More stories

  • in

    Co-haplotyping symbiont and host to unravel invasion pathways of the exotic pest Halyomorpha halys in Italy

    1.
    Lashkari, M. et al. Global genetic variation in the Asian citrus psyllid Diaphorina citri (Hemiptera: Liviidae) and the endosymbiont Wolbachia: links between Iran and the USA detected. Pest Manag. Sci. 70, 1033–1040 (2014).
    CAS  PubMed  Article  Google Scholar 
    2.
    Wang, Y. et al. Phylogeography of Diaphorina citri (Hemiptera: Liviidae) and its primary endosymbiont, ‘Candidatus Carsonella ruddii’: An evolutionary approach to host–endosymbiont interaction. Pest Manag. Sci. 74, 2185–2194 (2018).
    CAS  Article  Google Scholar 

    3.
    Otero-Bravo, A. & Sabree, Z. L. Comparing the utility of host and primary endosymbiont loci for predicting global invasive insect genetic structuring and migration patterns. Biol. Control. 116, 1016 (2018).
    Article  Google Scholar 

    4.
    Savio, C. et al. Evidence of two lineages of the symbiont ‘Candidatus Erwinia dacicola’ in Italian populations of Bactrocera oleae (Rossi) based on 16S rRNA gene sequences. Int. J. Syst. Evol. Micr. 62, 179–187 (2012).
    CAS  Article  Google Scholar 

    5.
    Funk, D. J., Helbling, L., Wernegreen, J. J. & Moran, N. A. Intraspecific phylogenetic congruence among multiple symbiont genomes. P. Roy Soc. B-Biol. Sci. 267, 2517–2521 (2000).
    CAS  Article  Google Scholar 

    6.
    Criscione, C. D., Cooper, B. & Blouin, M. S. Parasite genotypes identify source populations of migratory fish more accurately than fish genotypes. Ecology 87, 823–828 (2006).
    PubMed  Article  Google Scholar 

    7.
    Brown, J. E. et al. Aedes aegypti mosquitoes imported into the Netherlands, 2010. Emerg. Infect. Dis. 17, 2335 (2011).
    PubMed  PubMed Central  Article  Google Scholar 

    8.
    Simberloff, D. et al. Impacts of biological invasions: What’s what and the way forward. Trends Ecol. Evol. 28, 58–66 (2013).
    PubMed  Article  Google Scholar 

    9.
    Maynard, A. J. et al. Tiger on the prowl: Invasion history and spatio-temporal genetic structure of the Asian tiger mosquito Aedes albopictus (Skuse 1894) in the Indo-Pacific. PLoS Neglect. Trop. D. 11, e0005546 (2017).
    Article  Google Scholar 

    10.
    Bansal, R., Michel, A. & Sabree, Z. The crypt-dwelling primary bacterial symbiont of the polyphagous pentatomid pest Halyomorpha halys (Hemiptera: Pentatomidae). Environ. Entomol. 43, 617–625 (2014).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    11.
    Kenyon, L. J., Meulia, T. & Sabree, Z. L. Habitat visualization and genomic analysis of “Candidatus Pantoea carbekii”, the primary symbiont of the brown marmorated stink bug. Genome Biol. Evol. 7, 620–635 (2015).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    12.
    Leskey, T. C. & Nielsen, A. L. Impact of the invasive brown marmorated stink bug in North America and Europe: History, biology, ecology, and management. Annu. Rev. Entomol. 63, 599–618 (2018).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    13.
    Moore, L. et al. Characterizing damage potential of the brown marmorated stink bug in cherry orchards in Italy. Entomol. Gen. 3–4, 271–283 (2019).
    Article  Google Scholar 

    14.
    Rice, K. B. et al. Biology, ecology, and management of brown marmorated stink bug (Hemiptera: Pentatomidae). J. Integr. Pest Manag. 5, 1–13 (2014).
    Article  Google Scholar 

    15.
    Wermelinger, B., Wyniger, D. & Forster, B. First records of an invasive bug in Europe: Halyomorpha halys Stål (Heteroptera: Pentatomidae), a new pest on woody ornamentals and fruit trees?. Mitt. Schweiz. Entomol. Ges. 81, 1–8 (2008).
    Google Scholar 

    16.
    Arnold, K. Halyomorpha halys (Stål, 1855), Eine fur die europaische Fauna neu nachgewiesene Wanzenart (Insecta: Heteroptera: Pentatomidae: Cappaeini). Mitt. Thür. Entomologogenverbande. 16, 19 (2009).
    Google Scholar 

    17.
    Heckmann, R. Erster nachweis von Halyomorpha halys (Stål, 1855) (Heteroptera: Pentatomidae) für deutschland. Heteropteron 36, 17–18 (2012).
    Google Scholar 

    18.
    Callot, H. & Brua, C. Halyomorpha halys (Stål, 1855), la Punaise diabolique, nouvelle espece pour la faune de France (Heteroptera Pentatomidae). Entomologiste. 69, 69–71 (2013).
    Google Scholar 

    19.
    Maistrello, L. et al. First records in Italy of the Asian stinkbug Halyomorpha halys a new threat to fruit crops. Atti Giornate Fitopatologiche. 1, 283–288 (2014).
    Google Scholar 

    20.
    Gariepy, T. D., Haye, T., Fraser, H. & Zhang, J. Occurrence, genetic diversity, and potential pathways of entry of Halyomorpha halys in newly invaded areas of Canada and Switzerland. J. Pest Sci. 87, 17–28 (2014).
    Article  Google Scholar 

    21.
    Milonas, P. G. & Partsinevelos, G. K. First report of brown marmorated stink bug Halyomorpha halys Stål (Hemiptera: Pentatomidae) in Greece. EPPO Bull. 44, 183–186 (2014).
    Article  Google Scholar 

    22.
    Vétek, G., Papp, V., Haltrich, A. & Rédei, D. First record of the brown marmorated stink bug, Halyomorpha halys (Hemiptera: Heteroptera: Pentatomidae), in Hungary, with description of the genitalia of both sexes. Zootaxa. 3780, 194–200 (2014).
    PubMed  Article  PubMed Central  Google Scholar 

    23.
    Cesari, M. et al. A pest alien invasion in progress: potential pathways of origin of the brown marmorated stink bug Halyomorpha halys populations in Italy. J. Pest Sci. 88, 1–7 (2015).
    Article  Google Scholar 

    24.
    Macavei, L. et al. First detection of Halyomorpha halys, a new invasive species with a high potential of damage on agricultural crops in Romania. Lucr. Stiint. Ser. Agron. 58, 105–108 (2015).
    Google Scholar 

    25.
    Rabitsch, W. & Griebe, G. J. From the west and from the east? First records of Halyomorpha halys (Stål, 1855) (Hemiptera: Heteroptera: Pentatomidae) in Vorarlberg and Vienna, Austria. Beitr. Entomof. 16, 115–139 (2015).
    Google Scholar 

    26.
    Šeat, J. Halyomorpha halys (Stål, 1855) (Heteroptera: Pentatomidae) a new invasive species in Serbia. Acta Entomol. Serbica. 20, 167–171 (2015).
    Google Scholar 

    27.
    Maurel, J. P., Blaye, G., Valladares, L., Roinel, É. & Cochard, P. O. Halyomorpha halys (Stål, 1855), la punaise diabolique en France, a Toulouse (Heteroptera; Pentatomidae). Carnets Nat. 3, 2125 (2016).
    Google Scholar 

    28.
    Dioli, P., Leo, P. & Maistrello, L. Prime segnalazioni in Spagna e in Sardegna della specie aliena Halyomorpha halys (Stål, 1855) e note sulla sua distribuzione in Europa (Hemiptera, Pentatomidae). Rev. gaditana Entomol. 7, 539–548 (2016).
    Google Scholar 

    29.
    Simov, N. The invasive brown marmorated stink bug Halyomorpha halys (Stål, 1855) (Heteroptera: Pentatomidae) already in Bulgaria. Ecol. Montenegrina. 9, 51–53 (2016).
    Article  Google Scholar 

    30.
    Hemala, V. & Kment, P. First record of Halyomorpha halys and mass occurrence of Nezara viridula in Slovakia. Plant Prot. Sci. 53, 1–7 (2017).
    Article  Google Scholar 

    31.
    Claerebout, S., Haye, T., Olafsson, E., Pannier, E. & Bultot, J. Premières occurrences de Halyomorpha halys (Stål, 1855) pour la Belgique et actualisation de sa répartition en Europe (Hemiptera: Heteroptera: Pentatomidae). Bull. Soc. R. belge entomolog. 154, 205–227 (2019).
    Google Scholar 

    32.
    Tassini, C. & Mifsud, D. The brown marmorated stink bug, Halyomorpha halys (Hemiptera: Heteroptera: Pentatomidae) in Malta. EPPO Bull. 49, 132–136 (2019).
    Article  Google Scholar 

    33.
    Zovko, M., Ostojić, I., Jurković, D., & Karić, N. (2019). First report of the brown marmorated stink bug, Halyomorpha halys (Stål, 1855) in Bosnia and Herzegovina. Radovi Poljoprivredno Prehrambenog Fakulteta Univerziteta u SarajevuWorks of the Faculty of Agricultural and Food Sciences University of Sarajevo. 64, 68–78 (2019).

    34.
    Gapon, D. A. First records of the brown marmorated stink bug, Halyomorpha halys (Stål, 1855) (Heteroptera: Pentatomidae) in Russia, Abkhazia, and Georgia. Entomol. Rev. 96, 1086–1088 (2016).
    Article  Google Scholar 

    35.
    Musolin, D. L. et al. Invasive brown marmorated stink bug Halyomorpha halys (Stål) (Heteroptera: Pentatomidae) in Russia, Abkhazia, and Serbia: history of invasion, range expansion, early stages of establishment, and first records of damage to local crops. Arthropod.-Plant Interact. 12, 517–529 (2018).
    Article  Google Scholar 

    36.
    Kriticos, D. J. The potential global distribution of the brown marmorated stink bug, Halyomorpha halys, a critical threat to plant biosecurity. J. Pest Sci. 90, 1033–1043 (2017).
    Article  Google Scholar 

    37.
    Kapantaidaki, D. E. et al. Halyomorpha halys (Hemiptera: Pentatomidae) genetic diversity in North America and Europe. Insects 10, 174. https://doi.org/10.3390/insects10060174 (2019).
    Article  PubMed Central  PubMed  Google Scholar 

    38.
    Xu, J., Fonseca, D. M., Hamilton, G. C., Hoelmer, K. A. & Nielsen, A. L. Tracing the origin of US brown marmorated stink bugs, Halyomorpha halys. Biol. Invas. 16, 153–166 (2014).
    CAS  Article  Google Scholar 

    39.
    Cesari, M. et al. Genetic diversity of the brown marmorated stink bug Halyomorpha halys in the invaded territories of Europe and its patterns of diffusion in Italy. Biol. Invas. 20, 1073–1092 (2018).
    Article  Google Scholar 

    40.
    Zhu, G. P. et al. Range wide molecular data and niche modeling revealed the Pleistocene history of a global invader (Halyomorpha halys). Sci. Rep. 6, 23192 (2016).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    41.
    Valentin, R. E., Nielsen, A. L., Wiman, N. G., Lee, D. H. & Fonseca, D. M. Global invasion network of the brown marmorated stink bug, Halyomorpha halys. Sci. Rep. 7, 9866 (2017).
    ADS  PubMed  PubMed Central  Article  Google Scholar 

    42.
    Lee, W. et al. Genetic diversity of Halyomorpha halys (Hemiptera, Pentatomidae) in South Korea and comparison with COI sequence datasets from East Asia, Europe, and North America. Fla. Entomol. 101, 49–55 (2018).
    Article  Google Scholar 

    43.
    Morrison, W. R. et al. Attraction of Halyomorpha halys (Hemiptera: Pentatomidae) haplotypes in North America and Europe to baited traps. Sci. Rep. 7, 16941. https://doi.org/10.1038/s41598-017-17233-0 (2017).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    44.
    Dlugosch, K. M. & Parker, I. M. Founding events in species invasions: Genetic variation, adaptive evolution, and the role of multiple introductions. Mol. Ecol. 17, 431–449 (2008).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    45.
    Freers, A. Blinde Passagiere: Stinkwanzen, Marmorierte Baumwanze—Halyomorpha halys. https://ssl.bremen.de/lmtvet/sixcms/media.php/13/Blinde_Passagiere_Stinkwanze_3_2012.pdf. (2012).

    46.
    Gariepy, T. D., Bruin, A., Haye, T., Milonas, P. & Vétek, G. Occurrence and genetic diversity of new populations of Halyomorpha halys in Europe. J. Pest Sci. 88, 451–460 (2015).
    Article  Google Scholar 

    47.
    Croteau, E. K. Causes and consequences of dispersal in plants and animals. Nat. Educ. Know. 3, 12 (2010).
    Google Scholar 

    48.
    Schellhorn, N. A., Bianchi, F. J. J. A. & Hsu, C. L. Movement of entomophagous arthropods in agricultural landscapes: Links to pest suppression. Annu. Rev. Entomol. 59, 559–581 (2014).
    CAS  PubMed  Article  Google Scholar 

    49.
    Rassati, D., Faccoli, M., Petrucco Toffolo, E., Battisti, A. & Marini, L. Improving the early detection of alien wood-boring beetles in ports and surrounding forests. J. Appl. Ecol. 52, 50–58 (2015).
    CAS  Article  Google Scholar 

    50.
    Duthie, C., Michael, T., Stephenson, B., Yamoah, E. & McDonald, B. Risk analysis of Halyomorpha halys (brown marmorated stink bug) on all pathways (Ministry for Primary Industries, Wellington, 2012).
    Google Scholar 

    51.
    Haye, T. et al. Range expansion of the invasive brown marmorated stinkbug, Halyomorpha halys: an increasing threat to field, fruit and vegetable crops worldwide. J. Pest Sci. 88, 665–673 (2015).
    Article  Google Scholar 

    52.
    Vandervoet, T. F., Bellamy, D. E., Anderson, D. & MacLellan, R. Trapping for early detection of the brown marmorated stink bug, Halyomorpha halys, New Zealand. N. Z. Plant Prot. 72, 36–43 (2019).
    Google Scholar 

    53.
    Hosokawa, T., Matsuura, Y., Kikuchi, Y. & Fukatsu, T. Recurrent evolution of gut symbiotic bacteria in pentatomid stinkbugs. Zool. Let. 2, 24 (2016).
    Article  Google Scholar 

    54.
    Allen, J. M., Light, J. E., Perotti, M. A., Braig, H. R. & Reed, D. L. Mutational meltdown in primary endosymbionts: Selection limits Muller’s ratchet. PLoS ONE 4, 2 (2009).
    Google Scholar 

    55.
    Perotti, M. A., Allen, J. M., Reed, D. L. & Braig, H. R. Host-symbiont interactions of the primary endosymbiont of human head and body lice. FASEB J. 21, 1058–1066 (2007).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    56.
    Ohbayashi, T. et al. Insect’s intestinal organ for symbiont sorting. Proc. Natl. Acad. Sci. USA 112, E5179–E5188 (2015).
    CAS  PubMed  Article  Google Scholar 

    57.
    Liu, Q. et al. A chromosomal-level genome assembly for the insect vector for Chagas disease, Triatoma rubrofasciata. GigaScience 8, 89 (2019).
    ADS  Article  CAS  Google Scholar 

    58.
    Maistrello, L., Dioli, P., Bariselli, M., Mazzoli, G. L. & Giacalone-Forini, I. Citizen science and early detection of invasive species: Phenology of first occurrences of Halyomorpha halys in Southern Europe. Biol. Invas. 18, 3109–3116 (2016).
    Article  Google Scholar 

    59.
    Folmer, O., Black, M., Hoeh, W., Lutz, R. & Vrijenhoek, R. DNA primers for amplification of mitochondrial cytochrome c oxidase subunit I from diverse metazoan invertebrates. Mol. Mar. Biol. Biotechnol. 3, 294–299 (1994).
    CAS  PubMed  PubMed Central  Google Scholar 

    60.
    Kearse, M. et al. Geneious basic: An integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 28, 1647–1649 (2012).
    PubMed  PubMed Central  Article  Google Scholar 

    61.
    Kumar, S., Stecher, G., Li, M., Knyaz, C. & Tamura, K. MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol. Biol. Evol. 35, 1547–1549 (2018).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    62.
    Schloss, P. D., Delalibera, I., Handelsman, J. & Raffa, K. F. Bacteria associated with the guts of two wood-boring beetles: Anoplophora glabripennis and Saperda vestita (Cerambycidae). Environ. Entomol. 35, 625–629 (2006).
    Article  Google Scholar 

    63.
    Excoffier, L. & Lischer, H. E. L. Arlequin suite ver 35: A new series of programs to perform population genetics analyses under Linux and Windows. Mol. Ecol. Resour. 10, 564–567 (2010).
    PubMed  Article  PubMed Central  Google Scholar 

    64.
    Leigh, J. W. & Bryant, D. POPART: Full-feature software for haplotype network construction. Methods Ecol. Evol. 6, 1110–1116 (2015).
    Article  Google Scholar 

    65.
    Clement, M., Posada, D. & Crandall, K. A. TCS: A computer program to estimate gene genealogies. Mol. Ecol. 9, 1657–1659 (2000).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    66.
    Blüthgen, N., Menzel, F. & Blüthgen, N. Measuring specialization in species interaction networks. BMC Ecol. 6, 9 (2006).
    PubMed  PubMed Central  Article  Google Scholar 

    67.
    Dáttilo, W. Different tolerances of symbiotic and nonsymbiotic ant-plant networks to species extinctions. Netw. Biol. 2, 127 (2012).
    Google Scholar 

    68.
    Almeida-Neto, M., Guimaraes, P., Guimaraes, P. R. Jr., Loyola, R. D. & Ulrich, W. A consistent metric for nestedness analysis in ecological systems: Reconciling concept and measurement. Oikos 117, 1227–1239 (2008).
    Article  Google Scholar 

    69.
    Dormann, C. F., Gruber, B. & Fründ, J. Introducing the bipartite package: Analysing ecological networks. R news 8, 8–11 (2008).
    Google Scholar 

    70.
    Tillman, P. G., Greenstone, M. H. & Hu, J. S. Predation of stink bugs (Hemiptera: Pentatomidae) by a complex of predators in cotton and adjoining soybean habitats in Georgia, USA. Fla. Entomol. 2, 1114–1126 (2015).
    Article  Google Scholar 

    71.
    Šapina, I. & Jelaska, L. Š. First report of invasive brown marmorated stink bug Halyomorpha halys (Stål, 1855) in Croatia. EPPO Bull. 48, 138–143 (2018).
    Article  Google Scholar 

    72.
    Dhami, M. K., Dsouza, M., Waite, D. W., Anderson, D. & Li, D. Real-Time PCR assay for the identification of the Brown Marmorated Stink Bug (Halyomorpha halys). Front. Mol. Biosci. 3, 5 (2016).
    PubMed  PubMed Central  Article  CAS  Google Scholar  More

  • in

    Treeline ecotones shape the distribution of avian species richness and functional diversity in south temperate mountains

    1.
    Pearse, W. D. et al. Building up biogeography: pattern to process. J. Biogeogr. 45, 1223–1230 (2018).
    Article  Google Scholar 
    2.
    Wiens, J. J. The niche, biogeography and species interactions. Philos. Trans. R. Soc. 366, 2336–2350 (2011).
    Article  Google Scholar 

    3.
    Blyth, S., Groombridge, B., Lysenko, I., Miles, L. & Newton, A. Mountain watch: environmental change and sustainable developmental in mountains. UNEP World Conserv. Monit. Centre https://doi.org/10.5962/bhl.title.44936 (2002).
    Article  Google Scholar 

    4.
    Nagy, L. & Grabherr, G. The Biology of Alpine Habitats (Oxford University Press, Oxford, 2009).
    Google Scholar 

    5.
    Boyle, W. A., Sandercock, B. K. & Martin, K. Patterns and drivers of intraspecific variation in avian life history along elevational gradients: a meta-analysis. Biol. Rev. 91, 469–482 (2016).
    Article  Google Scholar 

    6.
    Martin, K. et al. Effects of severe weather on reproduction for sympatric songbirds in an alpine environment: interactions of climate extremes influence nesting success. Auk 134, 696–709 (2017).
    Article  Google Scholar 

    7.
    Freeman, B. G., Lee-Yaw, J. A., Sunday, J. M. & Hargreaves, A. L. Expanding, shifting and shrinking: the impact of global warming on species’ elevational distributions. Glob. Ecol. Biogeogr. https://doi.org/10.1111/geb.12774 (2018).
    Article  Google Scholar 

    8.
    Scridel, D. et al. A review and meta-analysis of the effects of climate change on Holarctic mountain and upland bird populations. Ibis (Lond. 1859) 160, 489–515 (2018).
    Article  Google Scholar 

    9.
    Rabosky, D. L. & Hurlbert, A. H. Species richness at continental scales is dominated by ecological limits. Am. Nat. 185, 572–583 (2015).
    PubMed  Article  PubMed Central  Google Scholar 

    10.
    Ibarra, J. T. & Martin, K. Biotic homogenization: loss of avian functional richness and habitat specialists in disturbed Andean temperate forests. Biol. Conserv. 192, 418–427 (2015).
    Article  Google Scholar 

    11.
    MacArthur, R. H. & MacArthur, J. W. On bird species diversity. Ecology 42, 594–598 (1961).
    Article  Google Scholar 

    12.
    Terborgh, J. Distribution on environmental gradients: theory and a preliminary interpretation of distributional patterns in the avifauna of the cordillera vilcabamba, Peru. Ecology 52, 23–40 (1971).
    Article  Google Scholar 

    13.
    Ghalambor, C. K., Huey, R. B., Martin, P. R., Tewksbury, J. J. & Wang, G. Are mountain passes higher in the tropics? Janzen’s hypothesis revisited. Integr. Comp. Biol. 46, 5–17 (2006).
    PubMed  Article  PubMed Central  Google Scholar 

    14.
    Jankowski, J. E., Ciecka, A. L., Meyer, N. Y. & Rabenold, K. N. Beta diversity along environmental gradients: Implications of habitat specialization in tropical montane landscapes. J. Anim. Ecol. 78, 315–327 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    15.
    Jankowski, J. E. et al. The relationship of tropical bird communities to tree species composition and vegetation structure along an Andean elevational gradient. J. Biogeogr. 40, 950–962 (2013).
    Article  Google Scholar 

    16.
    Janzen, D. H. Why mountain passes are higher in the tropics?. Am. Nat. 101, 233–249 (1967).
    Article  Google Scholar 

    17.
    Freeman, B. G. Lower elevation animal species do not tend to be better competitors than their higher elevation relatives. Glob. Ecol. Biogeogr. https://doi.org/10.1111/geb.13014 (2019).
    Article  Google Scholar 

    18.
    He, X., Luo, K., Brown, C. & Lin, L. A taxonomic, functional, and phylogenetic perspective on the community assembly of passerine birds along an elevational gradient in southwest China. Ecol. Evol. 8, 2712–2720 (2018).
    PubMed  PubMed Central  Article  Google Scholar 

    19.
    Terborgh, J. & Weske, J. S. The role of competition in the distribution of Andean birds. Ecology 56, 562–576 (1975).
    Article  Google Scholar 

    20.
    Vuilleumier, F. Forest birds of patagonia: ecological geography, speciation, endemism, and faunal history. Ornithol. Monogr. https://doi.org/10.2307/40168287 (1985).
    Article  Google Scholar 

    21.
    Estades, C. F., Condor, T. & Aug, N. Bird-habitat relationships in a vegetational gradient in the andes of central Chile. Condor 99, 719–727 (1997).
    Article  Google Scholar 

    22.
    Meynard, C. N. & Quinn, J. F. Bird metacommunities in temperate South American forest: vegetation structure, area, and climate effects. Ecology 89, 981–990 (2008).
    PubMed  Article  PubMed Central  Google Scholar 

    23.
    Devictor, V. et al. Spatial mismatch and congruence between taxonomic, phylogenetic and functional diversity: the need for integrative conservation strategies in a changing world. Ecol. Lett. 13, 1030–1040 (2010).
    PubMed  PubMed Central  Google Scholar 

    24.
    Díaz, S. & Cabido, M. Vive la différence: plant functional diversity matters to ecosystem processes. Trends Ecol. Evol. 16, 646–655 (2001).
    Article  Google Scholar 

    25.
    Cadotte, M. W., Carscadden, K. & Mirotchnick, N. Beyond species: functional diversity and the maintenance of ecological processes and services. J. Appl. Ecol. 48, 1079–1087 (2011).
    Article  Google Scholar 

    26.
    Flynn, D. F. B. et al. Loss of functional diversity under land use intensification across multiple taxa. Ecol. Lett. 12, 22–33 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    27.
    Mayfield, M. M. et al. What does species richness tell us about functional trait diversity? Predictions and evidence for responses of species and functional trait diversity to land-use change. Glob. Ecol. Biogeogr. 19, 423–431 (2010).
    Google Scholar 

    28.
    Boersma, K. S. et al. Linking multidimensional functional diversity to quantitative methods: a graphical hypothesis-evaluation framework. Ecology 97, 583–593 (2016).
    PubMed  Article  PubMed Central  Google Scholar 

    29.
    Borges, S. H. et al. Dissecting bird diversity in the Pantepui area of endemism, northern South America. J. Ornithol. 159, 1073–1086 (2018).
    Article  Google Scholar 

    30.
    Herzog, S. K., Kessler, M. & Bach, K. The elevational gradient in Andean bird species richness at the local scale: a foothill peak and a high-elevation plateau. Ecography (Cop.) 28, 209–222 (2005).
    Article  Google Scholar 

    31.
    Araneda, P., Sielfeld, W., Bonacic, C. & Ibarra, J. T. Bird diversity along elevational gradients in the Dry Tropical Andes of northern Chile: the potential role of Aymara indigenous traditional agriculture. PLoS ONE 13, e0207544 (2018).
    PubMed  PubMed Central  Article  Google Scholar 

    32.
    Caviedes, J. & Ibarra, J. T. Influence of anthropogenic disturbances on stand structural complexity in Andean temperate forests: implications for managing key habitat for biodiversity. PLoS ONE 12, e0169450 (2017).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    33.
    Altamirano, T. A., Ibarra, J. T., Martin, K. & Bonacic, C. The conservation value of tree decay processes as a key driver structuring tree cavity nest webs in South American temperate rainforests. Biodivers. Conserv. 26, 2453–2472 (2017).
    Article  Google Scholar 

    34.
    Blake, J. G. & Loiselle, B. A. Diversity of birds along an elevational gradient in the Cordillera Central, Costa Rica. Auk 117, 663–686 (2000).
    Article  Google Scholar 

    35.
    Ibarra, J. T., Martin, M., Cockle, K. L. & Martin, K. Maintaining ecosystem resilience: Functional responses of tree cavity nesters to logging in temperate forests of the Americas. Sci. Rep. 7, 1–9 (2017).
    CAS  Article  Google Scholar 

    36.
    Ibarra, J. T. et al. Seasonal dynamics of avian guilds inside and outside core protected areas in an Andean Biosphere Reserve of southern Chile. Bird Study 64, 410 (2017).
    Article  Google Scholar 

    37.
    Lindenmayer, D. B. Conserving large old trees as small natural features. Biol. Conserv. 211, 51–59 (2017).
    Article  Google Scholar 

    38.
    Salguero-Gómez, R., Violle, C., Gimenez, O. & Childs, D. Delivering the promises of trait-based approaches to the needs of demographic approaches, and vice versa. Funct. Ecol. 32, 1424–1435. https://doi.org/10.1111/1365-2435.13148 (2018).
    Article  PubMed  PubMed Central  Google Scholar 

    39.
    Stork, N. E. et al. Consistency of effects of tropical-forest disturbance on species composition and richness relative to use of indicator taxa. Conserv. Biol. 31, 924–933 (2017).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    40.
    Ibarra, J. T. & Martin, K. Beyond species richness: an empirical test of top predators as surrogates for functional diversity and endemism. Ecosphere 6, 1–15 (2015).
    Article  Google Scholar 

    41.
    Jankowski, J. E., Robinson, S. K. & Levey, D. J. Squeezed at the top: interspecific aggression may constrain elevational ranges in tropical birds. Ecology 91, 1877–1884 (2010).
    PubMed  Article  PubMed Central  Google Scholar 

    42.
    Mouchet, M. A., Villéger, S., Mason, N. W. H. & Mouillot, D. Functional diversity measures: an overview of their redundancy and their ability to discriminate community assembly rules. Funct. Ecol. 24, 867–876 (2010).
    Article  Google Scholar 

    43.
    Huey, R. B., Hertz, P. E. & Sinervo, B. Behavioral drive versus behavioral inertia in evolution: a null model approach. Am. Nat. 161, 357–366 (2003).
    PubMed  Article  PubMed Central  Google Scholar 

    44.
    Martin, T. E. Evolutionary determinants of clutch size in cavity-nesting birds: nest predation or limited breeding opportunities?. Am. Nat. 142, 937–946 (1993).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    45.
    Jackson, M. M., Gergel, S. E. & Martin, K. Effects of climate change on habitat availability and configuration for an endemic coastal alpine bird. PLoS ONE 10(11), e0142110. https://doi.org/10.1371/journal.pone.0146838 (2015).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    46.
    Martin, K. The ecological values of mountain environments and wildlife. In The Impacts of Skiing and Related Winter Recreational Activities on Mountain Environments (eds Rixen, C. & Rolando, A.) 3–29 (Bentham Science Publishers, Sharjah, 2013). https://doi.org/10.2174/9781608054886113010004.
    Google Scholar 

    47.
    Chen, I.-C., Hill, J. K., Ohlemüller, R., Roy, D. B. & Thomas, C. D. Rapid range shifts of species associated with high levels of climate warming. Science (80-.) 333, 1024–1026 (2011).
    ADS  CAS  Article  Google Scholar 

    48.
    Harsch, M. A., Hulme, P. E., McGlone, M. S. & Duncan, R. P. Are treelines advancing? A global meta-analysis of treeline response to climate warming. Ecol. Lett. 12, 1040–1049 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    49.
    Braunisch, V. et al. Temperate mountain forest biodiversity under climate change: Compensating negative effects by increasing structural complexity. PLoS ONE 9, e97718 (2014).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    50.
    Boyle, W. A. & Martin, K. The conservation value of high elevation habitats to North American migrant birds. Biol. Conserv. 192, 461–476 (2015).
    Article  Google Scholar 

    51.
    Buckland, S. T. et al. Introduction to Distance Sampling: Estimating Abundance of Biological Populations (Oxford University Press, Oxford, 2001).
    Google Scholar 

    52.
    Bibby, C. J., Burgess, N. D., Hill, D. A. & Mustoe, S. Bird Census Techniques (Academic Press, Cambridge, 2000).
    Google Scholar 

    53.
    Petchey, O. L. & Gaston, K. J. Functional diversity: back to basics and looking forward. Ecol. Lett. 9, 741–758 (2006).
    PubMed  Article  PubMed Central  Google Scholar 

    54.
    Laiolo, P., Pato, J. & Obeso, J. R. Ecological and evolutionary drivers of the elevational gradient of diversity. Ecol. Lett. 21, 1022–1032 (2018).
    PubMed  Article  PubMed Central  Google Scholar 

    55.
    Ikin, K. et al. Avian functional responses to landscape recovery. Proc. R. Soc. B Biol. Sci. 286, 20190114 (2019).
    Article  Google Scholar 

    56.
    Barbe, L., Morel, R., Rantier, Y., Lebas, J. F. & Butet, A. Bird communities of a temperate forest: spatio-temporal partitioning between resident and migratory species. J. Ornithol. 159, 457–469 (2018).
    Article  Google Scholar 

    57.
    Luck, G. W., Carter, A. & Smallbone, L. Changes in bird functional diversity across multiple land uses: interpretations of functional redundancy depend on functional group identity. PLoS ONE 8, e63671 (2013).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    58.
    Brown, J. H., Gillooly, J. F., Allen, A. P., Savage, V. M. & West, G. B. Toward a metabolic theory of ecology. Ecology 85, 1771–1789 (2004).
    Article  Google Scholar 

    59.
    Woodward, G. et al. Body size in ecological networks. Trends Ecol. Evol. 20, 402–409 (2005).
    PubMed  Article  PubMed Central  Google Scholar 

    60.
    Royle, J. A., Dawson, D. K. & Bates, S. Modelling abundance effects in distance sampling. Ecology 85, 1591–1597 (2004).
    Article  Google Scholar 

    61.
    Marques, T. A., Thomas, L., Fancy, S. G. & Buckland, S. T. Improving estimates of bird density using multiple-covariate distance sampling. Auk 124, 1229–1243 (2007).
    Article  Google Scholar 

    62.
    Chandler, R. Distance Sampling Analysis in Unmarked (ISGS Patuxent Wildlife Research Center, Laurel, 2014).
    Google Scholar 

    63.
    Harrison, X. A. et al. A brief introduction to mixed effects modelling and multi-model inference in ecology. PeerJ 6, 1–32. https://doi.org/10.7717/peerj.4794 (2018).
    Article  Google Scholar 

    64.
    Drever, M. C. & Martin, K. Response of woodpeckers to changes in forest health and harvest: Implications for conservation of avian biodiversity. For. Ecol. Manage. 259, 958–966 (2010).
    Article  Google Scholar 

    65.
    Burnham, K. P. & Anderson, D. R. Model Selection and Multimodel Inference: A Practical Information-Theoretic Approach. Ecological Modelling Vol. 172 (Springer-Verlag, Berlin, 2002).
    Google Scholar 

    66.
    Mason, N. W. H., Mouillot, D., Lee, W. G. & Wilson, J. B. Functional richness, functional evenness and functional divergence: the primary components of functional diversity. Oikos 111, 112–118 (2005).
    Article  Google Scholar 

    67.
    De Cáceres, M., Legendre, P. & He, F. Dissimilarity measurements and the size structure of ecological communities. Methods Ecol. Evol. 4, 1167–1177 (2013).
    Article  Google Scholar 

    68.
    Anderson, M. J. Distance-based tests for homogeneity of multivariate dispersions. Biometrics 62, 245–253 (2006).
    MathSciNet  PubMed  MATH  Article  PubMed Central  Google Scholar 

    69.
    Laliberté, E. & Legendre, P. A distance-based framework for measuring functional diversity from multiple traits. Ecology 91, 299–305 (2010).
    PubMed  Article  PubMed Central  Google Scholar 

    70.
    R Core Team. R: a language and environment for statistical computing (2018).

    71.
    Jaksic, F. M. & Feinsinger, P. Bird assemblages in temperate forests of North and South America: a comparison of diversity, dynamics, guild structure, and resource use. Rev. Chil. Hist. Nat. 64, 491–510 (1991).
    Google Scholar 

    72.
    Martínez, D. & González, G. Aves de Chile. Guías de campo y breve historia natural (Ediciones del Naturalista, Madrid, 2017).
    Google Scholar 

    73.
    Altamirano, T. A. et al. Hábitos de nidificación de las aves del bosque templado andino de Chile (2012).

    74.
    Jaramillo, A., Burke, P. & Beadle, D. Birds of Chile (Christopher Helm, London, 2003).
    Google Scholar 

    75.
    Altamirano, T. A., Ibarra, J. T., de la Maza, M., Navarrete, S. A. & Bonacic, C. Reproductive life-history variation in a secondary cavity-nester across an elevational gradient in Andean temperate ecosystems. Auk 132, 826–835 (2015).
    Article  Google Scholar 

    76.
    Medrano, F., Barros, R., Norambuena, H. V., Matus, R. & Schmitt, F. Atlas de las aves nidificantes de Chile (Red de Observadores de Aves y Vida Silvestre de Chile, Chile, 2018).
    Google Scholar 

    77.
    del Hoyo, J., Elliott, A., Sargatal, J., Christie, D. A. & Kirwan, G. Handbook of the birds of the World Alive (Lynx Editions, Barcelona, 2019).
    Google Scholar 

    78.
    Dunning, J. B. CRC Handbook of Avian Body Masses (CRC Press, London, 2007).
    Google Scholar  More

  • in

    Drivers of tropical soil invertebrate community composition and richness across tropical secondary forests using DNA metasystematics

    1.
    Trumbore, S., Brando, P. & Hartmann, H. Forest health and global change. Science 349, 814–818 (2015).
    ADS  CAS  PubMed  Article  Google Scholar 
    2.
    Gibbs, H. K. et al. Tropical forests were the primary sources of new agricultural land in the 1980s and 1990s. Proc. Natl. Acad. Sci. 107, 16732–16737 (2010).
    ADS  CAS  PubMed  Article  Google Scholar 

    3.
    Laurance, W. F. Have we overstated the tropical biodiversity crisis?. Trends Ecol. Evol. 22, 65–70 (2007).
    PubMed  Article  Google Scholar 

    4.
    Brienen, R. J. W. et al. Long-term decline of the Amazon carbon sink. Nature 519, 344–348 (2015).
    ADS  CAS  PubMed  Article  Google Scholar 

    5.
    Nottingham, A. T. et al. Climate warming and soil carbon in tropical forests: insights from an elevation gradient in the Peruvian Andes. Bioscience 65, 906–921 (2015).
    PubMed  PubMed Central  Article  Google Scholar 

    6.
    Seymour, F. & Busch, J. Why Forests? Why Now? The Science, Economics and Politics of Tropical Forests and Climate Change 1–450 (Center for Global Development, Washington, 2016).
    Google Scholar 

    7.
    de Quadros, P. D. et al. Coal mining practices reduce the microbial biomass, richness and diversity of soil. Appl. Soil Ecol. 98, 195–203 (2016).
    Article  Google Scholar 

    8.
    Beer, C. et al. Terrestrial gross carbon dioxide uptake: global distribution and covariation with climate. Science 329, 834–838 (2010).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    9.
    Jobbágy, E. G. & Jackson, R. B. The vertical distribution of soil organic carbon and its relation to climate and vegetation. Ecol. Appl. 10, 423–426 (2000).
    Article  Google Scholar 

    10.
    Chazdon, R. L. et al. Carbon sequestration potential of second-growth forest regeneration in the Latin American tropics. Sci. Adv. 2, e1501639–e1501639 (2016).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    11.
    Chazdon, R. L. Beyond deforestation: restoring forests and ecosystem services on degraded lands. Science 320, 1458–1460 (2008).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    12.
    Chazdon, R. L. Tropical forest recovery: legacies of human impact and natural disturbances. Perspect. Plant Ecol. 6, 51–71 (2003).
    Article  Google Scholar 

    13.
    Arroyo-Rodríguez, V. et al. Multiple successional pathways in human-modified tropical landscapes: new insights from forest succession, forest fragmentation and landscape ecology research. Biol. Rev. 92, 326–340 (2017).
    PubMed  Article  PubMed Central  Google Scholar 

    14.
    Vashum, K. T., Kasomwoshi, T. & Jayakumar, S. Soil organic carbon sequestration potential of primary and secondary forests in Northeast India. Proc. Int. Acad. Ecol. Environ. Sci. 6, 67 (2016).
    CAS  Google Scholar 

    15.
    Uriarte, M. et al. Natural disturbance and human land use as determinants of tropical forest dynamics: results from a forest simulator. New Phytol. 79, 423–443 (2009).
    Google Scholar 

    16.
    Calderón, K. et al. Effectiveness of ecological rescue for altered soil microbial communities and functions. ISME J. 11, 272–283 (2017).
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    17.
    Rodrigues, J. L. M. et al. Conversion of the Amazon rainforest to agriculture results in biotic homogenization of soil bacterial communities. Proc. Natl. Acad. Sci. USA 110, 988–993 (2013).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    18.
    Coyle, D. R. et al. Soil fauna responses to natural disturbances, invasive species, and global climate change: current state of the science and a call to action. Soil Biol. Biochem. 110, 116–133 (2017).
    CAS  Article  Google Scholar 

    19.
    Brussaard, L. Biodiversity and ecosystem functioning in soil. Ambio 26, 563–570 (1997).
    Google Scholar 

    20.
    Brussaard, L., de Ruiter, P. C. & Brown, G. G. Soil biodiversity for agricultural sustainability. Agric. Ecosyst. Environ. 121, 233–244 (2007).
    Article  Google Scholar 

    21.
    Lavelle, P. et al. Soil invertebrates and ecosystem services. Eur. J. Soil Biol. 42, S3–S15 (2006).
    Article  Google Scholar 

    22.
    Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet J. 17, 10–12 (2011).
    Article  Google Scholar 

    23.
    Eisenhauer, N., Sabais, A. C. W. & Scheu, S. Collembola species composition and diversity effects on ecosystem functioning vary with plant functional group identity. Soil Biol. Biochem. 43, 1697–1704 (2011).
    CAS  Article  Google Scholar 

    24.
    Griffiths, B. S., de Groot, G. A., Laros, I., Stone, D. & Geisen, S. The need for standardisation: exemplified by a description of the diversity, community structure and ecological indices of soil nematodes. Ecol. Indic. 87, 43–46 (2018).
    Article  Google Scholar 

    25.
    Mulder, C. Driving forces from soil invertebrates to ecosystem functioning: the allometric perspective. Naturwissenschaften 93, 467–479 (2006).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    26.
    Menta, C. Soil fauna diversity. In Biodiversity Conservation and Utilization in a Diverse World (ed. Lameed, G. A.) (InTech, London, 2012). https://doi.org/10.5772/51091.
    Google Scholar 

    27.
    Errington, I. et al. The influence of vegetation and soil properties on springtail communities in a diesel-contaminated soil. Sci. Total Environ. 619–620, 1098–1104 (2018).
    ADS  PubMed  Article  CAS  PubMed Central  Google Scholar 

    28.
    Korboulewsky, N., Perez, G. & Chauvat, M. How tree diversity affects soil fauna diversity: a review. Soil Biol. Biochem. 94, 94–106 (2016).
    CAS  Article  Google Scholar 

    29.
    Beng, K. C. et al. The utility of DNA metabarcoding for studying the response of arthropod diversity and composition to land-use change in the tropics. Sci. Rep. 6, 1–13 (2016).
    Article  CAS  Google Scholar 

    30.
    McGee, K. M., Robinson, C. V. & Hajibabaei, M. Gaps in DNA-based biomonitoring across the globe. Front. Ecol. Evolut. 7, 337 (2019).
    Article  Google Scholar 

    31.
    Ji, Y. et al. Reliable, verifiable and efficient monitoring of biodiversity via metabarcoding. Ecol. Lett. 16, 1245–1257 (2013).
    PubMed  Article  PubMed Central  Google Scholar 

    32.
    Liu, S. et al. SOAP Barcode: revealing arthropod biodiversity through assembly of Illumina shotgun sequences of PCR amplicons. Methods Ecol. Evol. 4, 1142–1150 (2013).
    Article  Google Scholar 

    33.
    Bienert, F. et al. Tracking earthworm communities from soil DNA: tracking earthworm communities from soil DNA. Mol. Ecol. 21, 2017–2030 (2012).
    CAS  PubMed  Article  Google Scholar 

    34.
    Dell’Anno, A., Carugati, L., Corinaldesi, C., Riccioni, G. & Danovaro, R. Unveiling the biodiversity of deep-sea nematodes through metabarcoding: are we ready to bypass the classical taxonomy?. PLoS ONE 10, e0144928 (2015).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    35.
    Fonseca, V. G. et al. Metagenetic analysis of patterns of distribution and diversity of marine meiobenthic eukaryotes: macroecology of microscopic marine eukaryotes. Glob. Ecol. Biogeogr. 23, 1293–1302 (2014).
    Article  Google Scholar 

    36.
    Pansu, J. et al. Landscape-scale distribution patterns of earthworms inferred from soil DNA. Soil Biol. Biochem. 83, 100–105 (2015).
    CAS  Article  Google Scholar 

    37.
    Saitoh, S. et al. A quantitative protocol for DNA metabarcoding of springtails (Collembola). Genome 59, 705–723 (2016).
    CAS  PubMed  Article  Google Scholar 

    38.
    Baird, D. J. & Hajibabaei, M. Biomonitoring 2.0: a new paradigm in ecosystem assessment made possible by next-generation DNA sequencing: news and views: opinion. Mol. Ecol. 21, 2039–2044 (2012).
    PubMed  Article  Google Scholar 

    39.
    Edge, T. A. et al. The ecobiomics project: advancing metagenomics assessment of soil health and freshwater quality in Canada. Sci. Total Environ. 710, 135906 (2020).
    ADS  CAS  PubMed  Article  Google Scholar 

    40.
    Porter, T. M. & Hajibabaei, M. Scaling up: a guide to high-throughput genomic approaches for biodiversity analysis. Mol. Ecol. 27, 313–338 (2018).
    PubMed  Article  PubMed Central  Google Scholar 

    41.
    Porter, T. M. & Hajibabaei, M. Automated high throughput animal CO1 metabarcode classification. Sci. Rep. 8, 1–10 (2018).
    Article  CAS  Google Scholar 

    42.
    Porter, T. M. & Hajibabaei, M. Over 2.5 million COI sequences in GenBank and growing. PLoS ONE 13, e0200177 (2018).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    43.
    Fernandes, K. et al. DNA metabarcoding-a new approach to fauna monitoring in mine site restoration. Restor. Ecol. 26, 1098–1107 (2018).
    Article  Google Scholar 

    44.
    Gibson, J. F. et al. Large-scale biomonitoring of remote and threatened ecosystems via high-throughput sequencing. PLoS ONE 10, e0138432-e138515 (2015).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    45.
    Martin, G. K., Adamowicz, S. J. & Cottenie, K. Taxonomic resolution based on DNA barcoding affects environmental signal in metacommunity structure. Freshw. Sci. 35, 701–711 (2016).
    Article  Google Scholar 

    46.
    Banerjee, S. et al. Determinants of bacterial communities in Canadian agroforestry systems. Environ. Microbiol. 18, 1805–1816 (2016).
    PubMed  Article  PubMed Central  Google Scholar 

    47.
    Chikoski, J. M., Ferguson, S. H. & Meyer, L. Effects of water addition on soil arthropods and soil characteristics in a precipitation-limited environment. Acta Oecologica 30, 203–211 (2006).
    ADS  Article  Google Scholar 

    48.
    Wall, D. H. et al. Soil Ecology and Ecosystem Services (Oxford University Press, Oxford, 2012).
    Google Scholar 

    49.
    Yu, Y. et al. Shifts in microbial community function and structure along the successional gradient of coastal wetlands in Yellow River Estuary. Eur. J. Soil Biol. 49, 12–21 (2012).
    Article  Google Scholar 

    50.
    Convey, P., Block, W. & Peat, H. J. Soil arthropods as indicators of water stress in Antarctic terrestrial habitats?. Glob. Change Biol. 9, 1718–1730 (2003).
    ADS  Article  Google Scholar 

    51.
    Marra, J. L. & Edmonds, R. L. Soil arthropod responses to different patch types in a mixed-conifer forest of the Sierra Nevada. For. Sci. 51, 255–265 (1998).
    Google Scholar 

    52.
    McCluney, K. E. & Sabo, J. L. Sensitivity and tolerance of riparian arthropod communities to altered water resources along a drying river. PLoS ONE 9, e109276 (2014).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    53.
    Villani, M. G. & Wright, R. J. Environmental influences on soil macroarthropod behavior in agricultural systems 35, 249–269 (1990).
    Google Scholar 

    54.
    Zahran, H. H., Moharram, A. M. & Mohammad, H. A. Some ecological and physiological studies on bacteria isolated from salt-affected soils of Egypt. J. Basic Microbiol. 32, 405–413 (1992).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    55.
    McGee, K. M., Eaton, W. D., Shokralla, S. & Hajibabaei, M. Determinants of soil bacterial and fungal community composition toward carbon-use efficiency across primary and secondary forests in a costa rican conservation area. Microb. Ecol. 10, 423 (2018).
    Google Scholar 

    56.
    Oba, Y., Ôhira, H., Murase, Y., Moriyama, A. & Kumazawa, Y. DNA barcoding of Japanese click beetles (Coleoptera, Elateridae). PLoS ONE 10, e0116612 (2015).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    57.
    Ensafi, P. et al. Soil type mediates the effectiveness of biological control against Limonius californicus (Coleoptera: Elateridae). J. Econ. Entomol. 111, 2053–2058 (2018).
    CAS  PubMed  PubMed Central  Google Scholar 

    58.
    Poggi, S. et al. Relative influence of climate and agroenvironmental factors on wireworm damage risk in maize crops. J. Pest. Sci. 91, 585–599 (2018).
    Article  Google Scholar 

    59.
    Di, S., Huang, L., Diao, J. & Zhou, Z. Selective bioaccumulation and elimination of hexachlorocyclohexane isomers in Tubifex tubifex (Oligochaeta, Tubificidae). Environ. Sci. Pollut. Res. 23, 6990–6998 (2016).
    CAS  Article  Google Scholar 

    60.
    Pelegrí, S. & Blackburn, T. H. Effects of Tubifex tubifex (Oligochaeta:Tubificidae) on N-mineralization in freshwater sediments, measured with 15-N isotopes. Aquat. Microb. Ecol. 9, 289–294 (1995).
    Article  Google Scholar 

    61.
    Saaltink, R. M. Respiration and aeration by bioturbating Tubificidae alter biogeochemical processes in aquatic sediment. Aquat. Sci. 81, 1–13 (2019).
    Article  CAS  Google Scholar 

    62.
    Booth, M. S., Stark, J. M. & Rastetter, E. Controls on nitrogen cycling in terrestrial ecosystems: a synthetic analysis of literature data. New Phytol. 75, 139–157 (2005).
    Google Scholar 

    63.
    Feldpausch, T. R., Rondon, M. A., Fernandes, E. C. M., Riha, S. J. & Wandelli, E. Carbon and nutrient accumulation in secondary forests regenerating on pastures in central Amazonia. Ecol. Appl. 14, 164–176 (2004).
    Article  Google Scholar 

    64.
    Gehring, C., Vlek, P. L. G., de Souza, L. A. G. & Denich, M. Biological nitrogen fixation in secondary regrowth and mature rainforest of central Amazonia. Agric. Ecosyst. Environ. 111, 237–252 (2005).
    CAS  Article  Google Scholar 

    65.
    Guariguata, M. R. & Ostertag, R. Neotropical secondary forest succession: changes in structural and functional characteristics. For. Ecol. Manag. 148, 185–206 (2001).
    Article  Google Scholar 

    66.
    Chassot, O. & Monge, G. Connectivity conservation of the great green macaw’s landscape in costa rica and nicaragua (1994–2012). Parks 18, 1–10 (2012).
    Google Scholar 

    67.
    Hartshorn, G. S. et al. Vegetation Types and Floristic Patterns. La Selva: Ecology and Natural History of a Neotropical Rain Forest (The University of Chicago Press, Chicago, 1994).
    Google Scholar 

    68.
    van der Gast, C. J., Gosling, P., Tiwari, B. & Bending, G. D. Spatial scaling of arbuscular mycorrhizal fungal diversity is affected by farming practice: spatial scaling of arbuscular mycorrhizal fungi. Environ. Microbiol. 13, 241–249 (2011).
    PubMed  Article  PubMed Central  Google Scholar 

    69.
    McGee, K. M., Eaton, W. D., Porter, T. M., Shokralla, S. & Hajibabaei, M. Soil microbiomes associated with two dominant Costa Rican tree species, and implications for remediation. A case study from a Costa Rican conservation area. Appl. Soil Ecol. 137, 139–153 (2019).
    Article  Google Scholar 

    70.
    Eaton, W. D. et al. Differences in the soil microbial community and carbon-use efficiency following development of Vochysia guatemalensis tree plantations in unproductive pastures in Costa Rica. Restor. Ecol. https://doi.org/10.1111/rec.12978 (2019).
    Article  Google Scholar 

    71.
    Folmer, O., Black, M., Hoeh, W., Lutz, R. & Vrijenhoek, R. DNA primers for amplification of mitochondrial cytochrome c oxidase subunit I from diverse metazoan invertebrates. Mol. Mar. Biol. Biochem. 3, 294–299 (1994).
    CAS  Google Scholar 

    72.
    Hajibabaei, M., Spall, J. L., Shokralla, S. & van Konynenburg, S. Assessing biodiversity of a freshwater benthic macroinvertebrate community through non-destructive environmental barcoding of DNA from preservative ethanol. BMC Ecol. 12, 28 (2012).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    73.
    Gibson, J. et al. Simultaneous assessment of the macrobiome and microbiome in a bulk sample of tropical arthropods through DNA metasystematics. Proc. Nat. Acad. Sci. 111, 8007–8012 (2014).
    ADS  CAS  PubMed  Article  Google Scholar 

    74.
    St. John, J. SeqPrep. Retrieved https://github.com/jstjohn/SeqPrep (2016).

    75.
    Edgar, R. C. UNOISE2: improved error-correction for Illumina 16S and ITS amplicon sequencing. bioRxiv. https://doi.org/10.1101/081257 (2016).
    Article  Google Scholar 

    76.
    Callahan, B. J., McMurdie, P. J. & Holmes, S. P. Exact sequence variants should replace operational taxonomic units in marker-gene data analysis. ISME J. 11, 2639–2643 (2017).
    PubMed  PubMed Central  Article  Google Scholar 

    77.
    Reeder, J. & Knight, R. The ‘rare biosphere’: a reality check. Nat. Methods 6, 636–637 (2009).
    CAS  PubMed  Article  Google Scholar 

    78.
    Tedersoo, L. et al. 454 Pyrosequencing and Sanger sequencing of tropical mycorrhizal fungi provide similar results but reveal substantial methodological biases. New Phytol. 188, 291–301 (2010).
    CAS  PubMed  Article  Google Scholar 

    79.
    Weiss, S. et al. Normalization and microbial differential abundance strategies depend upon data characteristics. Microbiome 5, 27 (2017).
    PubMed  PubMed Central  Article  Google Scholar 

    80.
    Anderson, M. J., Gorley, R. N. & Clarke, R. K. PERMANOVA+ for Primer: guide to Software and Statistical Methods (PRIMER-E Ltd., Plymouth, UK, 2008).
    Google Scholar 

    81.
    Clarke, K. R. Non-parametric multivariate analyses of changes in community structure. Aust. J. Ecol. 18, 117–143 (1993).
    Article  Google Scholar 

    82.
    Anderson, M. J. & Willis, T. J. Canonical analysis of principal coordinates: a useful method of constrained ordination for ecology. Ecology 84, 511–525 (2003).
    Article  Google Scholar 

    83.
    DiStefano, J., Fidler, F. & Cumming, G. Effect size estimates and confidence intervals: An alternative focus for the presentation and interpretation of ecological data. In New Trends in Ecology Research. 1st edn. (ed. Burk, A.). (Nova Science Publishers Inc., New York, 2005).
    Google Scholar 

    84.
    Clarke, K. R., & Gorley, R. N. PRIMER v6: user manual/tutorial (Plymouth routines in multivariate ecological research) (PRIMER-E Ltd., Plymouth, UK, 2006).
    Google Scholar 

    85.
    Burnham, K. P. & Anderson, D. R. Model selection and multimodel inference: a practical information-theoretic approach. In Model Selection and Multimodel Inference (eds Burnham, K. P. & Anderson, D. R.) 1–515 (Springer, Berlin, 2003).
    Google Scholar 

    86.
    Legendre, P. & Anderson, M. J. Distance-based redundancy analysis: testing multispecies responses in multifactorial ecological experiments. New Phytol. 69, 1–24 (1999).
    Google Scholar  More

  • in

    Changes in coupled carbon‒nitrogen dynamics in a tundra ecosystem predate post-1950 regional warming

    1.
    Hugelius, G. et al. Estimated stocks of circumpolar permafrost carbon with quantified uncertainty ranges and identified data gaps. Biogeosciences 11, 6573–6593 (2014).
    Article  Google Scholar 
    2.
    Olefeldt, D. et al. Circumpolar distribution and carbon storage of thermokarst landscapes. Nat. Commun. 7, https://doi.org/10.1038/ncomms13043 (2016).

    3.
    McGuire, A. D. et al. Sensitivity of the carbon cycle in the Arctic to climate change. Ecol. Monogr. 79, 523–555 (2009).
    Article  Google Scholar 

    4.
    Schuur, E. A. G. et al. Climate change and the permafrost carbon feedback. Nature 520, 171–179 (2015).
    CAS  Article  Google Scholar 

    5.
    Chapin, D. M. & Bledsoe, C. S. Nitrogen fixation in Arctic plant communitiesin (eds. Chapin, F. S., Jefferies, R. L., Reynolds, J. F., Shaver, G. R. & Svoboda, J.) Arctic Ecossytems in a changing cliamte: an ecophysiological perspective 1992.

    6.
    Chapin, F. S. III. et al. The changing global carbon cycle: linking plant-soil carbon dynamics to global consequences. J. Ecol. 97, 840–850 (2009).
    CAS  Article  Google Scholar 

    7.
    Elser, J. J. et al. Global analysis of nitrogen and phosphorus limitation of primary producers in freshwater, marine and terrestrial ecosystems. Ecol. Lett. 10, 1135–1142 (2007).
    Article  Google Scholar 

    8.
    Wookey, P. A. et al. Ecosystem feedbacks and cascade processes: understanding their role in the responses of Arctic and alpine ecosystems to environmental change. Glob. Change Biol. 15, 1153–1172 (2009).
    Article  Google Scholar 

    9.
    Marin-Spiotta, E. et al. Paradigm shifts in soil organic matter research affect interpretations of aquatic carbon cycling: transcending disciplinary and ecosystem boundaries. Biogeochemistry 117, 279–297 (2014).
    CAS  Article  Google Scholar 

    10.
    Tape, K. D., Hallinger, M., Welker, J. M. & Ruess, R. W. Landscape heterogeneity of shrub expansion in arctic Alaska. Ecosystems 15, 711–724 (2012).
    CAS  Article  Google Scholar 

    11.
    Myrstener, M. et al. Persistent nitrogen limitation of stream biofilm communities along climate gradients in the Arctic. Glob Change Biol. 24, 3680–3691 (2018).
    Article  Google Scholar 

    12.
    Salmon, V. G. et al. Nitrogen availability increases in a tundra ecosystem during five years of experimental permafrost thaw. Glob. Change Biol. 22, 1927–1941 (2016).
    Article  Google Scholar 

    13.
    Kling, G. W., Kipphut, G. W. & Miller, M. C. Arctic lakes and streams as gas conduits to the atmosphere-implications for tundra carbon budgets. Science 251, 298–301 (1991).
    CAS  Article  Google Scholar 

    14.
    Kling, G. W., Kipphut, G. W., Miller, M. M. & O’Brien, W. J. Integration of lakes and streams in a landscape perspective: the importance of material processing on spatial patterns and temporal coherence. Freshwater Biol. 43, 477–497 (2000).
    Article  Google Scholar 

    15.
    DelSontro, T., Beaulieu, J. J. & Downing, J. A. Greenhouse gas emissions from lakes and impoundments: upscaling in the face of global change. Limnol. Oceanogr. Lett. 3, 64–75 (2018).
    CAS  Article  Google Scholar 

    16.
    Bring, A. et al. Arctic terrestrial hydrology: A synthesis of processes, regional effects, and research challenges. J. Geophys. Res. Biogeosci. 121, 621–649 (2016).
    Article  Google Scholar 

    17.
    Tank, S. E. et al. Landscape-level controls on dissolved carbon flux from diverse catchments of the circumboreal. Global Biogeochem. Cycles 26, https://doi.org/10.1029/2012gb004299 (2012).

    18.
    Kling, G. W., Kipphut, G. W. & Miller, M. C. The flux of CO2 and CH4 from lakes and rivers in arctic Alaska. Hydrobiologia 240, 23–36 (1992).
    CAS  Article  Google Scholar 

    19.
    Cory, R. M., Ward, C. P., Crump, B. C. & Kling, G. W. Sunlight controls water column processing of carbon in arctic fresh waters. Science 345, 925–928 (2014).
    CAS  Article  Google Scholar 

    20.
    Kortelainen, P. et al. Carbon evasion/accumulation ratio in boreal lakes is linked to nitrogen. Global Biogeochem. Cycles 27, 363–374 (2013).
    CAS  Article  Google Scholar 

    21.
    Schuur, E. A. G. et al. The effect of permafrost thaw on old carbon release and net carbon exchange from tundra. Nature 459, 556–559 (2009).
    CAS  Article  Google Scholar 

    22.
    Bowden, W. B. et al. Sediment and nutrient delivery from thermokarst features in the foothills of the North Slope, Alaska: Potential impacts on headwater stream ecosystems. J. Geophys. Res. Biogeosci. 113, https://doi.org/10.1029/2007jg000470 (2008).

    23.
    Curtis, J., Wendler, G., Stone, R. & Dutton, E. Precipitation decrease in the western arctic, with special emphasis on Barrow and Barter Island, Alaska. Int. J. Climatol. 18, 1687–1707 (1998).
    Article  Google Scholar 

    24.
    Yano, Y., Shaver, G. R., Giblin, A. E., Rastetter, E. B. & Nadelhoffer, K. J. Nitrogen dynamics in a small arctic watershed: retention and downhill movement of N-15. Ecol. Monogr. 80, 331–351 (2010).
    Article  Google Scholar 

    25.
    Tape, K., Sturm, M. & Racine, C. The evidence for shrub expansion in Northern Alaska and the Pan-Arctic. Glob. Change Biol. 12, 686–702 (2006).
    Article  Google Scholar 

    26.
    Stackpoole, S. M. et al. Inland waters and their role in the carbon cycle of Alaska. Ecol. Appl. 27, 1403–1420 (2017).
    Article  Google Scholar 

    27.
    Tank, S. E. et al. Landscape matters: Predicting the biogeochemical effects of permafrost thaw on aquatic networks with a state factor approach. Permafr. Periglac. Process. 31, 358–370 (2020).
    Article  Google Scholar 

    28.
    Hobbie, J. E. & Kling, G. W. Alaska’s Changing Arctic: Ecological Consequences for Tundra, Streams, and Lakes. (Oxford University Press, 2014).

    29.
    Kaufman, D. S. et al. Recent warming reverses long-term Arctic cooling. Science 325, 1236–1239 (2009).
    CAS  Article  Google Scholar 

    30.
    Meyers, P. A. & Ishiwatari, R. Lacustrine organic geochemistry-an overview of indicators of organic matter sources and diagenesis in lake sediments. Organ. Geochem. 20, 867–900 (1993).
    CAS  Article  Google Scholar 

    31.
    Vadeboncoeur, Y. et al. From Greenland to green lakes: cultural eutrophication and the loss of benthic pathways in lakes. Limnol. Oceanogr. 48, 1408–1418 (2003).
    Article  Google Scholar 

    32.
    Gettel, G. M., Giblin, A. E. & Howarth, R. W. Controls of benthic nitrogen fixation and primary production from nutrient enrichment of oligotrophic, Arctic lakes. Ecosystems 16, 1550–1564 (2013).
    CAS  Article  Google Scholar 

    33.
    Cornwell, J. C. & Banahan, S. A silicon budget for an alaskan arctic lake. Hydrobiologia 240, 37–44 (1992).
    CAS  Article  Google Scholar 

    34.
    France, R. L. C-13 enrichment in benthic compared to planktonic algae-foodweb implications. Mar. Ecol. Prog. Ser. 124, 307–312 (1995).
    Article  Google Scholar 

    35.
    Levine, M. A. & Whalen, S. C. Nutrient limitation of phytoplankton production in Alaskan Arctic foothill lakes. Hydrobiologia 455, 189–201 (2001).
    Article  Google Scholar 

    36.
    Hobbie, S. E., Schimel, J. P., Trumbore, S. E. & Randerson, J. R. Controls over carbon storage and turnover in high-latitude soils. Glob. Change Biol. 6, 196–210 (2000).
    Article  Google Scholar 

    37.
    Galman, V., Rydberg, J., de-Luna, S. S., Bindler, R. & Renberg, I. Carbon and nitrogen loss rates during aging of lake sediment: changes over 27 years studied in varved lake sediment. Limnol. Oceanogr. 53, 1076–1082 (2008).
    Article  Google Scholar 

    38.
    Fitzgerald, W. F. et al. Modern and historic atmospheric mercury fluxes in northern Alaska: Global sources and Arctic depletion. Environ. Sci. Technol. 39, 557–568 (2005).
    CAS  Article  Google Scholar 

    39.
    Holtgrieve, G. W. et al. A coherent signature of anthropogenic nitrogen deposition to remote watersheds of the northern hemisphere. Science 334, 1545–1548 (2011).
    CAS  Article  Google Scholar 

    40.
    Hobbs, W. O. et al. Nitrogen deposition to lakes in national parks of the western Great Lakes region: Isotopic signatures, watershed retention, and algal shifts. Global Biogeochem. Cycles 30, 514–533 (2016).
    CAS  Article  Google Scholar 

    41.
    Anderson, N. J. et al. Regional variability in the atmospheric nitrogen deposition signal and its transfer to the sediment record in Greenland lakes. Limnol. Oceanogr. 63, 2250–2265 (2018).
    CAS  Article  Google Scholar 

    42.
    Brock, C. S., Leavitt, P. R., Schindler, D. E., Johnson, S. P. & Moore, J. W. Spatial variability of stable isotopes and fossil pigments in surface sediments of Alaskan coastal lakes: Constraints on quantitative estimates of past salmon abundance. Limnol. Oceanogr. 51, 1637–1647 (2006).
    CAS  Article  Google Scholar 

    43.
    Curtis, C. J. et al. Spatial variations in snowpack chemistry, isotopic composition of NO3- and nitrogen deposition from the ice sheet margin to the coast of western Greenland. Biogeosciences 15, 529–550 (2018).
    CAS  Article  Google Scholar 

    44.
    Post, E. et al. Ecological dynamics across the Arctic associated with recent climate change. Science 325, 1355–1358 (2009).
    CAS  Article  Google Scholar 

    45.
    Jones, V. J. et al. The influence of Holocene tree-line advance and retreat on an arctic lake ecosystem: a multi-proxy study from Kharinei Lake, North Eastern European Russia. J. Paleolimnol. 46, 123–137 (2011).
    Article  Google Scholar 

    46.
    Kittel, T. G. F., Baker, B. B., Higgins, J. V. & Haney, J. C. Climate vulnerability of ecosystems and landscapes on Alaska’s North Slope. Reg. Environ. Change 11, S249–S264 (2011).
    Article  Google Scholar 

    47.
    McKane, R. B. et al. Reconstruction and analysis of historical changes in carbon storage in arctic tundra. Ecology 78, 1188–1198 (1997).
    Article  Google Scholar 

    48.
    Leavitt, P. R. et al. Paleolimnological evidence of the effects on lakes of energy and mass transfer from climate and humans. Limnol. Oceanogr. 54, 2330–2348 (2009).
    CAS  Article  Google Scholar 

    49.
    Tye, A. M. & Heaton, T. H. E. Chemical and isotopic characteristics of weathering and nitrogen release in non-glacial drainage waters on Arctic tundra. Geochim. Cosmochim. Acta 71, 4188–4205 (2007).
    CAS  Article  Google Scholar 

    50.
    Hobbie, J. E. et al. Impact of global change on biogeochemistry and ecosystems of an arctic freshwater system. Polar Res. 18, 207–214 (1999).
    Article  Google Scholar 

    51.
    Whalen, S. C. & Cornwell, J. C. Nitrogen, phosphorus, and organic-carbon cycling in an Arctic lake. Can. J. Fish. Aquat. Sci. 42, 797–808 (1985).
    CAS  Article  Google Scholar 

    52.
    Mantua, N. J., Hare, S. R., Zhang, Y. & Wallace, J. M. Pacific interdecadal climate oscillation with impacts on salmon production. Bull. Am. Meteorol. Soc. 78, 1069–1079 (1997).
    Article  Google Scholar 

    53.
    Galloway, J. N. et al. Transformation of the nitrogen cycle: Recent trends, questions, and potential solutions. Science 320, 889–892 (2008).
    CAS  Article  Google Scholar 

    54.
    Dean, W. E. Determination of carbonate and organic-matter in calcareous sediments and sedimentary-rocks by loss on ignition-comparison with other methods. J. Sediment. Petrol. 44, 242–248 (1974).
    CAS  Google Scholar 

    55.
    Appleby, P. G. In Tracking Environmental Change Using Lake Sediments Volume 1: Basin Analysis, Coring, and Chronological Techniques (eds. Last, W. M. & Smol, J. P.) 171–203 (Kluwer, 2001).

    56.
    Engstrom, D. R. & Rose, N. L. A whole-basin, mass-balance approach to paleolimnology. J. Paleolimnol. 49, 333–347 (2013).
    Article  Google Scholar 

    57.
    Dietz, R. D., Engstrom, D. R. & Anderson, N. J. Patterns and drivers of change in organic carbon burial across a diverse landscape: insights from 116 Minnesota lakes. Global Biogeochem. Cycles 29, 708–727 (2015).
    CAS  Article  Google Scholar 

    58.
    DeMaster, D. J. The supply and accumulation of silica in the marine environment. Geochim. Cosmochim. Acta 45, 1715–1732 (1981).
    CAS  Article  Google Scholar 

    59.
    Conley, D. J. & Schelske, C. L. In Tracking Environmental Change Using lake Sediments: Biological Methods and Indicators (eds. Smol, J. P., Birks, H. J. B. & Last, W. M.) 281–293 (Kluwer, 2001).

    60.
    Renberg, I. A procedure for preparing large sets of diatom slides from sediment cores. J. Paleolimnol. 4, 87–90 (1990).
    Article  Google Scholar 

    61.
    Savage, C., Leavitt, P. R. & Elmgren, R. Distribution and retention of effluent nitrogen in surface sediments of a coastal bay. Limnol. Oceanogr. 49, 1503–1511 (2004).
    CAS  Article  Google Scholar 

    62.
    Wood, S. N. Generalized Additive Models: An Introduction with R, 2nd edn. (Chapman and Hall/CRC, 2017).

    63.
    Wood, S. N. Thin-plate regression splines. J. R. Stat. Soc. (B) 65, 95–114 (2003).
    Article  Google Scholar 

    64.
    Simpson, G. L. Modelling palaeoecological time series using generalized additive models. Front. Ecol. Evol. 6, 149 (2018).
    Article  Google Scholar 

    65.
    Bai, J. & Perron, P. Computation and analysis of multiple structural change models. J Appl. Econ. 181, 1–22 (2003).
    Article  Google Scholar 

    66.
    R Core Team. A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, http://www.R-project.org/ (2013).

    67.
    Zeileis, A., Leisch, F., Hornik, K. & Kleiber, C. strucchange: An R package for testing for structural change in linear regression models. J. Stat. Softw. 7, 1–38 (2002).
    Article  Google Scholar 

    68.
    Anderson, N. J., Engstrom, D. R., Leavitt, P. R., Flood, S. M. & Heathcote, A. J. Mendeley, https://doi.org/10.17632/xfmpvjmrby.1 (2020).

    69.
    The Alaska Climate Research Center. Climatological Data-Alaska’s Arctic, http://oldclimate.gi.alaska.edu/Climate/Location/TimeSeries/Data/brwT (2010).

    70.
    Hastings, M. G., Steig, E. J. & Sigman, D. M. Seasonal variations in N and O isotopes of nitrate in snow at Summit, Greenland: Implications for the study of nitrate in snow and ice cores. J. Geophys. Res. Atmos. 109, https://doi.org/10.1029/2004jd004991 (2004).

    71.
    Tseng, C.-M., Lamborg, C. H., Fitzgerald, W. F. & Engstrom, D. R. Cycling of dissolved elemental mercury in Arctic Alaskan lakes. Geochim. Cosmochim. Acta 68, 1173–1184 (2004).
    CAS  Article  Google Scholar  More

  • in

    Analysis of molecular diversity within single cyanobacterial colonies from environmental samples

    Genotypic heterogeneity in single Rivularia-like colonies
    Rivularia-like colonies have a global distribution, occurring in marine or freshwater habitats, where they are usually attached to a rocky substrate; however, many studies have reported that Rivularia spp. are associated with unpolluted environments14. In addition, the relationships between some morphological or physiological features and the environment make these species excellent environmental indicators of changes in running water quality, mainly related to eutrophication processes14,30. Therefore, they have been included in biomonitoring programs21,31,32. On the other hand, because Rivularia colonies sometimes persist for very long periods, avoiding grazing, the toxicity of these colonies is being investigated33. It is undoubted that in all of these approaches, where genera and species must be strictly identified from environmental samples, accurate cyanobacterial characterization is essential.
    Traditional identification of cyanobacteria involves assigning a colony to a morphospecies, and conventionally, a bacterial colony is defined as a visible mass of clonal microorganisms, all of which originated from a single cell. However, the results from the present study show that the majority of the analyzed colonies consist of different clones growing together. Among the 28 Rivularia-like colonies, the phylotype corresponding to Rivularia sp. was present in 19 colonies, with abundances ranging from 59.4 to 99.8% depending on the studied colony. Nevertheless, it should also be noted that in most of the colonies, this phylotype dominated, whereby in 14 colonies, it presented an abundance of ≥ 90% (and within 7 of these colonies, the abundance was close to 99%). However, in three colonies, the abundance ranged from 72 to 85%, and in two of them, the abundance decreased to approximately 60%. The other highly abundant phylotypes found in these colonies, which reached abundances up to approximately 21%, corresponded to Calothrix sp. and Oculatella sp., the latter a genus morphologically similar to Leptolyngbya but separated from it because of genetic differences34. These results indicated great variability in the abundance of the phylotype corresponding to Rivularia depending on the analyzed colony, as well as variation in the other phylotypes and their abundances found in these colonies.
    One of the surprising findings was that among the twenty-eight analyzed Rivularia-like colonies, seven corresponded to the new, recently described genus Cyanomargarita, which as the authors described, is virtually indistinguishable from Rivularia in field samples15. In these colonies, genotypic heterogeneity was also found, in which the abundance of the phylotype corresponding to Cyanomargarita varied from 57,28% in a colony with clear lamination resembling R. haematites (see Fig. 3b) to 99.2% in a soft colony resembling R. biasolettiana. Interestingly, in these colonies, Phormidium sp. was the dominant nonheterocystous cyanobacterium instead of Oculatella from Rivularia colonies, but the phylotype corresponding to Calothrix was also found.
    Furthermore, phylotypes corresponding to Cyanomargarita and Rivularia were never found together in the same colony, although both types of colonies coexisted in the same rivers (e.g., Gordale Beck and Endrinales). Allelopathic effects could explain these results, as previously suggested for other cyanobacteria35. In fact, García-Espín et al.33 showed that extracts obtained from Rivularia colonies affected the photosynthetic activity of several diatoms and a red alga. Further experiments with extracts from both colonies would confirm this possible effect.
    Another very surprising finding was that two Rivularia-like colonies did not present any phylotypes corresponding to Rivularia or Cyanomargarita (or contained them at an abundance ≤ 0.7%). In one of these colonies (colony BAT4), five different phylotypes were found at similar abundances (approximately 15–20%), of which three corresponded to different Calothix spp. and the others corresponded to other Nostocaceae and Leptolyngbyaceae. In the other colony (BAT13), the dominant phylotype corresponded to the new genus Macrochaete16. This genus has been described only from cultures, so to the best of our knowledge, this is the first report in which a natural population is morphologically and genetically characterized. Nevertheless, it is noteworthy that the morphological characteristics of filaments and trichomes in this environmental sample were different from those reported in the description of this new genus, in which the phenotypic features resembled those of Calothrix. However, these features corresponded only to isolated strains, which are known to exhibit morphological variability and differences from natural populations7,12,13.
    R. biasolettiana vs R. haematites
    However, what was very interesting and deserves to be highlighted is that when we tried to differentiate the two typical Rivularia colonies found in calcareous streams, R. biasolettiana and R. haematites, we did not find genetic differences, at least at the studied level, the 16S rRNA gene.
    16S rRNA is the most widely used marker gene36,37, which fits the criteria of ubiquity, regions of strong conservation, and regions of hypervariability38,39. This gene is supported by reference databases containing over a million full-length 16S rRNA sequences, therefore spanning a broad phylogenetic spectrum40. The 16S rRNA gene has served as the general framework and as the benchmark for the taxonomy of prokaryotes41. Advances in high-throughput sequencing technologies have enabled almost comprehensive descriptions of bacterial diversity through 16S rRNA gene amplicons, which have been used in surveys of microbial communities to characterize the composition of microorganisms present in environments worldwide42,43,44,45. Although some issues have been raised, such as identification of metabolic or other functional capabilities of microorganisms when studies focus only on this gene, recent studies have shown that the phylogenetic information contained in 16S marker gene sequences is sufficiently well correlated with genomic content to yield accurate predictions when related reference genomes are available46,47,48,49. Therefore, the 16S rRNA gene continues to be the mainstay of sequence-based bacterial analysis, vastly expanding our understanding of the microbial world50.
    In particular, in cyanobacteria, as in other prokaryotes, the 16S rDNA gene is currently the most commonly used marker for molecular and phylogenetic studies51,52. The information obtained from 16S rDNA gene phylogenetic reconstructions, together with morphological, ultrastructural, and ecological data, led Komárek et al.53 to propose the current accepted classification of cyanobacteria. There have also been specific studies by this group concerning the problems associated with single-gene phylogenies, in which robust phylogenomic trees of cyanobacteria derived from multiple conserved proteins have also shown congruence between the multilocus and 16S rRNA gene phylogenies, which once again demonstrates the considerable strength of the 16S rRNA gene for phylogenetic inference and evaluation of prokaryote diversity54,55,56,57.
    In this study, in contrast to the genetic identity found in R. biasolettiana and R. haematites colonies, showing a dominance of OTU1, the remainder of the studied representatives of Rivulariaceae showed a wide range of variation in the 16S rDNA sequences and with OTU1. Sequence identity between OTU1 and the remaining OTUs belonging to this family was as low as approximately 90%, ranging from 90.73 to 93.41%, and when it was compared with other Rivulariaceae from the databases, in the different clusters of the phylogenetic tree, this value ranged from 87.12 to 93.90%. A large difference between the sequences of this gene was also found in other studies on Rivulariaceae15,16,17,29,58. In fact, several new genera are emerging on the basis of these differences15,16,17. Comparisons of phylogenies using other markers, such as the phycocyanin operon and the intervening intergenic spacer (cpcBA-IGS) with the 16S rRNA gene in previous studies in Rivulariaceae, have shown largely consistent results, with a high level of divergence between the components of this family11.
    In addition, the results of the present study showed correlations between morphological characteristics and the analyzed genes in all the cyanobacterial colonies/tufts, except for those of R. biasolettiana and R. haematites. In these two cyanobacteria, only distinct macroscopic phenotypic features were observed due to zonation and different degrees of calcification since no significant differences were found in the size measurements or other microscopic characteristics.
    Therefore, although the remainder of the genome has not been studied in these populations, the genetic identity of the studied marker, phenotypic features, together with environmental preferences point out that R. biasolettiana and R. haematites are ecotypes of the same species, as previously suggested59.
    R. biasolettiana and R. haematites have very similar morphotypes, and traditional taxonomical classification and studies have distinguished them primarily by their degrees of calcification. R. biasolettiana-type colonies are described as more gelatinous and less calcified, and the crystals are disseminated; however, R. haematites colonies are very hard and exhibit extensive calcification in concentric zones, which leads to clear lamination24,25,60,61. Because of its heavy mineralization, R. haematites is a model for stromatolite-binding organisms25,26.
    Microscopic observations from this study showed that some colonies presented typical R. haematites morphology with concentric bands of intense calcification (see Fig. 2a,b), and others were soft and less calcified, such as R. biasolettiana, although all of them presented the same dominant phylotype. Many others with this dominant phylotype have also shown ambiguous morphology with no clear lamination, although some dark/light zones could be observed (see, e.g., Fig. 4b,d, f). Even in Cyanomargarita colonies, whose genotype was clearly separated from that of Rivularia, concentric zones and extensive calcification could be observed (see, e.g., Fig. 3b,d). These results suggested that these phenotypic features are not diagnostic characteristics for further identification.
    In a two-year study, Obenlüneschloss and Schneider61 found that not all analyzed R. haematites colonies showed distinct concentric calcification layers. In the stromatolites of both types of Rivularia, the same lamination was observed, and the differences in calcification appeared later60. Pentecost and Franke26 compared populations of R. biasolettiana and R. haematites and argued that although both could be distinguished by their form of calcification and their trichome diameter, some populations of R. biasolettiana were more intensely calcified than others, suggesting that a continuity of forms may exist, even within the same stream, and therefore, a continuum of colony forms probably occurs between these taxa.
    Differences in the calcification pattern have been attributed to seasonality and cyanobacterial activity, in particular to photosynthesis24,26,62. The calcification in R. haematites occurred in concentric bands, which varied in thickness and the density of crystals. Since characteristic zonation is formed by filaments of different successive generations, the thickness will vary depending on the growth rate, while crystal density will depend on the rate of calcification. Calcification is the result of photosynthesis (with a maximum of 14%) and evaporation during the warmer seasons, while it is entirely abiogenic during winter as a result of CO2 evasion63. Therefore, dense calcified bands similar to those formed in winter have been described that are caused by a reduction in trichome growth and EPS production, allowing the development of abiotic surface precipitate, and less calcified layers are formed during spring and summer, when calcification is associated with photosynthesis in zones of growth with cell division24,26. Thus, differences in climatic conditions and/or biological activity seem to lead to differences in the degrees of zonation and calcification.
    The growth of Rivularia colonies is seasonal and strongly correlated with water temperature24,26. The colony growth rates were 12–14 µm/day in summer and 2 µm/day in winter24. The occurrence of R. biasolettiana was more closely related to high temperatures than that of R. haematites21. Moreover, colonies of R. haematites were generally collected under temperatures below 15 °C in mountain running waters64, and R. haematites stromatolites have been described as preferentially developed in wet periods, particularly in autumn and winter60. Our own field observations during the sampling for this and previous studies were that the gelatinous and weakly calcified R. biasolettiana type was more abundant in warmer locations, and in contrast, R. haematites was dominant in cold locations (data not shown).
    One possible explanation for the results found in this study could be related to these differences in the degree of zonation and calcification in relation to climate, which could include microclimatic conditions. In warmer sites or climatic conditions, when growth is rapid, the number of filaments will increase, moving towards the surface in a weakly dense and unaligned arrangement, on which calcite crystals spread, providing a lighter and less calcified structure. Thus, increased growth of Rivularia colonies can lead to the R. biasolettiana type. Under colder conditions, such as in winter, or microclimatic conditions, when growth slows down for other reasons, such as low light, filaments become more densely packed, allowing the development of extensive precipitates and leading to a dark band. When these conditions change, e.g., in the spring and summer, increases in temperatures and/or light will result in increasing and faster growth, leading to a less calcified new layer, and successive seasonal and/or microenvironmental changes will result in the typical lamination of R. haematites. Therefore, warmer places with high temperatures and/or light will allow the occurrence of the R. biasolettiana type, while in colder sites and/or sites with alternating environmental conditions, the R. haematites type will develop. Shaded colonies and colonies that lie in the supratidal spraywater zone often contain small, irregular and more densely packed crystals61.
    Cyanobacteria are known to modify EPS production, pigments, and morphology under environmental stimuli6. The production of EPS also varies depending on the cyanobacteria, whereby Rivularia has shown a well-developed exopolymer layer65, which is of great importance for this epilithic cyanobacteria, as it acts as an adhesive that allows cells to stick to the stones in the running waters, and it holds the filaments together, minimizing cell damage during intermittent drying exposure to the air and evaporation in the warmer seasons66. The C/N ratio is an important parameter for the variation in EPS production since high amounts of fixed C compared to N levels drive EPS synthesis to store excess C67,68. Therefore, Rivularia colonies that are exposed, in spring and summer, to high light intensities and temperatures will increase their photosynthetic rates and therefore the amount of EPS, as shown by the R. biasolettiana morphotype. In addition, most of the analyzed populations were dark in color, probably in relation to the accumulation of the yellow–brown scytonemin pigment in the sheaths or EPS, as previously observed in shallow and clear oligotrophic ecosystems, where water clarity allows UV radiation to penetrate well, protecting the cells from the damaging effects of this radiation69,70.
    In conclusion, environmental factors can lead to differences in macroscopic phenotypic features, such as those found in the Rivularia colonies studied here. However, further sampling under different climatic conditions and/or microenvironmental conditions or of Rivularia cultures grown under distinct temperature and/or illumination conditions, as well as analysis of other genes, are needed to confirm this hypothesis. More

  • in

    Discerning the thermodynamic feasibility of the spontaneous coexistence of multiple functional vegetation groups

    Experimental design
    A multi-layer canopy-root-soil model (MLCan)24,26,27 is used to calculate the energy and entropy fluxes for three climatologically-different ecosystems containing multiple functional groups: water-limited Santa Rita Mesquite (SRM), energy-limited Willow Creek (WCR), and nutrient-limited Tapajos National Forest (TAP)38.
    MLCan takes site-specific parameters and weather forcing data and computes the energy and entropy fluxes and temperatures for each of the ecosystem layers. Entropy calculations are based on both the energy fluxes and temperature of soil, air, and leaves (see Entropy Calculations). The model is run for a simulation period of 2 years (2004–2005) at a half-hourly timescale for SRM and WCR and an hourly timescale for TAP due to data availability. Weather forcing data was downloaded from FLUXNET2015: air temperature, air pressure, global radiation, precipitation, wind speed, friction velocity, and relative humidity32,33,34. Additional model input parameters can be found in Table S2 of the Supplementary Information.
    The initial soil moisture and temperature profiles for each of the sites—and snow properties for WCR—were produced from a spin-up of the model. The WCR and TAP sites used 2004 LAI with 2003 forcing data for a spin-up of 2 years to provide the initial conditions for the beginning of the 2004 simulation. For the SRM site, the FLUXNET2015 data was not available for 2003, so 2004 data was used instead.
    At each site, the model splits up the vegetation into plant functional groups. Domingues et al.35 demonstrates the importance of modeling ecosystems based on functional groups. For WCR and SRM, the vegetation is represented by understory herbaceous species and overstory trees. For TAP, a high biodiversity ecosystem in Amazonia, the vegetation is further divided and represented by four groups: understory tree, mid-canopy tree, upper-canopy tree, and upper-canopy liana35. See Table S1 of the Supplementary Information for functional group abbreviations.
    The LAI data for all sites are taken from MODIS39 and calibrated based on site documentation (Fig. S4 of the Supplementary Information). The LAI is then partitioned into two or four components based on the number of functional groups at each site. Additional LAI information can be found in the Supplementary Information.
    MLCan has been previously validated for each of the sites considered30,40. Since entropy cannot be directly measured, we provide a comparison of the model outputted latent heat fluxes with the observed fluxes at each site in Fig. S5 of the Supplementary Information for additional validation.
    Site descriptions
    The SRM site is located on the Santa Rita Experimental Range in southern Arizona ((31.8214^{circ }hbox{N}), (110.8661^{circ }hbox{W})). SRM has a hot semi-arid climate and consists of woody savannas with mesquite trees (Prosopis velutina Woot.) and C4 grasses and subshrubs40,41.
    The WCR (Willow Creek) site is located within the Chequamegon-Nicolet National Forest in northern Wisconsin ((45.8059^{circ }hbox{N}), (90.0799^{circ }hbox{W})) with a northern continental climate. It is a deciduous broadleaf forest dominated by sugar maple (Acer saccharum Marsh.) with understory shrubs, including bracken ferns (Pteridium aquilinum), and overstory seedlings and saplings42,43,44.
    The TAP (Tapajos National Forest) site data is taken from the Santarem Km 67 Primary Forest site located in Belterra, Pará, Brazil ((2.8567^{circ }hbox{S}), (54.9589^{circ }hbox{W})). This evergreen broadleaf forest in Amazonian Brazil has a tropical monsoon climate with vegetation consisting of dozens of known tree species and lianas30,35.
    Entropy calculations
    Entropy calculations are based on model-simulated temperature and energy at each of the 20 canopy layers and the soil-surface layer, and results are scaled up to the ecosystem level. No lateral exchange of fluxes are considered. The net sum of energy fluxes from all layers of the ecosystem is equivalent to the total flux of energy across the boundary of the control volume (Fig. S1 of the Supplementary Information). These energy fluxes include shortwave radiation (SW), longwave radiation (LW), latent heat (LE), and sensible heat (H). All results are categorized as the flux of energy at the boundary entering ((SW_{in}), (LW_{in})) or leaving ((SW_{out}), (LW_{out}), LE, H) the ecosystem. Because the total energy flux across the ecosystem boundary is equal to the sum across the canopy layers in the model, the total entropy flux across the boundary can also be taken as the cumulative sum of the entropy fluxes from all layers of the ecosystem.
    Entropy flux calculations are summarized in Table 1. All energy variables have units of (hbox{W/m}^2), entropy variables are in (hbox{W/m}^2hbox{K}), and temperatures are in K.
    Table 1 Entropy calculations
    Full size table

    Entropy for LE and H calculations are based on simple heat transfer. The change in entropy is:

    $$begin{aligned} dS=frac{dQ}{T} end{aligned}$$
    (1)

    where dQ is change in heat and T is temperature49. Thus, the flux of entropy for a given energy flux (E) across a boundary is:

    $$begin{aligned} J=frac{E}{T} . end{aligned}$$
    (2)

    However, thermal radiation (SW and LW) cannot be treated this simply. The entropy flux for blackbody radiation is:

    $$begin{aligned} J_{BR} = frac{4}{3} sigma T^3 = frac{4}{3} frac{E_{BR}}{T} end{aligned}$$
    (3)

    where (sigma) is the Stefan–Boltzmann constant, and (E_{BR}) is the blackbody radiation flux defined as (sigma T^4) from the Stefan–Boltzmann Law48,49.
    SW is considered blackbody radiation, and entropy fluxes for direct shortwave radiation ((J_{SW,direct})) can be obtained by Eq. 3. However, LW is considered non-blackbody radiation, also called ‘diluted blackbody radiation’, which must include an additional factor (X(epsilon )) to account for the entropy produced during the ‘diluted emission’ of radiation given by an object’s emissivity, (epsilon). This factor is defined as45,46:

    $$begin{aligned} X(epsilon ) = 1-Big [frac{45}{4pi ^4}ln {(epsilon )}(2.336-0.26epsilon )Big ]. end{aligned}$$
    (4)

    Although (SW_{diffuse}) is still a blackbody radiation, it has been demonstrated47 that the entropy flux due to (SW_{diffuse}) can be treated similarly to non-blackbody radiation with a new variable, (xi), in place of emissivity. (xi) is the ‘dilution factor’ of radiation due to scattering, meaning it is the ratio of diffuse solar radiance on Earth’s surface to solar radiance in extraterrestrial space47. Since diluted blackbody radiation ((SW_{diffuse})) is mathematically equivalent to non-blackbody radiation (LW) when the dilution factor is equal to the emissivity, (xi) can also be plugged into Eq. 4 to solve for the amplifying factor of entropy production due to scattering, (X(xi ))37,45,46,48.
    Each of the entropy calculations in Table 1 have a temperature value corresponding to the temperature of the energy’s source. For instance, shortwave radiation originates from the sun, so the source temperature in its entropy equations is (T_{sun}). Likewise, longwave radiation is assumed to originate from the atmosphere, leading to a corresponding temperature of (T_{atm}). However, (LW_{out}), LE, and H do not have a single source location, so we must calculate an equivalent temperature ((T_{eq})) for each energy category based on the modeled temperatures and weighted contribution of each layer to the total energy flux at the ecosystem boundary. The equivalent temperatures for these three energy categories are calculated as follows:

    $$begin{aligned} T_{eq,j} = sum _{k=1}^{21}[T_{k} times omega _{j, k}] end{aligned}$$
    (5)

    where (T_{eq,j}) is the equivalent temperature of energy category j such that (j in {LW_{out}, LE, H}). k refers to the layer in the ecosystem such that layers 1-20 are the canopy layers, and layer 21 refers to the ground surface. (T_k) is the temperature of layer k, and (omega _{j,k}) is the weight of energy category j coming from layer k given by:

    $$begin{aligned} omega _{j,k} = frac{E_{j,k}}{E_{j,eco}} end{aligned}$$
    (6)

    where (E_{j,k}) is the energy j leaving layer k, and (E_{j,eco}) is the total energy j leaving the ecosystem.
    The total entropy flux of the ecosystem ((J_{eco})) is calculated by summing the energy categories:

    $$begin{aligned} J_{eco} = sum J_j + J_{SWout} end{aligned}$$
    (7)

    where (J_{SWout}) is the entropy flux of diffuse shortwave radiation leaving the ecosystem. The entropy flux per unit energy (EUE) is another way to view the thermodynamic state of ecosystem vegetation. EUE is calculated as:

    $$begin{aligned} EUE_{j} = frac{J_{j}}{E_{j}} end{aligned}$$
    (8)

    where (EUE_{j}) is the entropy per unit energy in 1/K of energy category j. It follows that the corresponding (EUE_{SWout} = J_{SWout}/E_{SWout}), and the total ecosystem EUE is:

    $$begin{aligned} EUE_{eco} = frac{sum J_j + J_{SWout}}{sum E_j + E_{SWout}}. end{aligned}$$
    (9)

    Work calculations
    Work in an ecosystem represents the energy required to directly perform motion in the form of heat, effectively decreasing the temperature gradient within the ecosystem. We assume that LE and H are the primary regulators of temperature within a natural ecosystem, and (LW_{out}) is wasted energy. Additionally, we assume that the bottom of the control volume is sufficiently deep such that the temperature at the boundary is consistent and there is no loss of heat (i.e. ground heat flux is ignored). Thus, work is estimated and calculated directly from LE, H, and change in internal energy due to photosynthesis, (Delta Q):

    $$begin{aligned} W = LE+H+Delta Q end{aligned}$$
    (10)

    where (Delta Q) is significantly less than LE and H and can be ignored. So work can be simplified to:

    $$begin{aligned} W = LE+H. end{aligned}$$
    (11)

    Since work represents the ability of an ecosystem’s vegetation to deplete the driving temperature gradient imposed upon the ecosystem, our analysis compares work with temperature gradient. We define temperature gradient as:

    $$begin{aligned} frac{Delta T}{Delta z} = frac{T_{surf}-T_{air}}{h_e} end{aligned}$$
    (12)

    where (T_{surf}) is the temperature of the soil surface, (T_{air}) is the temperature of the air in the top layer of the ecosystem, and (h_e) is the ecosystem height (see Table S2 in the Supplementary Information).
    Work efficiency is the work performed for the amount of radiation entering the ecosystem defined as:

    $$begin{aligned} WE = frac{LE+H}{E_{SWin} + E_{LWin}} = frac{W}{E_{in}}. end{aligned}$$
    (13)

    Since each vegetation functional group partitions energy differently among the energy categories, work efficiency is a good way to compare thermodynamic behavior across model scenarios at each site in a normalized way.
    Statistical analysis
    To determine if the differences of entropy flux and work efficiency among scenarios at each site are statistically significant, we perform two separate tests for entropy flux and work efficiency. Since entropy flux distributions are positively skewed (Fig. 1a), we use the variance as an indicator of the difference between them. To this end we use the distribution-free Miller Jackknife (MJ) significance test50,51 for variance that does not assume that the distributions come from populations with the same median. However, the work efficiency distributions exhibit no such pattern (Fig. 1b), and, therefore, we use the two-sample Kolmogorov–Smirnov (KS) test, which measures the maximum absolute difference between two empirical cumulative distribution functions (CDF)52,53,54.
    First, the entropy flux variances are compared with the MJ test. Because functional group scenarios at each site are bounded on the lower end by similar values, if a distribution has a larger variance than another, then the two populations cannot be considered as coming from the same continuous distribution, and the distribution with a larger variance generally consists of larger values. For each site we test the null hypothesis, (H_0), that the distribution of multiple-functional-group entropy fluxes and the distribution for each of its single-functional-groups have the same variance. This is done with each functional group present at each site (Table S1). The alternate hypothesis, (H_{A1}), states that the distribution of entropy fluxes from the multiple-functional-group has a larger variance than that of the corresponding single-functional-group, meaning that the two populations do not belong to the same distribution and the multi-group scenario consists of larger values than the single-group scenario. The results from this test, shown in Table 2, indicate that (H_0) is rejected in favor of (H_{A1}) at the 5% level ((p More

  • in

    Satellite megaclusters could fox night-time migrations

    The brightness to the naked eye of giant megaconstellations of satellites could create the greatest alteration to the night sky’s appearance in human history (see Nature https://doi.org/fdz8; 2020). This could have potentially catastrophic effects on celestial navigation by wildlife, and therefore on terrestrial ecology.
    Migrating species such as birds, dung beetles and seals use stars as a source of directional information (see J. J. Foster et al. Proc. R. Soc B 285, 20172322; 2018). Some use bright objects as their main cue, and others rely on the starry sky’s centre of rotation, or the fainter band of the Milky Way.
    Constellations of satellites can form coherent patterns that could affect night-time migrations in a similar way to the Milky Way, for example. Such errors would have a global effect on migratory populations. Their energy balance could be altered, with long-term repercussions for survival and reproduction, as has been found for excessive or misdirected light (S. A. Cabrera-Cruz et al. Sci. Rep. 8, 3261; 2018).
    We call for astronomers and field biologists to identify and quantify the possible ecological effects of satellite megaconstellations. A regulatory framework could then be developed to control their proliferation, based on how species respond to spatio-temporal cues in the field. More

  • in

    Local community assembly mechanisms shape soil bacterial β diversity patterns along a latitudinal gradient

    1.
    Nemergut, D. R. et al. Patterns and processes of microbial community assembly. Microbiol Mol. Biol. Rev. 77, 342–356 (2013).
    PubMed  PubMed Central  Article  Google Scholar 
    2.
    Fierer, N. & Jackson, R. B. The diversity and biogeography of soil bacterial communities. Proc. Natl Acad. Sci. USA 103, 626–631 (2006).
    ADS  CAS  PubMed  Article  Google Scholar 

    3.
    Bahram, M. et al. Structure and function of the global topsoil microbiome. Nature 560, 233 (2018).
    ADS  CAS  PubMed  Article  Google Scholar 

    4.
    Martiny, J. B. H. et al. Microbial biogeography: putting microorganisms on the map. Nat. Rev. Microbiol. 4, 102 (2006).
    CAS  PubMed  Article  Google Scholar 

    5.
    Fierer, N. Embracing the unknown: disentangling the complexities of the soil microbiome. Nat. Rev. Microbiol. 15, 579 (2017).
    CAS  PubMed  Article  Google Scholar 

    6.
    Sul, W. J., Oliver, T. A., Ducklow, H. W., Amaral-Zettler, L. A. & Sogin, M. L. Marine bacteria exhibit a bipolar distribution. Proc. Natl Acad. Sci. USA 110, 2342–2347 (2013).
    ADS  CAS  PubMed  Article  Google Scholar 

    7.
    Delgado‐Baquerizo, M. et al. Carbon content and climate variability drive global soil bacterial diversity patterns. Ecol. Monogr. 86, 373–390 (2016).
    Article  Google Scholar 

    8.
    Martiny, J. B., Eisen, J. A., Penn, K., Allison, S. D. & Horner-Devine, M. C. Drivers of bacterial β-diversity depend on spatial scale. Proc. Natl Acad. Sci. USA 108, 7850–7854 (2011).
    ADS  CAS  PubMed  Article  Google Scholar 

    9.
    Cavender‐Bares, J., Kozak, K. H., Fine, P. V. & Kembel, S. W. The merging of community ecology and phylogenetic biology. Ecol. Lett. 12, 693–715 (2009).
    PubMed  Article  Google Scholar 

    10.
    Kraft, N. J. et al. Disentangling the drivers of β diversity along latitudinal and elevational gradients. Science 333, 1755–1758 (2011).
    ADS  CAS  PubMed  Article  Google Scholar 

    11.
    Myers, J. A., Chase, J. M., Crandall, R. M. & Jiménez, I. Disturbance alters beta‐diversity but not the relative importance of community assembly mechanisms. J. Ecol. 103, 1291–1299 (2015).
    Article  Google Scholar 

    12.
    Stegen, J. C. et al. Quantifying community assembly processes and identifying features that impose them. ISME J. 7, 2069–2079 (2013).
    PubMed  PubMed Central  Article  Google Scholar 

    13.
    Wang, X. B. et al. Habitat-specific patterns and drivers of bacterial β-diversity in China’s drylands. ISME J. 11, 1345 (2017).
    PubMed  PubMed Central  Article  Google Scholar 

    14.
    Ferrenberg, S. et al. Changes in assembly processes in soil bacterial communities following a wildfire disturbance. ISME J. 7, 1102–1111 (2013).
    PubMed  PubMed Central  Article  Google Scholar 

    15.
    Myers, J. A. et al. Beta-diversity in temperate and tropical forests reflects dissimilar mechanisms of community assembly. Ecol. Lett. 16, 151–157 (2013).
    PubMed  Article  Google Scholar 

    16.
    Catano, C. P., Dickson, T. L. & Myers, J. A. Dispersal and neutral sampling mediate contingent effects of disturbance on plant beta-diversity: a meta-analysis. Ecol. Lett. 20, 347–356 (2017).
    PubMed  Article  Google Scholar 

    17.
    Dini-Andreote, F., Stegen, J. C., Elsas, J. D. & Salles, J. F. Disentangling mechanisms that mediate the balance between stochastic and deterministic processes in microbial succession. Proc. Natl Acad. Sci. USA 112, 1326–1332 (2015).
    Article  CAS  Google Scholar 

    18.
    Tripathi, B. M. et al. Soil pH mediates the balance between stochastic and deterministic assembly of bacteria. ISME J. 12, 1072 (2018).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    19.
    Albright, M. B. N. & Martiny, J. B. H. Dispersal alters bacterial diversity and composition in a natural community. ISME J. 12, 296 (2018).
    PubMed  Article  Google Scholar 

    20.
    Cadotte, M. W. Dispersal and species diversity: a meta-analysis. Am. Nat. 167, 913–924 (2006).
    PubMed  Article  Google Scholar 

    21.
    Questad, E. J. & Foster, B. L. Coexistence through spatio‐temporal heterogeneity and species sorting in grassland plant communities. Ecol. Lett. 11, 717–726 (2008).
    PubMed  Article  Google Scholar 

    22.
    Segre, H. et al. Competitive exclusion, beta diversity, and deterministic vs. stochastic drivers of community assembly. Ecol. Lett. 17, 1400–1408 (2014).
    PubMed  Article  Google Scholar 

    23.
    Ranjard, L. et al. Turnover of soil bacterial diversity driven by wide-scale environmental heterogeneity. Nat. Commun. 4, 1434 (2013).
    ADS  CAS  PubMed  Article  Google Scholar 

    24.
    Qian, W. & Lin, X. Regional trends in recent temperature indices in China. Clim. Res. 27, 119–134 (2004).
    Article  Google Scholar 

    25.
    Ma, B. et al. Geographic patterns of co-occurrence network topological features for soil microbiota at continental scale in eastern China. ISME J. 10, 1891 (2016).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    26.
    Lozupone, C. A. & Knight, R. Global patterns in bacterial diversity. Proc. Natl Acad. Sci. USA 104, 11436–11440 (2007).
    ADS  CAS  PubMed  Article  Google Scholar 

    27.
    Morton, J. T. et al. Uncovering the Horseshoe effect in microbial analyses. mSystems 2, e00166–00116 (2017).
    PubMed  PubMed Central  Google Scholar 

    28.
    Zeng, J. et al. Nitrogen fertilization directly affects soil bacterial diversity and indirectly affects bacterial community composition. Soil Biol. Biochem. 92, 41–49 (2016).
    CAS  Article  Google Scholar 

    29.
    Chisholm, R. A. & Pacala, S. W. Theory predicts a rapid transition from niche-structured to neutral biodiversity patterns across a speciation-rate gradient. Theor. Ecol. 4, 195–200 (2011).
    Article  Google Scholar 

    30.
    Fierer, N. et al. Comparative metagenomic, phylogenetic and physiological analyses of soil microbial communities across nitrogen gradients. ISME J. 6, 1007–1017 (2012).
    CAS  PubMed  Article  Google Scholar 

    31.
    Rousk, J. et al. Soil bacterial and fungal communities across a pH gradient in an arable soil. ISME J. 4, 1340 (2010).
    PubMed  Article  Google Scholar 

    32.
    Bell, T. Experimental tests of the bacterial distance decay relationship. ISME J. 4, 1357 (2010).
    PubMed  Article  Google Scholar 

    33.
    Morlon, H. et al. A general framework for the distance–decay of similarity in ecological communities. Ecol. Lett. 11, 904–917 (2008).
    PubMed  PubMed Central  Article  Google Scholar 

    34.
    Dexter, K. G., Terborgh, J. W. & Cunningham, C. W. Historical effects on beta diversity and community assembly in Amazonian trees. Proc. Natl Acad. Sci. USA 109, 7787–7792 (2012).
    ADS  CAS  PubMed  Article  Google Scholar 

    35.
    Chase, J. M. & Myers, J. A. Disentangling the importance of ecological niches from stochastic processes across scales. Philos. Trans. R. Soc. B Biol. Sci. 366, 2351–2363 (2011).
    Article  Google Scholar 

    36.
    Mori, A. S., Fujii, S., Kitagawa, R. & Koide, D. Null model approaches to evaluating the relative role of different assembly processes in shaping ecological communities. Oecologia 178, 261–273 (2015).
    ADS  PubMed  Article  Google Scholar 

    37.
    Zhou, J. et al. Stochasticity, succession, and environmental perturbations in a fluidic ecosystem. Proc. Natl Acad. Sci. USA 111, E836–E845 (2014).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    38.
    Dai, W. et al. Spatial variability of soil nutrients in forest areas: a case study from subtropical China. J. Plant Nutr. Soil Sci. 181, 827–835 (2018).
    CAS  Article  Google Scholar 

    39.
    Gravel, D., Canham, C. D., Beaudet, M. & Messier, C. Reconciling niche and neutrality: the continuum hypothesis. Ecol. Lett. 9, 399–409 (2006).
    PubMed  Article  PubMed Central  Google Scholar 

    40.
    Durant, S. M. Competition refuges and coexistence: an example from Serengeti carnivores. J. Anim. Ecol. 67, 370–386 (1998).
    Article  Google Scholar 

    41.
    Langenheder, S., Berga, M., Östman, Ö. & Székely, A. J. Temporal variation of β-diversity and assembly mechanisms in a bacterial metacommunity. ISME J. 6, 1107 (2012).
    CAS  PubMed  Article  Google Scholar 

    42.
    Wu, S. H. et al. Patterns of plant invasions in China: taxonomic, biogeographic, climatic approaches and anthropogenic effects. Biol. Invasions 12, 2179–2206 (2010).
    Article  Google Scholar 

    43.
    Soon, Y. K. & Abboud, S. A comparison of some methods for soil organic carbon determination. Commun. Soil Sci. Plant Anal. 22, 943–954 (1991).
    CAS  Article  Google Scholar 

    44.
    Wright, A. F. & Bailey, J. S. Organic carbon, total carbon, and total nitrogen determinations in soils of variable calcium carbonate contents using a Leco CN-2000 dry combustion analyzer. Commun. Soil Sci. Plant Anal. 32, 3243–3258 (2001).
    CAS  Article  Google Scholar 

    45.
    Gianello, C. & Bremner, J. M. Comparison of chemical methods of assessing potentially available organic nitrogen in soil. Commun. Soil Sci. Plant Anal. 17, 215–236 (1986).
    CAS  Article  Google Scholar 

    46.
    Tamaki, H. et al. Analysis of 16S rRNA amplicon sequencing options on the Roche/454 next-generation titanium sequencing platform. PloS ONE 6, e25263 (2011).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    47.
    Caporaso, J. G. et al. QIIME allows analysis of high-throughput community sequencing data. Nat. Methods 7, 335 (2010).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    48.
    Zhang, X. et al. Changes of soil prokaryotic communities after clear-cutting in a karst forest: evidences for cutting-based disturbance promoting deterministic processes. FEMS Microbiol. Ecol. 92, fiw026 (2016).
    PubMed  Article  CAS  Google Scholar 

    49.
    Edgar, R. C. et al. UCHIME improves sensitivity and speed of chimera detection. Bioinformatics 27, 2194–2200 (2011).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    50.
    Gilbert, J. A. et al. The seasonal structure of microbial communities in the Western English Channel. Environ. Microbiol. 11, 3132–3139 (2009).
    CAS  PubMed  Article  Google Scholar 

    51.
    McMurdie, P. J. & Holmes, S. Waste not, want not: why rarefying microbiome data is inadmissible. PLoS Comput. Biol. 10, e1003531 (2014).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    52.
    Rideout, J. R. et al. Subsampled open-reference clustering creates consistent, comprehensive OTU definitions and scales to billions of sequences. PeerJ 2, e545 (2014).
    PubMed  PubMed Central  Article  Google Scholar 

    53.
    Edgar, R. C. UPARSE: highly accurate OTU sequences from microbial amplicon reads. Nat. Methods 10, 996 (2013).
    CAS  PubMed  Article  Google Scholar 

    54.
    Wang, Q., Garrity, G. M., Tiedje, J. M. & Cole, J. R. Naive Bayesian classifier for rapid assignment of rRNA sequences into the new bacterial taxonomy. Appl. Environ. Microb. 73, 5261 (2007).
    CAS  Article  Google Scholar 

    55.
    Székely, A. J. & Langenheder, S. The importance of species sorting differs between habitat generalists and specialists in bacterial communities. FEMS Microbiol. Ecol. 87, 102–112 (2014).
    PubMed  Article  CAS  Google Scholar 

    56.
    Vannette, R. L. & Fukami, T. Dispersal enhances beta diversity in nectar microbes. Ecol. Lett. 20, 901–910 (2017).
    PubMed  Article  Google Scholar 

    57.
    Raup, D. M. & Crick, R. E. Measurement of faunal similarity in paleontology. J. Paleontol. 53, 1213–1227 (1979).

    58.
    Chase, J. M., Kraft, N. J., Smith, K. G., Vellend, M. & Inouye, B. D. Using null models to disentangle variation in community dissimilarity from variation in α-diversity. Ecosphere 2, 1–11 (2011).
    Article  Google Scholar 

    59.
    Tello, J. S. et al. Elevational gradients in β-diversity reflect variation in the strength of local community assembly mechanisms across spatial scales. PLoS ONE 10, 0121458 (2015).
    Article  CAS  Google Scholar 

    60.
    Moran, P. A. Notes on continuous stochastic phenomena. Biometrika 37, 17–23 (1950).
    MathSciNet  CAS  PubMed  MATH  Article  Google Scholar 

    61.
    Blanchet, F. G., Legendre, P. & Borcard, D. Forward selection of explanatory variables. Ecology 89, 2623–2632 (2008).
    PubMed  Article  Google Scholar  More