.
    Oksanen, J. et al. vegan: Community Ecology Package (2019).

    63.
    McMurdie, P. J. & Holmes, S. phyloseq: An R package for reproducible interactive analysis and graphics of microbiome census data. PLoS ONE 8, e61217 (2013).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    64.
    Martin, B. D., Witten, D. & Willis, A. D. Modeling microbial abundances and dysbiosis with beta-binomial regression. Ann. Appl. Stat. 14, 94–115 (2020).
    MathSciNet  MATH  Google Scholar  More

  • in

    Robotic environmental DNA bio-surveillance of freshwater health

    ESP sample processing
    The ESP operated autonomously, needing only power, communications and fluid connections through its waterproof pressure housing (Fig. 1). Prior to sample initiation, the ESP was purged completely with nitrogen to reduce oxidative reactions (i.e., DNA degradation) from occurring. At the initiation of sampling, a puck (Fig. 1A cutout) loaded with filter material was placed within a clamp. Valves open to the outside allowed a syringe to sequentially pull water through the puck. Once the target volume was filtered, or the filter was loaded with biomass (i.e., ‘clogged’), filtering stopped and excess water was cleared. Five mL’s of RNAlater preservative was then added to the puck, soaking the filter for 10 min before the excess was evacuated and the puck was returned to storage. Preserved pucks were stored at the ESP temperature, which were similar to ambient air temperatures. The upper limit on the amount of time that an ESP device can operate in the field before DNA quality on a puck is comprised is not known but is at least 21 days10. A constant humidity kept the pucks moist, allowing for easy filter removal once the instrument was recovered.
    Figure 1

    The ESP is an electro-mechanical robot that can autonomously filter and preserve samples. (A) About the size of a 50-gal barrel, the ESP carries 132 ‘pucks’ (inset), each designed to hold 25 mm filters. (B) The ESP installed in a USGS streamgage station. (C) Streamgage station showing tubing run (white pipe) that contained pump and tubing to deliver stream water to the ESP. The ESP communicated via cell phone, and was powered during the deployment via either line power or portable solar arrays. Photo credits: U.S. Geological Survey.

    Full size image

    To get water to the ESP, we designed an external sampling module from which the ESP drew water11. The sampling module was self-draining, and fed by a submersible pump (WSP-12 V-2 M, Waterra USA Inc., Bellingham, W, USA) installed approximately 0.5 to 2 m below the river water line at each deployment site. To reduce possible carry-over contamination, the sampling pumps, tubing and external sampling modules were flushed with river water for 10 min prior to every sample collection. The sampling port of the ESP itself was cleaned with 10% bleach and a 10% tween-20 solution between samples. At the end of each ESP deployment, pucks were manually removed and filters were aseptically recovered into 2.0 mL screw cap centrifuge tubes and stored at − 80 ºC until molecular analyses were performed.
    Field deployments
    We performed initial ESP feasibility studies in Yellowstone National Park (USA; Fig. 2) in September 2017. Here, our goal was to determine if the ESP could be used to sample DNA of the waterborne protozoa, Naegleria spp., from a freshwater river where these organisms had previously been detected using standard techniques12. We filled 60-L sterilized carboys with water from the confluence of the Boiling and Gardner rivers. Carboys were transported to a lab at Montana State University (Bozeman, Montana) and connected to ESP samplers via tubing and syringe pumps. Water was passed through each filter (5-µm Diapore filters) until the filter became clogged; six samples were filtered.
    Figure 2

    Map of ESP water sampling locations. The inset map shows the location of the Upper Yellowstone River and Upper Snake River in the United States. The larger map shows the sample site locations (filled red circles) on each river relative to Yellowstone National Park and Grand Teton National Park (outlined in green).

    Full size image

    We then integrated the ESPs into two USGS streamgages on the Yellowstone River in 2018 and one USGS streamgage on the Snake River in 2019, (Fig. 1B,C) where we tested for DNA of the fish pathogen, Tetracapsuloides bryosalmonae, the causative agent of salmonid fish Proliferative Kidney Disease (PKD). On the Yellowstone River, we installed ESPs at the streamgage near the upstream and downstream extents of a recent PKD outbreak13, USGS 06191500 Yellowstone River at Corwin Springs MT and USGS 06192500 Yellowstone River near Livingston MT, described below as Corwin Springs and Carters Bridge, respectively (Fig. 2). On the Snake River, we installed one ESP at the streamgage 1.5 km downstream of Palisades Reservoir near the upstream extent of a recent PKD outbreak, USGS 13032500 Snake River near Irwin ID. The ESP pucks were loaded with 1.2-µm cellulose nitrate filters. We ran two negative controls (1 L of molecular grade water) through the ESP prior to and at the conclusion of deployment to assess for contamination.
    Yellowstone River
    The ESPs were programmed to collect 1-L samples every 12 h, from Jul 24 to Aug 26 2018, and every 3 h from Aug 27 to Sep 7 2018. The average (± 1 SE) volume filtered per sample was 639 (± 11) mL, indicating that most filters clogged prior to reaching the 1-L target volume. Filter samples were collected at ambient air temperatures ranging from 9.6 to 35.8 °C ((overline{x})  = 18.9) at Carter’s Bridge and 8.3–29.0 °C ((overline{x})  = 17.1) at Corwin Springs. We compared T. bryosalmonae ESP detections to those from manually collected grab samples from shore (6, 250-mL samples per site filtered through 1.2-µm cellulose nitrate filters) collected at weekly frequencies for the entire length of the ESP deployments and at daily frequencies between Aug 27 and Aug 30. Thus, ESP and manual eDNA samples collected at different temporal intervals (3 h, 12 h or weekly) allowed us to evaluate the added value of higher frequency sampling.
    We also evaluated the utility of automated high frequency sampling to detect a new invasion by introducing novel DNA of Scomber japonicas (mackerel fish) 100 m upstream of each Yellowstone River streamgage. On Aug 27, we introduced 3 kg of canned S. japonicas 100 m upstream of the water sampling inlet for each ESP. S. japonicas was blended with water, frozen and then placed within metal-wire minnow traps and anchored to the river’s bottom with cement pavers. The ESPs were programmed to sample every 3 h from Aug 27 to Sep 7. Manual grab samples (600 mL) were collected 10 m (n = 3), 100 m (n = 6), and 400 m (n = 3) downstream of the S. japonicas in order to test that S. japonicas DNA was transported downstream past the water sampling inlet of each ESP. Manual grab samples were collected immediately prior to S. japonicas introductions, 3 h post-introduction and then every 24 h for 3 days.
    Snake River
    The ESPs were programmed to collect 2-L samples every 12 h from Jul 17 to Sep 09 and then every 4 h from Sep 10 to Oct 1, 2019. Manually collected grab samples (three, 2-L samples filtered through 1.5-µm glass fiber filters) and negative field controls (1, 2-L sample of deionized water filtered through 1.5-µm glass fiber filters) were collected every 2 weeks following methods in Sepulveda et al.7. Filter samples were collected at ambient air temperatures ranging from 3.9 to 30.2 °C ((overline{x})  = 20.6). To broaden our taxonomic assessment, we tested these samples for T. bryosalmonae DNA, and also for kokanee salmon (Oncorhynchus nerka) and dreissenid mussel (Dreissena spp.) DNA. O. nerka only occur upstream in Palisades Reservoir and at such low abundances that they are not captured by resource managers in annual population surveys7. Dreissenid mussels have not yet been observed, but are the principal focus of aquatic invasive species monitoring programs in this region7.
    Molecular analyses
    Filters were removed from the pucks and then shipped frozen to the USGS Upper Midwest Environmental Science Center (LaCrosse, Wisconsin) for DNA extraction and quantitative PCR analyses. Filters were handled and stored in a dedicated room that is physically separated from rooms where high-quantity DNA extraction and PCR product or high-quality DNA is handled. We used the FastDNA SPIN kit for soil to extract DNA on samples from the Boiling River-Gardiner River confluence, following modifications described in Barnhart et al.14. To extract DNA from Yellowstone River and Snake River samples, we used the Investigator Lyse & Spin Basket Kit (Qiagen, Hilden, Germany) in concert with the gMax Mini genomic DNA kit (IBI Scientific), following manufacturer’s instructions, and eluted in 200 µL of buffer. Samples were extracted as site specific batches and one extraction control was collected per batch. We used previously published assays, limits of detection and methods therein for analyses of Naegleria spp.12, T. bryosalmonae13, S. japonicas15, O. nerka7, and Dreissena spp.16 (Table 1).
    Table 1 Primers and probes used in this study.
    Full size table

    We analyzed all samples in four replicate 25 µL reactions containing 2 µL of template DNA, 1 × Perfecta Toughmix (Quantabio), 400 nM forward and reverse primers, and 100 nM probe. Each plate contained 10 no-template PCR controls (one for each sample) using 2 µL of molecular grade water as the template as well as a standard curve with two replicates of 20,000 and 2,000 copy standards and four replicates of 200 and 20 copy standards. The standards were prepared with synthetic gBlocks (Integrated DNA Technologies) containing the amplicon sequences for each assay. Each sample was also analyzed in three replicates with 200 copies of synthetic gBlock spiked in to check for PCR inhibition. Any sample that indicated less than an average of 60 to 70 copies of targeted DNA in these triplicate samples was considered inhibited. Field and extraction negative controls were analyzed as regular samples. No negative controls amplified.
    Analyses
    Samples were scored as positive when one or more PCR replicates amplified for the target DNA. We used McNemar’s Exact Test to compare binary qPCR data (detection/non-detection) of T. bryosalmonae and O. nerka DNA between ESP and manually collected samples in the Yellowstone and Snake rivers. More

  • in

    Hunting strategies to increase detection of chronic wasting disease in cervids

    1.
    Wasserberg, G., Osnas, E. E., Rolley, R. E. & Samuel, M. D. Host culling as an adaptive management tool for chronic wasting disease in white-tailed deer: a modelling study. J. Appl. Ecol. 46, 457–466 (2009).
    PubMed  Google Scholar 
    2.
    Heberlein, T. A. “Fire in the Sistine Chapel”: How Wisconsin responded to chronic wasting disease. Hum. Dimens Wildl. 9, 165–179 (2004).
    Google Scholar 

    3.
    Donnelly, C. A. & Woodroffe, R. Badger-cull targets unlikely to reduce TB. Nature 526, 640 (2015).
    ADS  CAS  PubMed  Google Scholar 

    4.
    Turner, W. C. et al. Fatal attraction: vegetation responses to nutrient inputs attract herbivores to infectious anthrax carcass sites. Proc. R. Soc. Lond. Ser. B 281, https://doi.org/10.1098/rspb.2014.1785 (2014).

    5.
    Uehlinger, F. D., Johnston, A. C., Bollinger, T. K. & Waldner, C. L. Systematic review of management strategies to control chronic wasting disease in wild deer populations in North America. BMC Vet. Res. 12, 1–16 (2016).
    Google Scholar 

    6.
    Tildesley, M. J., Bessell, P. R., Keeling, M. J. & Woolhouse, M. E. J. The role of pre-emptive culling in the control of foot-and-mouth disease. Proc. R. Soc. Lond. Ser. B 276, 3239 (2009).
    Google Scholar 

    7.
    te Beest, D. E., Hagenaars, T. J., Stegeman, J. A., Koopmans, M. P. & van Boven, M. Risk based culling for highly infectious diseases of livestock. Vet. Res. 42, 81 (2011).
    Google Scholar 

    8.
    Benestad, S. L., Mitchell, G., Simmons, M., Ytrehus, B. & Vikøren, T. First case of chronic wasting disease in Europe in a Norwegian free-ranging reindeer. Vet. Res. 47, 88 (2016).
    PubMed  PubMed Central  Google Scholar 

    9.
    Haley, N. J. & Hoover, E. A. Chronic wasting disease of cervids: current knowledge and future perspectives. Annu. Rev. Anim. Biosci. 3, 305–325 (2015).
    CAS  PubMed  Google Scholar 

    10.
    USGS. Expanding Distribution of Chronic Wasting Disease https://www.usgs.gov/centers/nwhc/science/expanding-distribution-chronic-wasting-disease?qt-science_center_objects=0#qt-science_center_objects (USGS, 2019).

    11.
    Edmunds, D. R. et al. Chronic wasting disease drives population decline of white-tailed deer. PLoS ONE 11, e0161127 (2016).
    PubMed  PubMed Central  Google Scholar 

    12.
    DeVivo, M. T. et al. Endemic chronic wasting disease causes mule deer population decline in Wyoming. PLoS ONE 12, e0186512 (2017).
    PubMed  PubMed Central  Google Scholar 

    13.
    Mysterud, A. & Rolandsen, C. M. A reindeer cull to prevent chronic wasting disease in Europe. Nat. Ecol. Evol. 2, 1343–1345 (2018).
    PubMed  Google Scholar 

    14.
    V. K. M. Ytrehus, et al. Factors that can Contribute to Spread of CWD—An Update on the Situation in Nordfjella, Norway (Opinion of the Panel on biological hazards. Norwegian Scientific Committee for Food and Environment (VKM), Oslo, Norway, 2018).

    15.
    Vors, L. S. & Boyce, M. S. Global declines of caribou and reindeer. Glob. Change Biol. 15, 2626–2633 (2009).
    ADS  Google Scholar 

    16.
    Diefenbach, D. R., Rosenberry, C. S. & Boyd, R. C. From the field: efficacy of detecting chronic wasting disease via sampling hunter-killed white-tailed deer. Wildl. Soc. Bull. 32, 267–272 (2004).
    Google Scholar 

    17.
    Rees, E. E. et al. Targeting the detection of chronic wasting disease using the hunter harvest during early phases of an outbreak in Saskatchewan, Canada. Prev. Vet. Med. 104, 149–159 (2012).
    PubMed  Google Scholar 

    18.
    Belsare, A. V. et al. An agent-based framework for improving wildlife disease surveillance: a case study of chronic wasting disease in Missouri white-tailed deer. Ecol. Model. 417, 108919 (2020).
    Google Scholar 

    19.
    Walsh, D. P. & Miller, M. W. A weighted surveillance approach for detecting chronic wasting disease foci. J. Wildl. Dis. 46, 118–135 (2010).
    PubMed  Google Scholar 

    20.
    Heisey, D. M. et al. Linking process to pattern: estimating spatiotemporal dynamics of a wildlife epidemic from cross-sectional data. Ecol. Monogr. 80, 221–240 (2010).
    Google Scholar 

    21.
    Miller, M. W. & Conner, M. M. Epidemiology of chronic wasting disease in free-ranging mule deer: Spatial, temporal, and demographic influences on observed prevalence patterns. J. Wildl. Dis. 41, 275–290 (2005).
    PubMed  Google Scholar 

    22.
    Samuel, M. D. & Storm, D. J. Chronic wasting disease in white-tailed deer: infection, mortality, and implications for heterogeneous transmission. Ecol. 97, 3195–3205 (2016).
    Google Scholar 

    23.
    Mysterud, A., Coulson, T. & Stenseth, N. C. The role of males in the population dynamics of ungulates. J. Anim. Ecol. 71, 907–915 (2002).
    Google Scholar 

    24.
    Ginsberg, J. R. & Milner-Gulland, E. J. Sex biased harvesting and population dynamics in ungulates: implications for conservation and sustainable use. Cons. Biol. 8, 157–166 (1994).
    Google Scholar 

    25.
    Milner-Gulland, E. J., Coulson, T. N. & Clutton-Brock, T. H. On harvesting a structured ungulate population. Oikos 88, 592–602 (2000).
    Google Scholar 

    26.
    Stärk, K. D. C. et al. Concepts for risk-based surveillance in the field of veterinary medicine and veterinary public health: review of current approaches. BMC Health Serv. Res. 6, 20 (2006).
    PubMed  PubMed Central  Google Scholar 

    27.
    Martin, P. A., Cameron, A. R. & Greiner, M. Demonstrating freedom from disease using multiple complex data sources. Prev. Vet. Med. 79, 71–97 (2007).
    CAS  PubMed  Google Scholar 

    28.
    Cannon, R. M. Demonstrating disease freedom-combining confidence levels. Prev. Vet. Med. 52, 227–249 (2002).
    CAS  PubMed  Google Scholar 

    29.
    Sutherland, W. J. et al. A 2018 horizon scan of emerging Issues for global conservation and biological diversity. Trends Ecol. Evol. 33, 47–58 (2018).
    PubMed  Google Scholar 

    30.
    EFSA Panel on Biological Hazards (BIOHAZ), Ricci, A. et al. Chronic wasting disease (CWD) in cervids. EFSA J. 15, 4667 (2016).
    Google Scholar 

    31.
    Vicente, J. et al. Science-based wildlife disease response. Science 364, 943 (2019).
    ADS  PubMed  Google Scholar 

    32.
    Schalk, G. & Forbes, M. R. Male biases in parasitism of mammals: effects of study type, host age, and parasite taxon. Oikos 78, 67–74 (1997).
    Google Scholar 

    33.
    Córdoba-Aguilar, A. & Munguía-Steyer, R. The sicker sex: understanding male biases in parasitic infection, resource allocation and fitness. Plos One 8, e76246 (2013).
    ADS  PubMed  PubMed Central  Google Scholar 

    34.
    Milner-Gulland, E. J. et al. Reproductive collapse in saiga antelope harems. Nature 422, 135 (2003).
    ADS  CAS  PubMed  Google Scholar 

    35.
    Sargeant, G. A., Weber, D. C. & Roddy, D. E. Implications of chronic wasting disease, cougar predation, and reduced recruitment for elk management. J. Wildl. Manag. 75, 171–177 (2011).
    Google Scholar 

    36.
    Monello, R. J. et al. Survival and population growth of a free-ranging elk population with a long history of exposure to Chronic wasting disease. J. Wildl. Manag. 78, 214–223 (2014).
    Google Scholar 

    37.
    Argue, C. K., Ribble, C., Lees, V. W., McLane, J. & Balachandran, A. Epidemiology of an outbreak of chronic wasting disease on elk farms in Saskatchewan. Can. Vet. J. 48, 1241–1248 (2007).
    PubMed  PubMed Central  Google Scholar 

    38.
    Delahay, R. J. Smith, G. C. & Hutchings, M. R. Management of Disease in Wild Mammals (Springer, Tokyo, Japan, 2009).

    39.
    Almberg, E. S., Cross, P. C., Johnson, C. J., Heisey, D. M. & Richards, B. J. Modeling routes of chronic wasting disease transmission: environmental prion persistence promotes deer population decline and extinction. PLoS ONE 6, e19896 (2011).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    40.
    Keeling, M. J. The effects of local spatial structure on epidemiological invasions. Proc. R. Soc. Lond. Ser. B 266, 859–867 (1999).
    CAS  Google Scholar 

    41.
    Joly, D. O., Samuel, M. D., Langenberg, J. A., Rolley, R. E. & Keane, D. P. Surveillance to detect chronic wasting disease in white-tailed deer in Wisconsin. J. Wildl. Dis. 45, 989–997 (2009).
    PubMed  Google Scholar 

    42.
    Nusser, S. M., Clark, W. R., Otis, D. L. & Huang, L. Sampling considerations for disease surveillance in wildlife populations. J. Wildl. Manag. 72, 52–60 (2008).
    Google Scholar 

    43.
    Osnas, E. E., Heisey, D. M., Rolley, R. E. & Samuel, M. D. Spatial and temporal patterns of chronic wasting disease: fine-scale mapping of a wildlife epidemic in Wisconsin. Ecol. Appl. 19, 1311–1322 (2009).
    PubMed  Google Scholar 

    44.
    Samuel, M. D. et al. Surveillance strategies for detecting chronic wasting disease in free-ranging deer and elk – results of a CWD surveillance workshop. https://pubs.er.usgs.gov/publication/70006758 (U.S. Geological Survey Conference publication, Madison, WI, 2003).

    45.
    Spraker, T. R. et al. Spongiform encephalopathy in free-ranging mule deer (Odocoileus hemionus), white-tailed deer (Odocoileus virginianus) and Rocky Mountain elk (Cervus elaphus nelsoni) in northcentral Colorado. J. Wildl. Dis. 33, 1–6 (1997).
    CAS  PubMed  Google Scholar 

    46.
    Panzacchi, M. et al. Predicting the continuum between corridors and barriers to animal movements using Step Selection Functions and Randomized Shortest Paths. J. Anim. Ecol. 85, 32–42 (2015).
    PubMed  Google Scholar 

    47.
    Ziller, M., Selhorst, T., Teuffert, J., Kramer, M. & Schlüter, H. Analysis of sampling strategies to substantiate freedom from disease in large areas. Prev. Vet. Med. 52, 333–343 (2002).
    CAS  PubMed  Google Scholar 

    48.
    Jongman, R. H. G. Homogenisation and fragmentation of the European landscape: ecological consequences and solutions. Landsc. Urban Plan. 58, 211–221 (2002).
    Google Scholar 

    49.
    Holand, Ø. et al. The effect of sex ratio and male age structure on reindeer calving. J. Wildl. Manag. 67, 25–33 (2003).
    Google Scholar 

    50.
    Sæther, B.-E., Solberg, E. J. & Heim, M. Effects of altering sex ratio structure on the demography of an isolated moose population. J. Wildl. Manag. 67, 455–466 (2003).
    Google Scholar 

    51.
    Morina, D. L., Demarais, S., Strickland, B. K. & Larson, J. E. While males fight, females choose: male phenotypic quality informs female mate choice in mammals. Anim. Behav. 138, 69–74 (2018).
    Google Scholar 

    52.
    Bro-Jørgensen, J. Overt female competition and preference for central males in a lekking antelope. Proc. Natl Acad. Sci. USA 99, 9290–9293 (2002).
    ADS  PubMed  Google Scholar 

    53.
    Andres, D. et al. Sex differences in the consequences of maternal loss in a long-lived mammal, the red deer (Cervus elaphus). Behav. Ecol. Sociobiol. 67, 1249–1258 (2013).
    Google Scholar 

    54.
    Ericsson, G. Reduced cost of reproduction in moose Alces alces through human harvest. Alces 37, 61–69 (2001).
    Google Scholar 

    55.
    Apollonio, M. Andersen, R. & Putman, R. European Ungulates and their Management in the 21st Century (Cambridge University Press, Cambridge, 2010).

    56.
    Mawson, P. R., Hampton, J. O. & Dooley, B. Subsidized commercial harvesting for cost-effective wildlife management in urban areas: a case study with kangaroo sharpshooting. Wildl. Soc. Bull. 40, 251–260 (2016).
    Google Scholar 

    57.
    Manjerovic, M. B., Green, M. L., Mateus-Pinilla, N. & Novakofski, J. The importance of localized culling in stabilizing chronic wasting disease prevalence in white-tailed deer populations. Prev. Vet. Med. 113, 139–145 (2014).
    PubMed  Google Scholar 

    58.
    Mateus-Pinilla, N., Weng, H. Y., Ruiz, M. O., Shelton, P. & Novakofski, J. Evaluation of a wild white-tailed deer population management program for controlling chronic wasting disease in Illinois, 2003-2008. Prev. Vet. Med. 110, 541–548 (2013).
    PubMed  Google Scholar 

    59.
    Vaske, J. J. Lessons learned from human dimensions of chronic wasting disease research. Hum. Dimens Wildl. 15, 165–179 (2010).
    Google Scholar 

    60.
    Mysterud, A., Strand, O. & Rolandsen, C. M. Efficacy of recreational hunters and marksmen for host culling to combat chronic wasting disease in reindeer. Wildl. Soc. Bull. 43, 683–692 (2019).
    Google Scholar 

    61.
    Gaydos, D. A., Petrasova, A., Cobb, R. C. & Meentemeyer, R. K. Forecasting and control of emerging infectious forest disease through participatory modelling. Philos. Trans. R. Soc. Lond. B Biol. Sci. 374, 20180283 (2019).
    PubMed  PubMed Central  Google Scholar 

    62.
    Strand, O., Nilsen, E. B., Solberg, E. J. & Linnell, J. D. C. Can management regulate the population size of wild reindeer (Rangifer tarandus) through harvest? Can. J. Zool. 90, 163–171 (2012).
    Google Scholar 

    63.
    Nilsen, E. B. & Strand, O. Integrating data from several sources for increased insight into demographic processes: Simulation studies and proof of concept for hierarchical change in ratio models. PLoS ONE 13, e0194566 (2018).
    PubMed  PubMed Central  Google Scholar 

    64.
    Viljugrein, H. et al. A method that accounts for differential detectability in mixed samples of long-term infections with applications to the case of chronic wasting disease in cervids. Methods Ecol. Evol. 10, 134–145 (2019).
    Google Scholar 

    65.
    Mysterud, A. et al. The demographic pattern of infection with chronic wasting disease in reindeer at an early epidemic stage. Ecosphere 10, e02931 (2019).
    Google Scholar 

    66.
    MacDiarmid, S. C. A theoretical basis for the use of a skin test for brucellosis surveillance in extensively-managed cattle herds. Rev. Sci. Tech. Int Epiz 6, 1029–1035 (1987).
    CAS  Google Scholar 

    67.
    Viljugrein, H. Accompanying Code for the Paper “Hunting Wildlife to Increase Disease Detection” Version v1.0.0, August 4-2020) https://doi.org/10.5281/zenodo.3972037 (Zenodo, 2020). More

  • in

    Photosynthetic base of reduced grain yield by shading stress during the early reproductive stage of two wheat cultivars

    Wheat cultivars and growing conditions
    In this study, pot and field experiments with the shade-tolerant cultivar Henong825 and the shade-sensitive cultivar Kenong9204 were performed. These two winter wheat cultivars were identified with different degrees of shade tolerance by our previous study17. Both cultivars are released by Hebei Province, China, which are the most widely planted wheat cultivars in North China Plain. The parental combination of Henong825 and Kenong9204 is Linyuan95-3091/Shi4185, SA502/6021, respectively. Henong825 is characterized by strong lodging resistance. Kenong9204 is characterized by suitable for moderate water and fertilizer. Two field experiments were conducted during the 2016–2017 and 2017–2018 wheat-growing seasons in the Luancheng agro-ecosystem experimental station of the Chinese Academy of Sciences, Hebei Province (37° 53′ N and 114° 41′ E; elevation at 50 m). The climate characterizing of the study region is summer monsoon. The mean temperature, total precipitation, and solar radiation in both the winter wheat-growing seasons are shown in Table 5. The soil used in the experiments was loam containing 21.41 g kg−1 organic matter, 109.55 mg kg−1 alkaline nitrogen (N), 1.44 g kg−1 total N, 15.58 mg kg−1 available phosphorus (P), and 220 mg kg−1 rapidly available potassium (K). In both seasons, soils were fertilized with urea (N, 46%) and complete fertilizer (N–P, 21–54%) at 300 kg ha−1 and 375 kg ha−1. Seeds were sown by hand on October 6, 2016 and October 17, 2017, then the seedlings emerged 1 week later. In 2017 growing season, the YM stage was on April 15, and anthesis stage was on May 1 in both cultivars. In 2018 growing season, the YM stage was on April 16, and anthesis stage was on May 2 in both cultivars. The seedling density was 166 m−2, which is the norm in this region.
    Table 5 The monthly mean temperature (°C), total precipitation (mm), and solar radiation (MJ m−2 day−1) during the two growing seasons of winter wheat in 2016–2017 and 2017–2018.
    Full size table

    Experimental design
    This study was a combination of field experiment and pot experiment to investigate the effect of different shading intensity and duration during YM stage on grain components and photosynthetic characteristics. Pot experiment was supplement to field experiment.
    Field experiments
    The experiments were arranged in a randomized split-split plot design with three replicates. The main plots were split into three subplots subjected to one of three shading intensities: 100% (CK, control), 40% (S1), and 10% (S2) of natural light. Each subplot was split into four sub-subplots, which were randomly allocated to one of four durations: 1 day (D1), 3 days (D3), 5 days (D5), and 7 days (D7) during the YM stage. The shading treatments were conducted in these periods and replicated three times. Each plot size was 6 m long and 2 m wide, with 40 rows. There were 72 plots. Different degrees of artificial shade were provided by using black polyethylene screens horizontally installed at a height of 2 m above the ground.
    Determination of YM stage
    The YM stage roughly corresponds to Zadok’s scale from Z37 (main stem with flag leaf is visible) to Z39 (flag leaf ligule is noticeable). According to previous researches of YM stage, the estimated measurement of the YM stage was based on the auricle distance (AD, the distance between the auricle of the flag leaf and the auricle of the penultimate leaf) of main stem43,44. In order to keep the relationship between the occurrence of YM and AD unchanged, the field management practices, adequate irrigation was the same in two growing-seasons. Moreover, for each experiment, at the onset of appearance of the flag leaf of the main stem, 30 anthers of ten main stem spike of wheat were randomly sampled to establish the timing of YM stage initiation1. The correlation of the AD with the development of the YM stage in the florets of the two cultivars was measured and observed using microscope (Fig. 9). The cultivar Henong825 reached the YM stage at 1–2 cm, whereas Kenong9204 reached the YM stage at − 1 to 0 cm. To capture the YM stage in the shading condition, the plants were subjected to shading stress ahead of the YM stage occurrence. When more than 50% of the plants in each plot reached − 2 cm in Henong825 and − 4 cm in Kenong9204, the main stem of the plants was tagged, and shading stress was applied in each plot. Each experimental plot for Henong825 and Kenong9204 was independently subjected to shading stress on April 15, 2017 and April 16, 2018. When the shading stress treatments ended, the shade screens were removed and were exposed to natural light until they matured. Air temperature, light intensity, and relative humidity above the canopy were recorded using a portable weather station (ECA-YW0501; Beijing, China) during the shading period. Light spectral was measured using a portable geographic spectrometer (PSR + 3500, USA). The irradiance of spectral wavelength ranging from 350 to 2,500 nm was measured. The proportions of blue light (B/T), green light (G/T), red light (R/T), far-red light (FR/T), and red/far red (R/FR) were calculated according to their irradiance at 400–500 nm, 500–600 nm, 600–700 nm, and 700–800 nm, respectively. Following the local field management practices, adequate irrigation was conducted three times during the overwinter, jointing, and anthesis stages of the wheat-growing season. Weeds, fungal diseases, and insect pests were controlled through spraying of conventional herbicides, fungicides, and insecticides, correspondingly.
    Figure 9

    The relationship between anther development and shading period in two wheat cultivars.

    Full size image

    Pot experiments
    The pot experiments were conducted in a temperature-controlled glasshouse. Vernalized seedlings of the two wheat cultivars were transplanted to pots (45 cm in length, 28.5 cm in width and 20 cm in height; 18 plants in each pot; three pots for each treatment group) containing a mixture of vermiculite and nutritional soil (1:1). All wheat seedlings were grown at a day temperature of 25 °C, night temperature of 15 °C, and light intensity of 800 μmol m−2 s−1. When the AD of the main stems of Henong825 and Kenong9204 cultivars were approximately − 2 cm and − 4 cm, respectively, the main stem of the plants was tagged, and shading stress was applied in each treatment. Shading treatments groups were the different shading intensities and shading durations previously mentioned. The shading condition in glasshouse was simulated with black polyethylene screen to keep up with the experimental methods in the field. After shading stress, the shading nets were removed, until the crops matured.
    Sampling and measurements
    Photosynthetic rate, stomatal conductance, intercellular carbon dioxide, and chlorophyll fluorescence parameters
    In field experiments, three randomly selected flag leaves on the tagged main stems of plants in each plot were analyzed to determine Pn, Gs, Ci, and chlorophyll fluorescence. For each shading treatment group, Pn, Gs, and Ci were measured using an LI-6400XT portable system (LI-COR Biosciences, Nebraska, USA), and the chamber of which was equipped with a red/blue LED light source (LI6400-02B) before the shading stress was removed. Before measurement, the machine was preheated for 30 min, and checked, adjusted to zero, calibrated according to the instructions. Moreover, the light intensity in measured chamber was equivalent of shading treatment conditions. The flow rates was set at 500 μmol s−1, The temperature in chamber was set 25 °C. The CO2 concentration was set to 400 μmol mol−1, which was provided by carbon dioxide cylinders to maintain a stable CO2 environment. The chlorophyll fluorescence of flag leaves on the tagged main stems of plants were measured using a modulate chlorophyll fluorescence imaging system (Imaging-PAM; Hansatech, UK) in each plot. The primary light energy conversion efficiency of PSII (Fv/Fm) and actual photochemical quantum efficiency (YII) were measured after 30 min of dark adaptation. The saturation irradiance (PARsat) and maximum electron transport (Jmax) of flag leaves in each treatment were calculated using a modified rectangular hyperbola. On the day next to shading removal, the Pn of three flag leaves from each replicate plot were measured.
    Chlorophyll content
    For glasshouse pot experiments, nine flag leaves (three leaves were randomly selected per pot from three pots in each treatment group) tagged main stems of plant were selected prior to the removal of shading. The flag leaves were then sliced following the removal of the main vein. After the sliced fresh leaves were weighed to 0.1 g, the chlorophyll content of leaves was extracted with 80% acetone for 48 h and analyzed through micro-determination (Thermo Varioskan Flash, USA). The absorbance of chlorophyll a (chl a) and chlorophyll b (chl b) was read at 663 and 646 nm, respectively (Thermo Varioskan Flash, USA), and the chlorophyll contents were calculated according to following equations: chl a (mg/g) = (12.7 × A663 nm–2.69 × A646 nm)/(100 × M); and chl b (mg/g) = (22.9 × A646 nm–4.68 × A663 nm)/(100 × M) where A663 and A646 are absorption levels at 663 and 646 nm, respectively; M is leaf fresh weight. The total chlorophyll (chl a + chl b) values were calculated by chl a and chl b values.
    Leaf anatomy and surface characteristics
    The approximately 2-mm2 leaf sections in D7 treatments and one day after recovery were harvested from the center of three flag leaves on the tagged main stems of plants using a scalpel and were rapidly fixed in electron microscope fixation fluid at 4 °C overnight. Stomatal apertures and chloroplast ultrastructure were observed by Servicebio (Wuhan) using a scanning electron microscope (SU8100; Hitachi) and a transmission electron microscope (HT7700; Hitachi). Simultaneously, the fully expanded flag leaves collected from plants in each treatment were fixed with FAA solution and embedded in paraffin to measure the leaf anatomical structure. The embedded wax block were sectioned to a thickness of 8 μm, then following dewaxing in environmental transparent solution and rehydration in a series of graded alcohol solutions. Finally, the tissue samples were stained with safranin and fast green, observed under a Leica DM6 microscope (Leica, Germany), and the respective images were obtained.
    Grain yield, yield components, and aboveground biomass
    At harvest in the field experiments during both growing seasons, 60 tagged plants per replicate were randomly sampled to determine grain yield components. The harvested plants were naturally dried to a grain water content of approximately 11%. Each tagged plant was then threshed using a single plant threshing machine to determine the grain number and grain yield needed for the estimation of the average grain weight. In addition, 30 tagged winter wheat plants were uprooted randomly and gradually by hand from each plot. Each plant was cut from the root and was dried at 80 °C. Aboveground biomass was measured using a precision digital balance (model BSA3202S; Sartorius, Germany) with a precision of 0.01 g.
    Statistical analysis
    The experimental data for grain yield, yield components, biomass and chlorophyll fluorescence parameters were analyzed using a general linear model procedure (GLM) in SPSS 22.0 for a split-split plot design. The significant differences among treatment mean values were determined by the least significance difference analysis (LSD, P  More