More stories

  • in

    Identification of plastic-associated species in the Mediterranean Sea using DNA metabarcoding with Nanopore MinION

    1.
    Windsor, F. M. et al. A catchment-scale perspective of plastic pollution. Glob. Change Biol. 25, 1207–1221 (2019).
    ADS  Article  Google Scholar 
    2.
    Boucher, J. & Billard, G. The challenges of measuring plastic pollution. Field Actions Sci. Rep. J. Field Actions 19, 68–75 (2019).
    Google Scholar 

    3.
    Jambeck, J. R. et al. Plastic waste inputs from land into the ocean. Science 347, 768–771 (2015).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    4.
    Worm, B., Lotze, H. K., Jubinville, I., Wilcox, C. & Jambeck, J. Plastic as a persistent marine pollutant. Annu. Rev. Environ. Resour. 42, 1–26 (2017).
    Article  Google Scholar 

    5.
    Amaral-Zettler, L. A., Zettler, E. R. & Mincer, T. J. Ecology of the plastisphere. Nat. Rev. Microbiol. 18, 139–151 (2020).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    6.
    Zettler, E. R., Mincer, T. J. & Amaral-Zettler, L. A. Life in the “plastisphere”: microbial communities on plastic marine debris. Environ. Sci. Technol. 47, 7137–7146 (2013).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    7.
    Dussud, C. et al. Evidence of niche partitioning among bacteria living on plastics, organic particles and surrounding seawaters. Environ. Pollut. 236, 807–816 (2018).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    8.
    Taberlet, P., Coissac, E., Pompanon, F., Brochmann, C. & Willerslev, E. Towards next-generation biodiversity assessment using DNA metabarcoding. Mol. Ecol. 21, 2045–2050 (2012).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    9.
    De Tender, C. A. et al. Bacterial community profiling of plastic litter in the Belgian part of the North Sea. Environ. Sci. Technol. 49, 9629–9638 (2015).
    ADS  PubMed  Article  CAS  PubMed Central  Google Scholar 

    10.
    Santos, A., van Aerle, R., Barrientos, L. & Martinez-Urtaza, J. Computational methods for 16S metabarcoding studies using Nanopore sequencing data. Comput. Struct. Biotechnol. J. 18, 296–305 (2020).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    11.
    Jacquin, J. et al. Microbial ecotoxicology of marine plastic debris: a review on colonization and biodegradation by the ‘plastisphere’. Front. Microbiol. 10, 865 (2019).
    PubMed  PubMed Central  Article  Google Scholar 

    12.
    Bleidorn, C. Third generation sequencing: technology and its potential impact on evolutionary biodiversity research. Syst. Biodivers. 14, 1–8 (2016).
    Article  Google Scholar 

    13.
    Krehenwinkel, H. et al. Nanopore sequencing of long ribosomal DNA amplicons enables portable and simple biodiversity assessments with high phylogenetic resolution across broad taxonomic scale. GigaScience 8, giz006 (2019).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    14.
    Pawlowski, J. et al. CBOL protist working group: barcoding eukaryotic richness beyond the animal, plant, and fungal kingdoms. PLoS Biol. 10, e1001419 (2012).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    15.
    Leray, M. & Knowlton, N. Censusing marine eukaryotic diversity in the twenty-first century. Philos. Trans. R. Soc. B Biol. Sci. 371, 20150331 (2016).
    Article  Google Scholar 

    16.
    Piganeau, G., Eyre-Walker, A., Grimsley, N. & Moreau, H. How and why DNA barcodes underestimate the diversity of microbial eukaryotes. PLoS ONE 6, e16342 (2011).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    17.
    Saunders, G. W. & Kucera, H. An evaluation of rbcL, tufA, UPA, LSU and ITS as DNA barcode markers for the marine green macroalgae. Cryptogamie Algologie 31, 487 (2010).
    Google Scholar 

    18.
    Schoch, C. L. et al. Nuclear ribosomal internal transcribed spacer (ITS) region as a universal DNA barcode marker for Fungi. Proc. Natl. Acad. Sci. 109, 6241–6246 (2012).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    19.
    Hebert, P. D., Cywinska, A., Ball, S. L. & Dewaard, J. R. Biological identifications through DNA barcodes. Proc. R. Soc. Lond. Ser. B Biol. Sci. 270, 313–321 (2003).
    CAS  Article  Google Scholar 

    20.
    Bahram, M., Anslan, S., Hildebrand, F., Bork, P. & Tedersoo, L. Newly designed 16S rRNA metabarcoding primers amplify diverse and novel archaeal taxa from the environment. Environ. Microbiol. Rep. 11, 487–494 (2019).
    PubMed  Article  PubMed Central  Google Scholar 

    21.
    Debeljak, P. et al. Extracting DNA from ocean microplastics: a method comparison study. Anal. Methods 9, 1521–1526 (2017).
    CAS  Article  Google Scholar 

    22.
    Weisburg, W. G., Barns, S. M., Pelletier, D. A. & Lane, D. J. 16S ribosomal DNA amplification for phylogenetic study. J. Bacteriol. 173, 697–703 (1991).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    23.
    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).
    PubMed  PubMed Central  Google Scholar 

    24.
    Hadziavdic, K. et al. Characterization of the 18S rRNA gene for designing universal eukaryote specific primers. PLoS ONE 9, e87624 (2014).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    25.
    Vieira, H. H. et al. tufA gene as molecular marker for freshwater Chlorophyceae. Algae 31, 155–165 (2016).
    CAS  Article  Google Scholar 

    26.
    De Beeck, M. O. et al. Comparison and validation of some ITS primer pairs useful for fungal metabarcoding studies. PLoS ONE 9, e97629 (2014).
    ADS  Article  Google Scholar 

    27.
    Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet J. 17, 10–12 (2011).
    Article  Google Scholar 

    28.
    De Coster, W., D’Hert, S., Schultz, D. T., Cruts, M. & Van Broeckhoven, C. NanoPack: visualizing and processing long-read sequencing data. Bioinformatics 34, 2666–2669 (2018).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    29.
    Schmieder, R. & Edwards, R. Quality control and preprocessing of metagenomic datasets. Bioinformatics 27, 863–864 (2011).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    30.
    Baloğlu, B. et al. A workflow for accurate metabarcoding using nanopore MinION sequencing. BioRxiv. https://doi.org/10.1101/2020.05.21.108852 (2020).
    Article  Google Scholar 

    31.
    Srivathsan, A. et al. A Min IONTM-based pipeline for fast and cost-effective DNA barcoding. Mol. Ecol. Resour. 18, 1035–1049 (2018).
    CAS  Article  Google Scholar 

    32.
    Maestri, S. et al. A rapid and accurate MinION-based workflow for tracking species biodiversity in the field. Genes 10, 468 (2019).
    CAS  PubMed Central  Article  Google Scholar 

    33.
    Bolyen, E. et al. Reproducible, interactive, scalable and extensible microbiome data science using QIIME 2. Nat. Biotechnol. 37, 852–857 (2019).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    34.
    Voorhuijzen-Harink, M. M. et al. Toward on-site food authentication using nanopore sequencing. Food Chem. X2 (2019).

    35.
    Nilsson, R. H. et al. The UNITE database for molecular identification of fungi: handling dark taxa and parallel taxonomic classifications. Nucleic Acids Res. 47, 259–264 (2019).
    Article  CAS  Google Scholar 

    36.
    Sauvage, T., Schmidt, W. E., Suda, S. & Fredericq, S. A metabarcoding framework for facilitated survey of endolithic phototrophs with tufA. BMC Ecol. 16, 8 (2016).
    PubMed  PubMed Central  Article  Google Scholar 

    37.
    Heller, P., Casaletto, J., Ruiz, G. & Geller, J. A database of metazoan cytochrome c oxidase subunit I gene sequences derived from GenBank with CO-ARBitrator. Sci. Data 5, 180156 (2018).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    38.
    McMurdie, P. J. & Holmes, S. phyloseq: an R package for reproducible interactive analysis and graphics of microbiome census data. PLoS ONE 8, e61217 (2013).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    39.
    Andersen, K. S., Kirkegaard, R. H., Karst, S. M. & Albertsen, M. ampvis2: an R package to analyse and visualise 16S rRNA amplicon data. BioRxiv, 299537 (2018).

    40.
    R Core Team. R: a language and environment for statistical computing (R Foundation for Statistical Computing, Vienna, 2014).
    Google Scholar 

    41.
    Mafune, K. K., Godfrey, B. J., Vogt, D. J. & Vogt, K. A. A rapid approach to profiling diverse fungal communities using the MinION™ nanopore sequencer. BioTechniques 68, 72–78 (2019).
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    42.
    Herbst, F. A. et al. Elucidation of in situ polycyclic aromatic hydrocarbon degradation by functional metaproteomics (protein-SIP). Proteomics 13, 2910–2920 (2013).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    43.
    Jin, H. M., Kim, J. M., Lee, H. J., Madsen, E. L. & Jeon, C. O. Alteromonas as a key agent of polycyclic aromatic hydrocarbon biodegradation in crude oil-contaminated coastal sediment. Environ. Sci. Technol. 46, 7731–7740 (2012).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    44.
    Lin, X., Yang, B., Shen, J. & Du, N. Biodegradation of crude oil by an Arctic psychrotrophic bacterium Pseudoalteromomas sp. P29. Curr. Microbiol. 59, 341–345 (2009).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    45.
    Hedlund, B. P. & Staley, J. T. Isolation and characterization of Pseudoalteromonas strains with divergent polycyclic aromatic hydrocarbon catabolic properties. Environ. Microbiol. 8, 178–182 (2006).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    46.
    Schneiker, S. et al. Genome sequence of the ubiquitous hydrocarbon-degrading marine bacterium Alcanivorax borkumensis. Nat. Biotechnol. 24, 997–1004 (2006).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    47.
    Yakimov, M. M. et al. Alcanivorax borkumensis gen. nov., sp. nov., a new, hydrocarbon-degrading and surfactant-producing marine bacterium. Int. J. Syst. Evolut. Microbiol. 48, 339–348 (1998).
    CAS  Google Scholar 

    48.
    Delacuvellerie, A., Cyriaque, V., Gobert, S., Benali, S. & Wattiez, R. The plastisphere in marine ecosystem hosts potential specific microbial degraders including Alcanivorax borkumensis as a key player for the low-density polyethylene degradation. J. Hazard. Mater. 380, 120899 (2019).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    49.
    Wangensteen, O. S. & Turon, X. Metabarcoding techniques for assessing biodiversity of marine animal forests. Mar. Anim. For. Ecol. Benthic Biodivers. Hotspots 1, 445–503 (2017).
    Article  Google Scholar 

    50.
    Truelove, N. K., Andruszkiewicz, E. A. & Block, B. A. A rapid environmental DNA method for detecting white sharks in the open ocean. Methods Ecol. Evol. 10, 1128–1135 (2019).
    Article  Google Scholar 

    51.
    Gillespie, R. et al. Nanopore sequencing of long ribosomal DNA amplicons enables portable and simple biodiversity assessments with high phylogenetic resolution across broad taxonomic scale. GigaScience 8, giz006 (2019).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    52.
    Kono, N. & Arakawa, K. Nanopore sequencing: review of potential applications in functional genomics. Dev. Growth Differ. 61, 316–326 (2019).
    PubMed  Article  PubMed Central  Google Scholar 

    53.
    Nair, S. A., Devassy, V., Dwivedi, S. & Selvakumar, R. Preliminary observations on tar-like material observed on some beaches. Curr. Sci. India 41, 766–767 (1972).
    Google Scholar 

    54.
    Kasai, Y. et al. Predominant growth of Alcanivorax strains in oil-contaminated and nutrient-supplemented sea water. Environ. Microbiol. 4, 141–147 (2002).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    55.
    Reisser, J. et al. Millimeter-sized marine plastics: a new pelagic habitat for microorganisms and invertebrates. PLoS ONE 9, e100289 (2014).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    56.
    Masó, M., Fortuño, J. M., de Juan, S. & Demestre, M. Microfouling communities from pelagic and benthic marine plastic debris sampled across Mediterranean coastal waters. Sci. Mar. 80, 117–127 (2016).
    Article  Google Scholar 

    57.
    Wang, S. et al. The interactions between microplastic polyvinyl chloride and marine diatoms: physiological, morphological, and growth effects. Ecotoxicol. Environ. Saf. 203, 111000 (2020).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    58.
    De Tender, C. et al. A review of microscopy and comparative molecular-based methods to characterize “Plastisphere” communities. Anal. Methods 9, 2132–2143 (2017).
    Article  CAS  Google Scholar  More

  • in

    Relationships between nitrogen cycling microbial community abundance and composition reveal the indirect effect of soil pH on oak decline

    1.
    van Mantgem PJ, Stephenson NL, Byrne JC, Daniels LD, Franklin JF, Fule PZ, et al. Widespread increase of tree mortality rates in the western United States. Science. 2009;323:521–4.
    PubMed  Article  CAS  PubMed Central  Google Scholar 
    2.
    Allen CD, Macalady AK, Chenchouni H, Bachelet D, McDowell N, Vennetier M, et al. A global overview of drought and heat-induced tree mortality reveals emerging climate change risks for forests. For Ecol Manag. 2010;259:660–84.
    Article  Google Scholar 

    3.
    Carnicer J, Coll M, Ninyerola M, Pons X, Sanchez G, Penuelas J. Widespread crown condition decline, food web disruption, and amplified tree mortality with increased climate change-type drought. Proc Natl Acad Sci USA. 2011;108:1474–8.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    4.
    Brown N, Vanguelova E, Parnell S, Broadmeadow S, Denman S. Predisposition of forests to biotic disturbance: predicting the distribution of Acute Oak Decline using environmental factors. For Ecol Manag. 2018;407:145–54.
    Article  Google Scholar 

    5.
    Denman S, Doonan J, Ransom-Jones E, Broberg M, Plummer S, Kirk S, et al. Microbiome and infectivity studies reveal complex polyspecies tree disease in Acute Oak Decline. ISME J. 2017. https://doi.org/10.1038/ismej.2017.170.

    6.
    Denman S, Barrett G, Kirk SA, McDonald JE, Coetzee MPA. Identification of Armillaria species on oak in Britain: implications for Oak Health. Forestry. 2017;90:148–61.
    Article  Google Scholar 

    7.
    Martınez-Vilalta J, Lloret F, Breshears DD. Drought-induced forest decline: causes, scope and implications. Biol Lett. 2012;8:689–91.
    PubMed  Article  PubMed Central  Google Scholar 

    8.
    McDowell NG, Ryan MG, Zeppel MJB, Tissue DT. Improving our knowledge of drought-induced forest mortality through experiments, observations, and modeling. N. Phytologist. 2013;200:289–93.
    Article  Google Scholar 

    9.
    Thomas FM, Blank R, Hartmann G. Abiotic and biotic factors and their interactions as causes of oak decline in central Europe. Pathol. 2002;32:277–307.
    Article  Google Scholar 

    10.
    Niinemets Ü. Responses of forest trees to single and multiple environmental stresses from seedlings to mature plants: past stress history, stress interactions, tolerance and acclimation. Ecol Manag. 2010;260:1623–39.
    Article  Google Scholar 

    11.
    Amoroso MM, Daniels LD, Larson BC. Temporal patterns of radial growth in declining Austrocedrus chilensis forests in Northern Patagonia: the use of treerings as an indicator of forest decline. Ecol Manag. 2012;265:62–70.
    Article  Google Scholar 

    12.
    Bansal S, Hallsby G, Löfvenius MO, Nilsson MC. Synergistic, additive and antagonistic impacts of drought on herbivory on Pinus sylvestris: leaf, tissue and whole-plant responses and recovery. Tree Physiol. 2013;33:451–63.
    CAS  PubMed  Article  Google Scholar 

    13.
    Whyte G, Howard K, Hardy GEStJ, Burgess T. The Tree Decline Recovery Seesaw; a conceptual model of the decline and recovery of drought stressed plantation trees. For Ecol Manag. 2016;370:102–13.
    Article  Google Scholar 

    14.
    Calder JA, Kirkpatrick JB. Climate change and other factors influencing the decline of the Tasmanian cider gum (Eucalyptus gunnii). Australian J Botany. 2008;56. https://doi.org/10.1071/BT08105.

    15.
    Avila JM, Gallardo A, Ibáñez B, Gómez‐Aparicio L. Quercus suber dieback alters soil respiration and nutrient availability in Mediterranean forests. J Ecol. 2016;104:1441–52.
    Article  Google Scholar 

    16.
    Crawford N. Nitrate: nutrient and signal for plant growth. Plant Cell. 1995;7:859–68.
    CAS  PubMed  PubMed Central  Google Scholar 

    17.
    Lovett GM, Arthur MA, Weathers KC, Griffin JM. Long-term changes in forest carbon and nitrogen cycling caused by an introduced pest/pathogen complex. Ecosystems. 2010;13:1188–1200.
    CAS  Article  Google Scholar 

    18.
    Throop H, Lerdau MT. Effects of nitrogen deposition on insect herbivory: Implications for community and ecosystem processes. Ecosystems. 2004;7:109–33.
    CAS  Article  Google Scholar 

    19.
    Thomas FM, Ahlers U. Effects of excess nitrogen on frost hardiness and freezing injury of above-ground tissue in young oaks (Quercus petraea and Q. robur). N. Phytologist. 1999;144:73–83.
    Article  Google Scholar 

    20.
    Hardham AR. The cell biology behind Phytophthora pathogenicity. Australas Plant Pathol. 2001;30:91–98.
    Article  Google Scholar 

    21.
    Brown N, Jeger M, Kirk S, Xu X, Denman S. Spatial and temporal patterns in symptom expression within eight woodlands affected by acute Oak Decline. For Ecol Manag. 2016;360:97–109.
    Article  Google Scholar 

    22.
    Scarlett K, Guest DI, Daniel R. Elevated soil nitrogen increases the severity of dieback due to Phytophthora cinnamomi. Australas Plant Pathol. 2013;42:155–62.
    Article  Google Scholar 

    23.
    Yao H, Bowman D, Shi W. Seasonal variations in soil microbial biomass and activity in warm and cool season turfgrass systems. Soil Biol Biochem. 2011;43:1536–43.
    CAS  Article  Google Scholar 

    24.
    Prosser JI, Nicol GW. Relative contributions of archaea and bacteria to aerobic ammonia oxidation in the environment. Environ Microbiol. 2008;10:2931–41.
    CAS  PubMed  Article  Google Scholar 

    25.
    Prosser JI, Nicol GW. Archaeal and bacterial ammonia oxidisers in soil: the quest for niche specialisation and differentiation. Trends Microbiol. 2012;20:523–31.
    CAS  PubMed  Article  Google Scholar 

    26.
    Erguder TH, Boon N, Wittebolle L, Marzorati M, Verstraete W. Environmental factors shaping the ecological niches of ammonia‐oxidizing archaea. FEMS Microbiol Rev. 2009;33:855–69.
    CAS  PubMed  Article  Google Scholar 

    27.
    Hink L, Lycus P, Gubry-Rangin C, Frostgard A, Nicol GW, Prosser JI, et al. Kinetics of NH3‐oxidation, NO‐turnover, N2O‐production and electron flow during oxygen depletion in model bacterial and archaeal ammonia oxidisers. Environ Microbiol. 2017;19:4882–96.
    CAS  PubMed  Article  Google Scholar 

    28.
    Gubry-Rangin C, Hai B, Quince C, Engel M, Thomson BC, James P, et al. Niche specialization of terrestrial archaeal ammonia oxidisers. Proc Natl Acad Sci. 2011;108:21206–11.
    CAS  PubMed  Article  Google Scholar 

    29.
    Leininger S, Schloter UT, Schwark I, Qi J, Nicol GW, Prosser JI, et al. Archaea predominate among ammonia-oxidizing prokaryotes in soils. Nature. 2006;442:806–9.
    CAS  PubMed  Article  Google Scholar 

    30.
    Gubry-Rangin C, Nicol GW, Prosser JI. Archaea rather than bacterial control nitrification in two agricultural acidic soils. FEMS Microbiol Ecol. 2010;74:566–74.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    31.
    Verhamme DT, Prosser JI, Nicol GW. Ammonia concentration determines differential growth of ammonia oxidizing archaeal and bacteria in soil microcosms. ISME J. 2011;5:1067–71.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    32.
    Di HJ, Cameron KC, Shen J-P, Winefield CS, O’Callaghan M, Bowatte S, et al. Ammonia-oxidizing bacteria and archaea grow under contrasting soil nitrogen conditions. FEMS Microbiol Ecol. 2010;72:386–94.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    33.
    Hink L, Gubry-Rangin C, Nicol GW, Prosser J. The consequences of niche and physiological differentiation of archaeal and bacterial ammonia oxidisers for nitrous oxide emissions. ISME J. 2018;12:1084–93.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    34.
    Clark D, McKew B, Dong L, Leung G, Dumbrell AJ, Stott A, et al. Mineralization and nitrification: archaea dominate ammonia-oxidising communities in grassland soils. Soil Biol Biochem. 2020;143:107725.
    CAS  Article  Google Scholar 

    35.
    Delgado-Baquerizo M, Maestre FT, Eldridge DJ, Singh BK. Microsite differentiation drives the abundance of soil ammonia oxidizing bacteria along aridity gradients. Front Microbiol. 2016;7:505.
    PubMed  PubMed Central  Article  Google Scholar 

    36.
    Martens-Habbena W, Berube PM, Urakawa H, de la Torre JR, Stahl DA. Ammonia oxidation kinetics determine niche separation of nitrifying Archaea and Bacteria. Nature. 2009;461:976–9.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    37.
    Barta J, Tahovska K, Santruckova H, Oulehle F. Microbial communities with distinct denitrification potential in spruce and beech soils differing in nitrate leaching. Sci Rep Nat. 2017;7:9738.
    Article  CAS  Google Scholar 

    38.
    Fierer N, Lauber CL, Ramirez KS, Zaneveld J, Bradford MA, Knight R. Comparative metagenomic, phylogenetic and physiological analyses of soil microbial communities across nitrogen gradients. ISME J. 2012;6:1007–17. https://doi.org/10.1038/ismej.2011.159.
    CAS  Article  PubMed  Google Scholar 

    39.
    Ramirez KS, Lauber CL, Knight R, Bradford MA, Fierer N. Consistent effects of nitrogen fertilization on soil bacterial communities in contrasting systems. Ecology. 2010;91:3463–70. https://doi.org/10.1890/10-0426.1.
    Article  PubMed  PubMed Central  Google Scholar 

    40.
    Cranfield University 2020. The Soils Guide. www.landis.org.uk. UK; Cranfield University.

    41.
    G Kerr, J Haufe. Thinning practice. A Silvicultural Guide. Bristol: Forestry Commission; 2011;1:54.

    42.
    Cools N, De Vos B Sampling and Analysis of Soil. In: Manual on methods and criteria for harmonized sampling, assessment, monitoring and analysis of the effects of air pollution on forests. Hamburg: UNECE, ICP Forests; 2010, pp. 208.

    43.
    MAFF. Code of good agricultural practice for the protection of soil. London, UK: Ministry of Agriculture, Fisheries and Food; 1993.
    Google Scholar 

    44.
    Li J, Nedwell DB, Beddow J, Dumbrell AJ, McKew BA, Thorpe EL, et al. amoA gene abundances and nitrification potential rates suggest that benthic ammonia-oxidizing bacteria (AOB) not archaea (AOA) dominate N cycling in the Colne estuary, UK. Appl Environ Microbiol. 2015;81:159–65.
    PubMed  Article  CAS  Google Scholar 

    45.
    Beddow J, Stolpe B, Cole PA, Lead JR, Sapp M, Lyons BP, et al. Nanosilver inhibits nitrification and reduces ammonia-oxidizing bacterial but not archaeal amoA gene abundance in estuarine sediments. Environ Microbiol. 2017;19:500–10.
    CAS  PubMed  Article  Google Scholar 

    46.
    Tourna M, Freitag TE, Nicol GW, Prosser JI. Growth, activity and temperature responses of ammonia-oxidizing archaea and bacteria in soil microcosms. Environ Microbiol. 2008;10:1357–64.
    CAS  PubMed  Article  Google Scholar 

    47.
    Rotthauwe JH, Witzel KP, Liesack W. The ammonia monooxygenase structural gene amoA as a functional marker: molecular fine-scale analysis of natural ammonia-oxidizing populations. Appl Environ Microbiol. 1997;63:4704–12.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    48.
    Throbäck IN, Enwall K, Jarvis A, Hallin S. Reassessing PCR primers targeting nirS, nirK and nosZ genes for community surveys of denitrifying bacteria with DGGE. FEMS Microbiol Ecol. 2004;49:401–17.
    PubMed  Article  CAS  Google Scholar 

    49.
    Braker G, Fesefeldt A, Witzel KP. Development of PCR primer systems for amplification of nitrite reductase genes (nirK and nirS) to detect denitrifying bacteria in environmental samples. Appl Environ Microbiol. 1998;64:3769–75.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    50.
    Henry S, Bru D, Stres B, Hallet S, Philippot L. Quantitative detection of the nosZ gene, encoding nitrous oxide reductase, and comparison of the abundances of 16S rRNA, narG, nirK, and nosZ Genes in Soils. Appl Environ Microbiol. 2006;72:5181–9.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    51.
    Herlemann D, Labrenz M, Jürgens K, Bertilsson S, Waniek J, Andersson A. Transitions in bacterial communities along the 2000 km salinity gradient of the Baltic Sea. ISME J. 2011;5:1571–9.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    52.
    Raskin L, Stromley JM, Rittmann BE, Stahl DA. Group-specific 16S rRNA hybridization probes to describe natural communities of methanogens. Appl Environ Microbiol. 1994;60:1232–40.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    53.
    Stahl DA, Amann R Development and application of nucleic acid probes. In: Nucleic acid techniques in bacterial systematics. Stackebrandt, E, Goodfellow M, editors. Chichester, UK: John Wiley & Sons Ltd; 1991. pp. 205–48.

    54.
    Dumbrell AJ, Ferguson RMW, Clark DR. Microbial community analysis by single-amplicon high-throughput next generation sequencing: Data analysis—from raw output to ecology. In: McGenity T, Timmis K, Nogales B, editors. Hydrocarbon and Lipid Microbiology Protocols. Berlin, Heidelberg: Springer Protocols Handbooks. Springer; 2016. 155–206.

    55.
    Joshi NA, Fass JN Sickle: A Sliding-Window, Adaptive, Quality-Based Trimming Tool for FastQ Files 2011; (Version 1.33).

    56.
    Nurk S, Bankevich A, Antipov D, Gurevich AA, Korobeynikov A, Lapidus A, et al. Assembling single-cell genomes and mini-metagenomes from chimeric MDA products. J Comput Biol. 2013;20:714–37.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    57.
    Nikolenko SI, Korobeynikov AI, Alekseyev MA. BayesHammer: Bayesian clustering for error correction in single-cell sequencing. BMC Genomics SP. 2013;14:S7.
    Article  Google Scholar 

    58.
    Rognes T, Flouri T, Nichols B, Quince C, Mahé F VSEARCH: a versatile open source tool for metagenomics. PeerJ. 2016;4. https://doi.org/10.7717/peerj.2584.

    59.
    Edgar RC, Haas BJ, Clemente JC, Quince C, Knight R. UCHIME improves sensitivity and speed of chimera detection. Bioinformatics. 2011;16:2194–200.
    Article  CAS  Google Scholar 

    60.
    Wang Q, Quensen JF, Fish JA, Lee TK, Sun Y, Tiedje JM, et al. Ecological patterns of nifH genes in four terrestrial climatic zones explored with targeted metagenomics using FrameBot, a new informatics tool. MBio. 2013;4:e00592–13.
    PubMed  PubMed Central  Google Scholar 

    61.
    Wang Q, Garrity GM, Tiedje JM, Cole JR. Naive Bayesian classifier for rapid assignment of rRNA sequences into the new bacterial taxonomy. Appl Environ Microbiol. 2007;16:5261–7.
    Article  CAS  Google Scholar 

    62.
    Edgar RC. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004;5:1792–7.
    Article  CAS  Google Scholar 

    63.
    Tamura K, Stecher G, Peterson D, Filipski A, Kumar S. MEGA6: molecular evolutionary genetics analysis version 6.0. Mol Biol Evol. 2013;12:2725–9.
    Article  CAS  Google Scholar 

    64.
    Fish J, Chai B, Wang Q, Sun Y, Brown CT, Tiedje JM, et al. FunGene: the functional gene pipeline and repository. Front Microbiol. 2013;4:291.
    PubMed  PubMed Central  Article  Google Scholar 

    65.
    Altschul S, Madden TL, Schäffer AA, Zhang J, Zhang Z, Miller W, et al. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 1997;25:3389–402.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    66.
    Lefcheck J. PIECEWISESEM: Piecewise structural equation modelling in R for ecology, evolution, and systematics. Methods Ecol Evol. 2016;7:573–9.
    Article  Google Scholar 

    67.
    Shipley B. A new inferential test for path models based on directed acyclic graphs. Struct Equ Modeling. 2000;7:206–18.
    Article  Google Scholar 

    68.
    Grace JB. Structural Equation Modelling and Natural Systems. New York, NY: Cambridge University Press; 2006.
    Google Scholar 

    69.
    Oksanen J, Blanchet FG, Friendly M, Kindt R, Legendre P, McGlinn D, et al. Vegan: Community Ecology Package. R package Version. 2017;2:4–3.
    Google Scholar 

    70.
    Leininger Wang Y, Naumann U, Wright S, Warton D. Mvabund—an R package for model‐based analysis of multivariate abundance data. Methods Ecol Evolution. 2012;3:471–4.
    Article  Google Scholar 

    71.
    Lehtovirta-Morley LE, Stoecker K, Vilcinskas A, Prosser JI, Nicol GW. Cultivation of an obligate acidophilic ammonia oxidizer from a nitrifying acid soil. Proc Natl Acad Sci USA. 2011;108:15892–7.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    72.
    Hu H, Zhang L, Dai Y, et al. pH-dependent distribution of soil ammonia oxidizers across a large geographical scale as revealed by high-throughput pyrosequencing. J Soils Sediment. 2013;13:1439–49.
    Article  CAS  Google Scholar 

    73.
    Hu BL, Liu S, Wang W, Shen LD, Lou LP, Liu WP, et al. pH-dominated niche segregation of ammonia-oxidising microorganisms in Chinese agricultural soils. FEMS Microbiol Ecol. 2014;90:290–9. https://doi.org/10.1111/1574-6941.12391.
    CAS  Article  Google Scholar 

    74.
    Hu H, Zhang L, Yuan C, Zheng Y, Wang J, Chen D, et al. The large-scale distribution of ammonia oxidizers in paddy soils is driven by soil pH, geographic distance, and climatic factors. Front Microbiol. 2015;6:938.
    PubMed  PubMed Central  Google Scholar 

    75.
    Delgado-Baquerizo M, Gallardo A, Wallenstein MD, Maestre FT. Vascular plants mediate the effects of aridity and soil properties on ammonia-oxidizing bacteria and archaea. FEMS Microbiol Ecol. 2013;13:273–82.
    Article  CAS  Google Scholar 

    76.
    Eldridge DJ, Beecham G, Grace J. Do shrubs reduce the adverse effects of grazing on soil properties? Ecohydrology. 2015;8:1503–13.
    Article  Google Scholar 

    77.
    Berdugo M, Soliveres S, Maestre FT. Vascular plants and biocrusts modulate how abiotic factors affect wetting and drying events in drylands. Ecosystems. 2014;17:1242–56.
    CAS  Article  Google Scholar 

    78.
    Köhler S, Levia DF, Jungkunst HF, Gerold G. An In Situ Method to Measure and Map Bark pH. J Wood Chem Technol. 2015;35:438–49.
    Article  CAS  Google Scholar 

    79.
    Matschonat G, Falkengren-Grerup U. Recovery of soil pH, Cation-exchange Capacity and the Saturation of Exchange Sites from Stemflow-induced Soil Acidification in Three Swedish Beech (Fagus sylvatica L.) Forests. Scand J For Res. 2000;15:39–48.
    Article  Google Scholar 

    80.
    Wang Y, Uchida Y, Shimomura U, Akiyama H, Hayatsu M. Responses of denitrifying bacterial communities to short-term waterlogging of soils. Sci Rep. 2017;7:803.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    81.
    Liu J, Yu Z, Yao Q, Sui Y, Shi Y, Chu H, et al. Ammonia-oxidizing Archaea show more distinct biogeographic distribution patterns than ammonia-oxidizing bacteria across the black soil zone of Northeast China. Front Microbiol. 2018;9:171.
    PubMed  PubMed Central  Article  Google Scholar 

    82.
    Shen C, Xiong J, Zhang H, Feng Y, Lin X, Li X, et al. Soil pH drives the spatial distribution of bacterial communities along elevation on Changbai Mountain. Soil Biol Biochem. 2013;57:204–11.
    CAS  Article  Google Scholar 

    83.
    Nicol GW, Leininger S, Schleper C, Prosser JI. The influence of soil pH on the diversity, abundance and transcriptional activity of ammonia oxidizing archaea and bacteria. Environ Microbiol. 2008;10:2966–78.
    CAS  PubMed  Article  Google Scholar 

    84.
    Rousk J, Baath E, Brookes PC, Lauber CL, Lozupone C, Caporaso JG, et al. Soil bacterial and fungal communities across a pH gradient in an arable soil. ISME J. 2010;4:1340–51.
    PubMed  Article  Google Scholar 

    85.
    Yuan YL, Si GC, Wang J, Luo TX, Zhang GX. Bacterial community in alpine grasslands along an altitudinal gradient on the Tibetan Plateau. FEMS Microbiol Ecol. 2014;87:121–32.
    CAS  PubMed  Article  Google Scholar 

    86.
    Kaiser k, Wemheuer B, Korolkow V, Wemheuer F, Nacke H, Schöning I, et al. Driving forces of soil bacterial community structure, diversity, and function in temperate grasslands and forests. Sci Rep. 2017;7:9738.
    Article  CAS  Google Scholar 

    87.
    Meaden S, Metcalf CJE, Koskella B. The effects of host age and spatial location on bacterial community composition in the English Oak tree (Quercus robur). Environ Microbiol Rep. 2016;8:649–58.
    CAS  PubMed  Article  Google Scholar 

    88.
    Patra AK, Abbadie L, Clays-Josserand A, Degrange V, Grayston SJ, Guillaumaud N, et al. Effects of management regime and plant species on the enzyme activity and genetic structure of N-fixing, denitrifying and nitrifying bacterial communities in grassland soils. Environ Microbiol. 2006;8:1005–16.
    CAS  PubMed  Article  Google Scholar 

    89.
    Bremer C, Braker G, Matthies D, Beierkuhnlein C, Conrad R. Plant presence and species combination, but not diversity, influence denitrifier activity and the composition of nirK-type denitrifier communities in grassland soil. FEMS Microbiol Ecol. 2009;70:377–87.
    CAS  PubMed  Article  Google Scholar  More

  • in

    The role of kinship and demography in shaping cooperation amongst male lions

    1.
    Hamilton, W. D. The genetic theory of social behaviour I and II. J. Theor. Biol. 7, 1–52 (1964).
    CAS  PubMed  Article  Google Scholar 
    2.
    Ward, A., & Webster, M. Sociality: the behaviour of group-living animals (Springer,2016).

    3.
    Lehmann, L., & Keller, L. The evolution of cooperation and altruism. A general framework and classification of models. J. Evol. Biol.19, 1365–1378 (2006).

    4.
    Sueur, C. et al. Collective decision-making and fission–fusion dynamics: a conceptual framework. Oikos 120, 1608–1617 (2011).
    Article  Google Scholar 

    5.
    Jones, T. B. et al. Consistent sociality but flexible social associations across temporal and spatial foraging contexts in a colonial breeder. Ecol. Lett. https://doi.org/10.1111/ele.13507 (2020).
    Article  PubMed  Google Scholar 

    6.
    Carter, G. G. & Wilkinson, G. S. Food sharing in vampire bats: reciprocal help predicts donations more than relatedness or harassment. Proc. R. Soc. B. 280, 20122573. https://doi.org/10.1098/rspb.2012.2573 (2013).
    Article  PubMed  Google Scholar 

    7.
    Baglione, V., Canestrari, D., Marcos, J. M. & Ekman, J. Kin selection in cooperative alliances of carrion crows. Science 300, 1947–1949 (2003).
    ADS  CAS  PubMed  Article  Google Scholar 

    8.
    Wahaj, S. A. et al. Kin discrimination in the spotted hyena (Crocuta crocuta): nepotism among siblings. Behav. Ecol. Sociobiol. 56, 237–247 (2004).
    Article  Google Scholar 

    9.
    East, M. L. et al. Maternal effects on offspring social status in spotted hyenas. Behav. Ecol. 20, 478–483 (2009).
    Article  Google Scholar 

    10.
    Komdeur, J., Burke, T., Dugdale, H.L., & Richardson, D.S. Seychelles warblers: Complexities of the helping paradox in Cooperative breeding in vertebrates: studies of ecology, evolution and behavior (ed. Koenig, W. D. & Dickinson, J. L.) 197–216 (Cambridge University Press, 2016).

    11.
    Krakauer, A. H. Kin selection and cooperative courtship in wild turkeys. Nature 434, 69–72 (2005).
    ADS  CAS  PubMed  Article  Google Scholar 

    12.
    De Moor, D., Roos, C., Ostner, J. & Schülke, O. Bonds of bros and brothers: kinship and social bonding in post-dispersal male macaques. Mol. Ecol. https://doi.org/10.1111/mec.15560 (2020).
    Article  PubMed  Google Scholar 

    13.
    Koykka, C. & Wild, W. Concessions, lifetime fitness consequences, and the evolution of coalitionary behaviour. Behav. Ecol. 28, 20–30 (2016).
    Article  Google Scholar 

    14.
    Clutton-Brock, T. Cooperation between non-kin in animal societies. Nature 46, 51–57 (2009).
    ADS  Article  CAS  Google Scholar 

    15.
    Schaller, G.B. The Serengeti Lion: a study of predator-prey relations. (University of Chicago Press, 1972).

    16.
    Bertram, B. C. Social factors influencing reproduction in wild lions. J. Zool. 177, 463–482 (1975).
    Article  Google Scholar 

    17.
    Bygott, J. D., Bertram, B. C. & Hanby, J. P. Male lions in large coalitions gain reproductive advantages. Nature 282, 839 (1979).
    ADS  Article  Google Scholar 

    18.
    Packer, C. & Pusey, A. E. Cooperation and competition within coalitions of male lions: Kin selection or game theory?. Nature 296, 740 (1982).
    ADS  Article  Google Scholar 

    19.
    Grinnell, J., Packer, C. & Pusey, A. E. Cooperation in male lions: kinship, reciprocity or mutualism?. Anim. Behav. 49, 95–105 (1995).
    Article  Google Scholar 

    20.
    Chakrabarti, S. & Jhala, Y. V. Selfish partners: resource partitioning in male coalitions of Asiatic lions. Behav. Ecol. 28, 1532–1539 (2017).
    PubMed  PubMed Central  Article  Google Scholar 

    21.
    Packer, C. et al. Reproductive success of lions in Reproductive success (ed. Clutton-Brock, T.H.) 363–383 (University of Chicago Press, 1988).

    22.
    Packer, C., Gilbert, D. A., Pusey, A. E. & O’Brien, S. J. A molecular genetic analysis of kinship and cooperation in African lions. Nature 351, 562–565 (1991).
    ADS  CAS  Article  Google Scholar 

    23.
    Connor, R. C., Smolker, R. A., & Richards, A. F. Two levels of alliance formation among male bottlenose dolphins (Tursiops sp.). PNAS. 89, 987–990 (1992).

    24.
    Parsons, K. M. et al. Kinship as a basis for alliance formation between male bottlenose dolphins, Tursiops truncatus, in the Bahamas. Anim. Behav. 66, 185–194 (2003).
    Article  Google Scholar 

    25.
    Widdig, A., Streich, W. J. & Tembrock, G. Coalition formation among male Barbary macaques (Macaca sylvanus). Am. J. Primatol. 50, 37–51 (2000).
    CAS  PubMed  Article  Google Scholar 

    26.
    Gottelli, D., Wang, J., Bashir, S. & Durant, S. M. Genetic analysis reveals promiscuity among female cheetahs. Proc. R. Soc. B. 274, 1993–2001 (2007).
    PubMed  Article  Google Scholar 

    27.
    Bertram, B.C. Pride of lions. (JM Dent and Sons Ltd, 1978).

    28.
    O’Brien, S.J. Prides and Prejudice in Tears of the cheetah and other tales from the genetic frontier: the genetic secrets of our animal ancestors (ed. O’Brien, S.J.) 35–55 (Thomas Dunne Books, 2003).

    29.
    de Manuel, M. et al. The evolutionary history of extinct and living lions. PNAS 117, 10927–10934 (2020).
    PubMed  Article  CAS  Google Scholar 

    30.
    Clutton-Brock, T. H. Reproductive skew, concessions and limited control. Trends Ecol. Evol. 13, 288–292 (1998).
    CAS  PubMed  Article  Google Scholar 

    31.
    Queller, D. C. & Keith, F. G. Estimating relatedness using genetic markers. Evolution 43, 258–275 (1989).
    PubMed  Article  Google Scholar 

    32.
    Wang, J. Estimating pairwise relatedness in a small sample of individuals. Heredity 119, 302–313 (2017).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    33.
    Sandel, A. A., Langergraber, K. E. & Mitani, J. C. Adolescent male chimpanzees (Pan troglodytes) form social bonds with their brothers and others during the transition to adulthood. Am. J. Primatol. 82, 23091. https://doi.org/10.1002/ajp.23091 (2020).
    Article  Google Scholar 

    34.
    Dal Pesco, F. Dynamics and fitness benefits of male-male sociality in wild Guinea baboons (Papio papio). (PhD thesis), Georg-August University, Göttingen, Germany.

    35.
    Christakis, N. A. & Fowler, J. H. Friendship and natural selection. PNAS 111, 10796–10801 (2014).
    ADS  CAS  PubMed  Article  Google Scholar 

    36.
    Engh, A. L. et al. Behavioural and hormonal responses to predation in female chacma baboons (Papio hamadryas ursinus). Proc. R. Soc. B. 273, 707–712 (2006).
    CAS  PubMed  Article  Google Scholar 

    37.
    Hill, K. R. et al. Co-residence patterns in hunter-gatherer societies show unique human social structure. Science 331, 1286–1289 (2011).
    ADS  CAS  PubMed  Article  Google Scholar 

    38.
    Silk, J.B. Practicing Hamilton’s Rule: kin selection in primate groups in Cooperation in primates and humans (ed. Kappeler, P.M., & van Schaik, C.P.) 25–46 (Springer, 2006).

    39.
    Chakrabarti, S. et al. Adding constraints to predation through allometric relation of scats to consumption. J. Anim. Ecol. 85, 660–670 (2016).
    PubMed  Article  Google Scholar 

    40.
    Møller, A. P. & Birkhead, T. R. Copulation behaviour in mammals: evidence that sperm competition is widespread. Biol. J. Linn. Soc. 38, 119–131 (1989).
    Article  Google Scholar 

    41.
    Chakrabarti, S. & Jhala, Y. V. Battle of the sexes: a multi-male mating strategy helps lionesses win the gender war of fitness. Behav. Ecol. 30, 1050–1061 (2019).
    Article  Google Scholar 

    42.
    Jhala, Y. V. et al. Asiatic lion: ecology, economics and politics of conservation. Front. Ecol. Evol. 7, 312 (2019).
    ADS  Article  Google Scholar 

    43.
    Boom, R. C. J. A. et al. Rapid and simple method for purification of nucleic acids. J. Clin. Microbiol. 28, 495–503 (1990).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    44.
    Antunes, A. et al. The evolutionary dynamics of the lion Panthera leo revealed by host and viral population genomics. PLoS Genet. 4, 1000251. https://doi.org/10.1371/journal.pgen.1000251 (2008).
    CAS  Article  Google Scholar 

    45.
    Singh, A., Shailaja, K., Gaur, A., & Singh., L. Development and characterization of novel microsatellite markers in the Asiatic lion (Panthera leo persica). Mol. Ecol. Notes. 2, 542–543 (2002).

    46.
    Gaur, A. et al. Twenty polymorphic microsatellite markers in the Asiatic lion (Panthera leo persica). Conserv. Genet. 7, 1005–1008 (2006).
    CAS  Article  Google Scholar 

    47.
    Menotti-Raymond, M. et al. A genetic linkage map of microsatellites in the domestic cat (Felis catus). Genomics 57, 9–23 (1999).
    CAS  PubMed  Article  Google Scholar 

    48.
    Menotti-Raymond, M. et al. An STR forensic typing system for genetic individualization of domestic cat (Felis catus) samples. J. Forensic Sci. 50, 1061–1070 (2005).
    CAS  PubMed  Article  Google Scholar 

    49.
    Williamson, J. E., Huebinger, R. M., Sommer, J. A., Louis, E. E. Jr. & Barber, R. C. Development and cross-species amplification of 18 microsatellite markers in the Sumatran tiger (Panthera tigris sumatrae). Mol. Ecol. Notes. 2, 110–112 (2002).
    CAS  Article  Google Scholar 

    50.
    Drummond, A.J.A.B. et. al. v5. 4. Auckland (2011).

    51.
    Matschiner, M. & Salzburger, W. TANDEM: integrating automated allele binning into genetics and genomics workflows. Bioinformatics 25, 1982–1983 (2009).
    CAS  PubMed  Article  Google Scholar 

    52.
    Taberlet, P. et al. Reliable genotyping of samples with very low DNA quantities using PCR. Nucleic Acids Res. 24, 3189–3194 (1996).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    53.
    Peakall, R.O.D., & Smouse, P.E. GENALEX 6: genetic analysis in Excel. Population genetic software for teaching and research. Mol. Ecol. Notes. 6, 288–295 (2006).

    54.
    Bergner, L. M., Jamieson, I. G. & Robertson, B. C. Combining genetic data to identify relatedness among founders in a genetically depauperate parrot, the Kakapo (Strigops habroptilus). Conserv. Genet. 15, 1013–1020 (2014).
    Article  Google Scholar 

    55.
    Gilbert, D. A., Packer, C., Pusey, A. E., Stephens, J. C. & O’Brien, S. J. Analytical DNA fingerprinting in lions: parentage, genetic diversity, and kinship. J. Hered. 82, 378–386 (1991).
    CAS  PubMed  Article  Google Scholar 

    56.
    Pemberton, J. M., Albon, S. D., Guinness, F. E., Clutton-Brock, T. H. & Dover, G. A. Behavioral estimates of male mating success tested by DNA fingerprinting in a polygynous mammal. Behav. Ecol. 3, 66–75 (1992).
    Article  Google Scholar 

    57.
    Dixson, A. F., Bossi, T. & Wickings, E. J. Male dominance and genetically determined reproductive success in the mandrill (Mandrillus sphinx). Primates 34, 525–532 (1993).
    Article  Google Scholar 

    58.
    Krebs, J.R., & Davies, N.B. An introduction to behavioural ecology. (Blackwell Scientific Publications, 1987).

    59.
    Smith, J. M. Group selection and kin selection. Nature 201, 1145–1147 (1964).
    ADS  Article  Google Scholar 

    60.
    Banerjee, K. & Jhala, Y. V. Demographic parameters of endangered Asiatic lions (Panthera leo persica) in Gir forests India. J. Mammal. 93, 1420–1430 (2012).
    Article  Google Scholar 

    61.
    Meena, V. Reproductive strategy and behaviour of male Asiatic lions. [Dissertation/Ph.D. thesis]. (Forest Research Institute University, 2008).

    62.
    Banerjee, K. Ranging patterns, habitat use and food habits of the satellite lion populations (Panthera leo persica) in Gujarat, India. [Dissertation/Ph.D. thesis]. (Forest Research Institute Deemed University, 2012).

    63.
    Gogoi, K., Kumar, U., Banerjee, K. & Jhala, Y. V. Spatially explicit density and its determinants for Asiatic lions in the Gir forests. PLoS ONE 15, 0228374. https://doi.org/10.1371/journal.pone.0228374 (2020).
    CAS  Article  Google Scholar 

    64.
    R Core Team. R: A language and environment for statistical computing. R Foundation for Statistical Computing. Vienna, Austria. (2019). More

  • in

    Benthic ecosystem cascade effects in Antarctica using Bayesian network inference

    The data for this study were collected in the austral summer of 2016, on board the BAS research ship RRS James Clark Ross38. Biological abundance data was taken from photographic images from Brasier et al.10 (see methodology within). The images were taken with a Shallow Underwater Camera System (SUCS), in transects of 10 photographs, 10 m apart. Replicant transects were separated by 100 m, at water depths of 500 m, 750 m and 1000 m10. Each photograph in the analysis was 0.51 m2 in area. In the analyses, ‘VME unknown’ was biological material unidentifiable to phylum, but distinguishable as VME taxa, e.g., branched or budding fragments which could be bryozoan or cnidarian species. The percentage cover of encrusting species was recorded by Brasier et al.10, as the number of individual colonies was not always possible to distinguish between colonising taxa10. Physical variables, such as substrate texture, were also observed from the SUCS images10.
    Taxonomic resolution of the data originally collected was dependent upon the ability to distinguish the organism in the images10. For our analysis, the taxonomic hierarchy analysed related to the abundance level of that group, as lower abundance, finer scale taxonomic groupings would be zero-inflated so not able to be used within our methodology (cf. Milns et al.39). Taxonomic groupings included were Annelida, Arthropoda, Cnidaria, Echinodermata, Mollusca, Bryozoa, Actinopteri, Porifera and Echinodermata. The echinoderm taxa Ophiuroides and Euryalids also occurred in large enough numbers to be analysed as separate groups. The taxa in each of these groupings can be seen in the Brasier et al.10
    The raw data (taxon ID from photographs) was highly zero-inflated (84.9% of entries), so data grouping was performed to capture the fine-scale (individual photographs at ∼cm scale) and large-scale (replicate transects at ∼km scale) ecology of the ecosystems. For the fine-scale network there were 527 samples (abundances from individual photographs) and twelve nodes. Four were physical nodes (Depth, Region, Substrate and Substrate Texture, Table 1), two were functions of the specimens present (percent encrusting), five were taxa classified to Phylum level (Arthopoda, Bryozoa, Porifera, Cnidaria and Echinodermata) and there was one bin-group (VME Unknown) Supplementary Table 2.
    Table 1 Table of network properties for the two networks found. Chains are defined as series of dependencies containing more than two nodes. Link density is mean number of connections per node, connectance is the number of connections/number of nodes*(number of nodes-1).
    Full size table

    For the large-scale network, the data from all the photographs in each event (up to three replicate transects) was combined to form 21 samples with 16 nodes. For factor nodes (Region, Texture and Substrate) the modal variable was taken, taxa nodes were summed, and the mean was used for Percent Encrusting and Depth. Texture was excluded from analyses due to the high zero-count. Two nodes were functions of the specimens present (percent encrusting), ten taxa were analysed (Annelid, Arthropod, Cnidarian, Echinoderm, Mollusc, Bryozoan, Actinopteri, Ophiuroidea, Euryalida, Porifera) and one bin-group (VME Unknown) was used. The combination of the two networks enabled the investigation of both physical and biotic interactions across multiple spatial scales.
    We chose these relatively coarse taxonomic groupings in order to maximise the statistical power of our analyses. This statistical power enabled us to reconstruct a complex network of dependencies, thus revealing the subtle relationships between different taxa. The coarse grouping is a limitation of this method, because the homogenisation of some diverse groups, such as echinoderms, groups together organisms with variable life modes and traits (and therefore differing relationships with other taxa). However, the coarse level of identification was necessary with this volume of data (over 500 photographs), and the patchiness and rarity of many taxa. Future studies involving a higher density of photographs will allow for genus or species level separation, and/or an analysis based upon functional traits classification.
    Analysis
    One approach to understanding how ecosystems function is to consider the ecosystem as a network (cf. Miln et al.38). Different taxa or groups of taxa are considered ‘nodes’ and their interactions are described as ‘connections’ (more usually known as ‘edges’ in Baysian networks), which link interacting taxa together. A node that the connections feed into are called ‘parents’. Work has centred on gene regulatory networks17, neural information flow networks, and with more recent applications to ecological and palaeoecological networks18,39,40,41. It is important to note that the structure produced by the BNIA reflects the associations caused by co-localisations (two taxa which both have a high abundance), not by a specific interaction, for example predation. By using BNIA, direct dependencies between taxa can be detected, minimising auto-correlation between two nodes. For example, if A depends on B which depends on C, there could be a correlation between A and C. However, this correlation would not represent an interaction or association between A and C, merely the two correlations between A and B and B and C. BNIA enables only the realised dependencies to be found, ensuring only actual interactions and associations are found.
    Bayesian network inference was performed in Banjo16, The BNIA software used was Banjo v2.0.0, a publicly available Java based algorithm39,41. Banjo uses uniform priors, so boundaries for the different discretized groups were chosen to ensure even splitting between the groups. Discretized data were input into Banjo, which then generated a random network based on the input variables. A greedy search was then performed to find a more likely network than the random one generated. This search was repeated 10 million times for each set of input data and the most probable network was then output. The maximum number of parents was set to 3 to limit artefacts19.
    The BNIA used require discrete data, which ensures data noise is masked, and only the relative densities of each taxon are important39. We split the data into three intervals; zero counts, low counts and high counts. Low counts consisted of counts below the median for the taxa group and high counts were counts over the median. Medians were used over means because for some groups the high counts were very high, and would result in a very small number of samples grouped in the highest interval (cf. Milns et al.39). A large amount of bins maintains the amount of information present in the dataset, while fewer bins provide more statistical power, and greater noise masking. Yu19 has shown that for ecological data sets three different bins is a good balance. Zero was treated as a separate entity because the presence of one individual is very different to a zero presence, in contrast to zero gene expression, for example. Data preparation for Banjo (grouping and discretization) was carried out in R42, as was the statistical analysis of the data. Further analysis of banjo outputs, when required, used the functional language Haskell43. The scripts are available on Github (github.com/egmitchell/bootstrap).
    To minimise outliers bias, 100 samples were bootstrapped at 95% level by randomly selecting 95% of the total number of samples for each analysis20,43, and then finding the subsample network using Banjo. For each connection calculated, the probability of occurrence was calculated, and the resultant distributions analysed to find the number of Gaussian sub-distributions using normal mixture models44. This probability distribution was bimodal for each data set, which suggests that there were two distributions of connections, those with low probability of occurrence, and those with highly probable connections. The final network for each area was taken to be those connections which were highly probable. The threshold for being labelled ‘highly probable’ depended on the network (as determined by the normal mixture modelling analyses): 53% for fine-scale network and 51% for the large-scale network. The magnitude of the occurrence rate is indicated in the network by the width of the line depicting the connection.
    The direction of the connection between nodes indicates which node (taxon) has a dependency on the other node (taxon). For each connection, the directionality was taken to be the direction that  occurred in the majority of bootstrapped networks. Where there was no majority (directional connections have a probability between 0.4 and 0.6), the connection was said to have bi-directionality, or indicated a mutual dependency.
    The IS can be used to gauge the type and strength of the interaction between two nodes. If the IS = 1, this corresponds with a positive correlation. When node 1 is high, node 2 will be high. An IS of −1 corresponds to a negative correlation: High node 1 corresponds to a low node 2. An IS = 0 does not mean there is no correlation between the two nodes. IS = 0 means that the interaction is non-monotonic. Sometimes node 1 will be positively correlated with node 2, sometimes negatively. The mean IS for each connection was calculated for each sample area.
    Contingency test filtering
    In order to avoid Type I errors introduced by high zero counts, which is common in ecological data sets, we excluded rare taxa, which were found in under 33% of the grid cells. Note, that this method of exclusion could potentially mean that a taxon with high abundance in a very limited area is excluded from analyses. To further guard against Type I errors, we also used a method of contingency test filtering that removed from consideration a connection between two variables whose joint distribution showed no evidence of deviation from the distribution expected from their combined marginal distributions (chi-squared tests, p  > 0.25)39. This threshold was used to ensure no chance of removing truly dependent dependencies, so that only artefacts such as those found between high zero counts were removed from consideration. These links were provided to the BNIA to exclude from consideration.
    Inference
    Inferring how one node (taxa or physical variable) is likely to change given another node being in a given state is done by calculating the probability of node A being in a given state given node B is in a given state. All nodes that have dependencies between A and B and so are included in the calculation:

    $${mathrm{P}}left( {{mathrm{A|B}}} right) = mathop {sum}limits_{n = 1}^{n = N – 1} {frac{{mathop {prod }nolimits_{s = 1}^{s = S} {mathrm{P}}({mathrm{B}}_n|A_{n + 1})}}{{mathop {sum}nolimits_{m = 1}^N {P(B|A_m)P(A_m)} }}}$$

    The N are the total number of nodes in the chain and n and m are the indices for the chain of nodes of length N connecting the first and end nodes. S are the number of discrete states for each node, which are indexed s. In order to infer the likely change of one taxon’s (A) abundance on another’s abundance (B), the probabilities of all taxa that occur in the network between A and B have to be taken into account. For example, for the probability that B is in a Zero abundance state given A is in a high abundance state, and given that B are connected through a dependency of B on C and C on A, the probability is calculated as follows: the probabilities of C existing in all states is calculated given a High abundance state for A, and then the probabilities for B existing in a Zero state is calculated for each state of C and then summed together to give the probability of B in Zero given A in high. To work out the inferred change in B given a change in A from High to Zero would involve taking the difference between the probabilities for each state of B given A is high with each state of B given A is low. For the code used to generate these inferred probabilities, please see ref. 45.
    Reporting summary
    Further information on research design is available in the Nature Research Reporting Summary linked to this article. More

  • in

    Population genomics in two cave-obligate invertebrates confirms extremely limited dispersal between caves

    1.
    Rétaux, S. & Casane, D. Evolution of eye development in the darkness of caves: adaptation, drift, or both?. EvoDevo 4, 1–12 (2013).
    Article  Google Scholar 
    2.
    Culver, D. C. & Pipan, T. The Biology of Caves and Other Subterranean Habitats (Oxford University Press, Oxford, 2019).
    Google Scholar 

    3.
    Poulson, T. L. & White, W. B. The cave environment. Science (80–) 165, 971–981 (1969).
    ADS  CAS  Article  Google Scholar 

    4.
    Peck, S. B. Evolution of adult morphology and life-history characters in cavernicolous Ptomaphagus beetles. Evolution (N. Y). 40, 1021–1030 (1986).
    Google Scholar 

    5.
    Trontelj, P., Borko, Š & Delić, T. Testing the uniqueness of deep terrestrial life. Sci. Rep. 9, 1–9 (2019).
    CAS  Article  Google Scholar 

    6.
    Polo-Cavia, N. & Gomez-Mestre, I. Pigmentation plasticity enhances crypsis in larval newts: associated metabolic cost and background choice behaviour. Sci. Rep. 7, 1–10 (2017).
    Article  CAS  Google Scholar 

    7.
    Cook, L. M. & Saccheri, I. J. The peppered moth and industrial melanism: evolution of a natural selection case study. Heredity (Edinb.). 110, 207–212 (2013).
    CAS  PubMed  Article  Google Scholar 

    8.
    Stevens, M. & Merilaita, S. Animal camouflage: current issues and new perspectives. Philos. Trans. R. Soc. B. Biol. Sci. 364, 423–427 (2009).
    Article  Google Scholar 

    9.
    Culver, D. C., Master, L. L., Christman, M. C. & Hobbs, H. H. Obligate cave fauna of the 48 contiguous United States. Conserv. Biol. 14, 386–401 (2000).
    Article  Google Scholar 

    10.
    Niemiller, M. L. & Zigler, K. S. Patterns of cave biodiversity and endemism in the Appalachians and Interior Plateau of Tennessee, USA. PLoS ONE 8, e64177 (2013).
    ADS  PubMed  PubMed Central  Article  Google Scholar 

    11.
    Christman, M. C., Culver, D. C., Madden, M. K. & White, D. Patterns of endemism of the eastern North American cave fauna. J. Biogeogr. 32, 1441–1452 (2005).
    Article  Google Scholar 

    12.
    Snowman, C. V., Zigler, K. S. & Hedin, M. Caves as islands: mitochondrial phylogeography of the cave-obligate spider species Nesticus barri (Araneae: Nesticidae). J. Arachnol. 38, 49–56 (2010).
    Article  Google Scholar 

    13.
    Dixon, G. B. & Zigler, K. S. Cave-obligate biodiversity on the campus of Sewanee: The University of the South, Franklin County, Tennessee. Northeast. Nat. 10, 251–266 (2011).
    Google Scholar 

    14.
    Christman, M. C. & Culver, D. C. The relationship between cave biodiversity and available habitat. J. Biogeogr. 28, 367–380 (2001).
    Article  Google Scholar 

    15.
    Zigler, K. S., Niemiller, M. L. & Fenolio, D. B. Cave Biodiversity of the Southern Cumberland Plateau. In 2014 National Speleological Society Convention Guidebook, 159–163 (National Speleological Society, 2014).

    16.
    Hedin, M. C. Speciational history in a diverse clade of habitat-specialized spiders (Araneae: Nesticidae: Nesticus): inferences from geographic-based sampling. Evolution (N. Y.). 51, 1929–1945 (1997).
    Google Scholar 

    17.
    Hedin, M. & Dellinger, B. Descriptions of a new species and previously unknown males of Nesticus (Araneae: Nesticidae) from caves in Eastern North America, with comments on species rarity. Zootaxa 19, 1–19 (2005).
    Article  Google Scholar 

    18.
    Carver, L. M., Perlaky, P., Cressler, A. & Zigler, K. S. Reproductive seasonality in Nesticus (Araneae: Nesticidae) cave spiders. PLoS ONE 11, 7–8 (2016).
    Article  CAS  Google Scholar 

    19.
    Leray, V. L., Caravas, J., Friedrich, M. & Zigler, K. S. Mitochondrial sequence data indicate “Vicariance by Erosion” as a mechanism of species diversification in North American Ptomaphagus (Coleoptera, Leiodidae, Cholevinae ) cave beetles. 57, 35–57 (2019).

    20.
    Wang, S., Meyer, E., McKay, J. K. & Matz, M. V. 2b-RAD: a simple and flexible method for genome-wide genotyping. Nat. Methods 9, 808–810 (2012).
    CAS  PubMed  Article  Google Scholar 

    21.
    Rokas, A. & Abbot, P. Harnessing genomics for evolutionary insights. Trends Ecol. Evol. 24, 192–200 (2009).
    PubMed  Article  Google Scholar 

    22.
    Nunziata, S. O. & Weisrock, D. W. Estimation of contemporary effective population size and population declines using RAD sequence data. Heredity 120, 196–207. https://doi.org/10.1038/s41437-017-0037-y (2018).
    CAS  Article  PubMed  Google Scholar 

    23.
    Ortuño, V. M. et al. The ‘alluvial mesovoid shallow substratum’, a new subterranean habitat. PLoS ONE 8, 1–16 (2013).
    Article  CAS  Google Scholar 

    24.
    Mammola, S. et al. Ecology and sampling techniques of an understudied subterranean habitat: the Milieu Souterrain Superficiel (MSS). Naturwissenschaften 103, 88 (2016).
    PubMed  Article  CAS  Google Scholar 

    25.
    Wakefield, K. R. & Zigler, K. S. Obligate subterranean fauna of Carter State Natural Area, Franklin County, Tennessee. Speleobiol. Notes 4, 24–28 (2012).
    Google Scholar 

    26.
    Dixon, G. B. et al. Genomic determinants of coral heat tolerance across latitudes. Science (80–). 348, 1460–1462 (2015).
    ADS  CAS  Article  Google Scholar 

    27.
    Matz, M. V., Treml, E. A., Aglyamova, G. V. & Bay, L. K. Potential and limits for rapid genetic adaptation to warming in a Great Barrier Reef coral. PLoS Genet. 14, 1–19 (2018).
    Article  CAS  Google Scholar 

    28.
    Matz, M. V. 2bRAD_denovo git repository. https://github.com/z0on/2bRAD_denovo (2019). Accessed 12 September 2020.

    29.
    Li, W. & Godzik, A. Cd-hit: a fast program for clustering and comparing large sets of protein or nucleotide sequences. Bioinformatics 22, 1658–1659 (2006).
    CAS  PubMed  Article  Google Scholar 

    30.
    Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359 (2012).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    31.
    Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    32.
    Nielsen, R., Korneliussen, T., Albrechtsen, A., Li, Y. & Wang, J. SNP calling, genotype calling, and sample allele frequency estimation from new-generation sequencing data. PLoS ONE 7, e37558 (2012).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    33.
    Danecek, P. et al. The variant call format and VCFtools. Bioinformatics 27, 2156–2158 (2011).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    34.
    Korneliussen, T. S. ANGSD web page. https://www.popgen.dk/angsd/index.php/ANGSD (2013). Accessed 12 September 2020.

    35.
    Korneliussen, T. S., Albrechtsen, A. & Nielsen, R. ANGSD: analysis of next generation sequencing data. BMC Bioinf. 15, 1–13 (2014).
    Article  Google Scholar 

    36.
    ANGSD. angsd git repository. https://github.com/ANGSD/angsd (2014). Accessed 12 September 2020.

    37.
    Dixon, G. caveRAD git repository. https://github.com/grovesdixon/caveRAD (2019). Accessed 12 September 2020.

    38.
    Dixon, G., Kitano, J. & Kirkpatrick, M. The origin of a new sex chromosome by introgression between two stickleback fishes. Mol. Biol. Evol. 36, 28–38 (2019).
    CAS  PubMed  Article  Google Scholar 

    39.
    Li, H. A statistical framework for SNP calling, mutation discovery, association mapping and population genetical parameter estimation from sequencing data. Bioinformatics 27, 2987–2993 (2011).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    40.
    Guo, Y. et al. The effect of strand bias in Illumina short-read sequencing data. BMC Genom. 13, 1–11 (2012).
    CAS  Article  Google Scholar 

    41.
    Vieira, F. G., Fumagalli, M., Albrechtsen, A. & Nielsen, R. Estimating inbreeding coefficients from NGS data: impact on genotype calling and allele frequency estimation. Genome Res. 23, 1852–1861 (2013).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    42.
    Fraley, C. & Raftery, A. E. Model-based methods of classification: using the mclust software in chemometrics. J. Stat. Softw. 18, 1–13 (2007).
    Article  Google Scholar 

    43.
    Skotte, L., Korneliussen, T. S. & Albrechtsen, A. Estimating individual admixture proportions from next generation sequencing data. Genetics 195, 693–702 (2013).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    44.
    Fumagalli, M. et al. Quantifying population genetic differentiation from next-generation sequencing data. Genetics 195, 979–992 (2013).
    PubMed  PubMed Central  Article  Google Scholar 

    45.
    Reynolds, J., Weir, B. S. & Cockerham, C. C. Estimation of the coancestry coefficient: basis for a short-term genetic distance. Genet. Soc. Am. 105, 767–779 (1983).
    CAS  Article  Google Scholar 

    46.
    Weir, B. S. & Cockerham, C. C. Estimating F-statistics for the analysis of population structure. Evolution (N. Y.). 38, 1358–1370 (1984).
    CAS  Google Scholar 

    47.
    Hahn, M. W. Molecular Population Genetics (Oxford University Press, Oxford, 2018).
    Google Scholar 

    48.
    Korneliussen, T. S., Moltke, I., Albrechtsen, A. & Nielsen, R. Calculation of Tajima’s D and other neutrality test statistics from low depth next-generation sequencing data. BMC Bioinf. 14, 1–14 (2013).

    49.
    Keightley, P. D., Ness, R. W., Halligan, D. L. & Haddrill, P. R. Estimation of the spontaneous mutation rate per nucleotide site in a Drosophila melanogaster full-sib family. Genetics 196, 313–320 (2014).
    CAS  PubMed  Article  Google Scholar 

    50.
    Wright, S. Variability Within and Among Natural Populations (University of Chicago Press, Chicago, 1978).
    Google Scholar 

    51.
    Kamimura, Y., Abe, J., Ferreira, R. L. & Yoshizawa, K. Microsatellite markers developed using a next-generation sequencing technique for Neotrogla spp. (Psocodea: Prionoglarididae), cave dwelling insects with sex-reversed genitalia. Entomol. Sci. 22, 48–55 (2019).
    Article  Google Scholar 

    52.
    Schäfer, M. A., Orsini, L., McAllister, B. F. & Schlötterer, C. Patterns of microsatellite variation through a transition zone of a chromosomal cline in Drosophila americana. Heredity (Edinb). 97, 291–295 (2006).
    PubMed  Article  CAS  Google Scholar 

    53.
    Bradic, M., Beerli, P., García-De Leán, F. J., Esquivel-Bobadilla, S. & Borowsky, R. L. Gene flow and population structure in the Mexican blind cavefish complex (Astyanax mexicanus). BMC Evol. Biol. 12, 9 (2012).
    PubMed  PubMed Central  Article  Google Scholar 

    54.
    Juan, C., Guzik, M. T., Jaume, D. & Cooper, S. J. B. Evolution in caves: Darwin’s ‘wrecks of ancient life’ in the molecular era. Mol. Ecol. 19, 3865–3880 (2010).
    PubMed  Article  Google Scholar 

    55.
    Pool, J. E., Hellmann, I., Jensen, J. D. & Nielsen, R. Population genetic inference from genomic sequence variation. Genome Res. 20, 291–300 (2010).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    56.
    Silva, M. S., Martins, R. P. & Ferreira, R. L. Cave lithology determining the structure of the invertebrate communities in the Brazilian Atlantic Rain Forest. Biodivers. Conserv. 20, 1713–1729 (2011).
    Article  Google Scholar 

    57.
    Brunet, A. K. & Medellín, R. A. The species-area relationship in bat assemblages of tropical caves. J. Mammal. 82, 1114–1122 (2001).
    Article  Google Scholar 

    58.
    Culver, D. C., Christman, M. C., Elliott, W. R., Hobbs, H. H. & Reddell, J. R. The North American obligate cave fauna: regional patterns. Biodivers. Conserv. 12, 441–468 (2003).
    Article  Google Scholar  More

  • in

    Substrate-dependent fish have shifted less in distribution under climate change

    Our results show that a species’ distribution—measured as change in mean center of biomass and change in northern and southern range extent—depends on the species relationship with oceanographic versus static environmental variables. In the fall, species whose distributions are most influenced by bottom salinity have shifted mean centroids of biomass significantly more than species whose distributions are most influenced by depth or substrate (nonparametric pairwise Wilcoxon test p = 0.00007 and p = 0.0013, Fig. 1a). Species whose distributions are most influenced by bottom temperature have shifted mean centroids of biomass significantly more than species whose distributions are most influenced by depth (nonparametric pairwise Wilcoxon test p = 0.0083, Fig. 1a) and shifted the northern and southern extent of their biomass weighted ranges significantly more than species whose distributions are most influenced by depth and substrate (nonparametric pairwise Wilcoxon test p = 0.00018 and p = 0.014, Fig. 1c and nonparametric pairwise Wilcoxon test p = 0.0038 and p = 0.024, Fig. 1d). Percentage change in biomass weighted range size did not depend on the strongest predictor variable for a species distribution (Fig. 1b). These patterns were not as strong in the spring (see Supplemental Fig. 1).
    Fig. 1: Historic distribution shifts versus the most important predictor variable for a species distribution in the fall.

    Shifts in mean centroid of biomass from first 5 years to last 5 years in kilometers (a), percentage change in biomass range size from first 5 years to last 5 years in meters squared (b). Change in northern (c) and southern (d) extent of the biomass weighted range from first 5 years to last 5 years (n = 93). Brackets and numbers represent p-value. Whiskers represent 1.5* interquartile range. Box represents interquartile range as distance between first and third quartiles. Line represents median, red point represents mean, and black points represent outliers (outside of 1.5*IQR).

    Full size image

    Our results show that in the spring and fall, pelagic species’ distributions are primarily influenced by ocean temperature and depth, while demersal species’ distributions are predominately influenced by ocean temperature and substrate, and benthic species’ distributions are influenced by substrate (Fig. 2).
    Fig. 2: Strongest predictor variable for each species group in the fall.

    Percentage of each species group that had substrate, depth, bottom temperature, and salinity as the strongest predictor variable in terms of deviance explained for the entire time series (n = 93). Results were similar across seasons (see Supplemental Fig. 2).

    Full size image

    Pelagic species have shifted mean center of biomass significantly more over the historic time period compared with benthic species in the fall and significantly more than demersal species in the spring (nonparametric pairwise Wilcoxon test p = 0.039 and p = 0.045, Figs. 3a and 4a). Similarly, pelagic species have expanded their range size significantly more than demersal species and have shifted the northern extent of their range boundary significantly more than demersal species in the spring (nonparametric pairwise Wilcoxon test p = 0.024 and p = 0.028, Fig. 4c, d).
    Fig. 3: Historic distribution shifts for three species types in the fall.

    Fall shifts in southern mean centroid (a), percentage change in range size (b), shifts in northern range boundary (c), and southern range boundary (d) from first 5 years and last 5 years in latitudinal degrees for each species group (n = 93). Brackets and numbers represent p-value. Whiskers represent 1.5* interquartile range. Box represents interquartile range as distance between first and third quartiles. Line represents median, red point represents mean, and black points represent outliers (outside of 1.5*IQR).

    Full size image

    Fig. 4: Historic distribution shifts for three species types in the spring.

    Spring shifts in southern mean centroid (a), percentage change in range size (b), shifts in northern range boundary (c), and southern range boundary (d) from first 5 years and last 5 years in latitudinal degrees for each species group (n = 91). Brackets and numbers represent p-value. Whiskers represent 1.5* interquartile range. Box represents interquartile range as distance between first and third quartiles. Line represents median, red point represents mean, and black points represent outliers (outside of 1.5*IQR).

    Full size image

    These results indicate that benthic species, more influenced by substrate than pelagic species, have retained their historical distributions in response to climate change, while pelagic species have shifted drastically. Exemplars of benthic fish that have retained their distributions include American plaice (Hippoglossoides platessoides), witch flounder (Glyptocephalus cynoglossus), and winter flounder (Pseudopleuronectes americanus) (Fig. 5a). These species distributions are strongly influenced by benthic substrate (Supplemental Data 1 and 2). Exemplars of pelagic fish that have shifted their distributions include rough scad (Trachurus lathami), Atlantic menhaden (Brevoortia tyrannus), and round herring (Etrumeus teres) (Fig. 5b). These species distributions are strongly influenced by bottom temperature and salinity (Supplemental Data 1 and 2).
    Fig. 5: Changes in the biomass weighted mean centroid of species distributions in the Fall.

    Biomass weighted mean centroids were calculated for two time periods: time period 1 (1986–1990) and time period 2 (2014–2018). Benthic species associated with bottom substrate retain their historical distributions (a) whereas pelagic species have shifted distributions (b). Species were selected to show extreme shifts and extreme retentions of distributions, and all shifts can be found in Supplemental Data 1 and 2.

    Full size image

    By linking the historic evidence of species distribution shifts with their relationship with bottom temperature, bottom salinity, and benthic substrate, we have identified a broad generalization on how species with specific life history traits may be influenced by future climate change. Recent work suggests that pelagic species may shift farther under climate change compared with benthic invertebrates14, and recent studies have included sediment type as a constraining variable in projections of future species distributions15. In conjunction with this work, our research highlights that these strong shifts have already occurred historically, and they are influenced by the strength of a pelagic fish species’ relationship with bottom temperature or salinity, compared with benthic species, which are more influenced by substrate. Given the importance of bottom substrate on benthic species distributions, we may see shifts in population dynamics, such as abundance and productivity as temperatures warm instead of geographic shifts in distributions. For, increased ocean warming that may go beyond their preferred thermal envelope is expected in the Northeast LME and Mid Atlantic9. If affinity for substrate type keeps benthic species in regions that become too warm, these species may find themselves in suboptimal conditions and will not be able to relocate as easily as pelagic species.
    Moreover, we examine demersal species that are influenced by both bottom temperature and substrate and have shifted moderately compared to benthic or pelagic species. Research in the North Sea suggests that demersal species are shifting to deeper waters, which may suggest an interaction between bottom temperature and depth that requires further research16. Research in the Northeast LME has linked the shrinking spatial distribution of cusk, a demersal species, to a combination of ocean warming and the ensuing fragmentation of suitable bottom habitat17, suggesting another interaction requiring future examination. As an intermediate case, these species may be the most unpredictable under climate change, and thus fisheries management will have to consider the varying nature of these species’ distribution shifts.
    These results provide historical evidence of pelagic species shifting distributions while benthic species remain more associated with their preferred substrate, which is most likely a result of the functional differences between these types of species. For example, the recruitment success of Atlantic menhaden (Brevoortia tyrannus), a pelagic schooling species we identified to shift historical distributions, is strongly related to ocean temperatures and larger climate dynamics such as the Atlantic Multidecadal Oscillation which influences temperatures and salinity18. The suitable habitat and migration timing of mackerel (Scomber scombrus), a pelagic schooling species, has been linked to changes in ocean temperatures and multidecadal variability19, relying on temperature cues for their seasonal migrations. Research in the North Sea and Baltic Sea suggest that the northward shift of anchovies (Engraulis encrasicolus) and sardines (Sardina pilchardus), two pelagic species, is strongly linked to temperature20. Benthic species, on the other hand, rely on structured biotic habitats, such as marshes, coral reefs, and submerged aquatic vegetation as well as abiotic sediment for their survival21. The importance of abiotic sediment stems from the productivity of these habitats, as they usually contain high levels of detritus, microbes, and microinvertebrates22.
    It is important to identify the limitations of the approaches used in this study. The original species-CPUE data were collected from North Carolina to the Gulf of Maine, meaning we sampled a realized niche of the studied species. In comparison, the fundamental niche of the species may extend beyond our study area, in both the southern and northern directions. This study examined the role of bottom temperature, salinity, depth and substrate on species distributions, but was unable to examine the potential effects of fishing pressure, interspecific interactions, demographic changes, population sizes, or larval dispersal on species distributions and abundance23. While research has demonstrated that simple area-weighted center of distributions can be biased, we attempted to account for this by using several metrics of distribution (area occupied, northern, and southern extents of ranges)24.
    Despite these limitations, we expect our results will be valuable for fisheries managers as they anticipate the likelihood of species distribution shifts in their management areas. While current work has examined the role of temperature in determining species shifts, our work highlights the importance of examining the differential role of other static ecological variables on determining a species likelihood of shifting distributions under climate change. By uncovering the differential effects of certain habitat constraints on pelagic versus demersal and benthic species, we can begin to understand why certain species have shifted dramatically over the last thirty years, while others have retained their historical distributions. These results highlight the need for stock assessments and future species distribution models that include the functional differences among species as well as the environmental variables that constrain their distributions in order to more appropriately understand the potential impacts of climate change on their future distributions. Only then can we understand how future fisheries will be impacted by the ecological effects of climate change. More

  • in

    Long-run bacteria-phage coexistence dynamics under natural habitat conditions in an environmental biotechnology system

    1.
    Bouvier T, del Giorgio PA. Key role of selective viral-induced mortality in determining marine bacterial community composition. Environ Microbiol. 2007;9:287–97.
    CAS  PubMed  Article  Google Scholar 
    2.
    Brown MR, Baptista JC, Lunn M, Swan DL, Smith SJ, Davenport RJ, et al. Coupled virus – bacteria interactions and ecosystem function in an engineered microbial system. Water Res. 2019;152:264–73.
    CAS  PubMed  Article  Google Scholar 

    3.
    Shapiro OH, Kushmaro A, Brenner A. Bacteriophage predation regulates microbial abundance and diversity in a full-scale bioreactor treating industrial wastewater. ISME J. 2010;4:327–36.
    PubMed  Article  Google Scholar 

    4.
    Buckling A, Rainey PB. Antagonistic coevolution between a bacterium and a bacteriophage. Proc R Soc B Biol Sci. 2002;269:931–6.
    Article  Google Scholar 

    5.
    Stern A, Sorek R. The phage-host arms race: shaping the evolution of microbes. BioEssays. 2011;33:43–51.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    6.
    Rostøl JT, Marraffini L. (Ph)ighting phages: how bacteria resist their parasites. Cell Host Microbe. 2019;25:184–94.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    7.
    Paez-Espino D, Sharon I, Morovic W, Stahl B, Thomas BC, Barrangou R, et al. CRISPR immunity drives rapid phage genome evolution in streptococcus thermophilus. MBio. 2015;6:1–9.
    CAS  Article  Google Scholar 

    8.
    Common J, Morley D, Westra ER, Van Houte S. CRISPR-Cas immunity leads to a coevolutionary arms race between Streptococcus thermophilus and lytic phage. Philos Trans R Soc B Biol Sci. 2019;374:20180098.
    CAS  Article  Google Scholar 

    9.
    Flores CO, Valverde S, Weitz JS. Multi-scale structure and geographic drivers of cross-infection within marine bacteria and phages. ISME J. 2013;7:520–32.
    PubMed  Article  Google Scholar 

    10.
    Breitbart M, Rohwer F. Here a virus, there a virus, everywhere the same virus? Trends Microbiol. 2005;13:278–84.
    CAS  PubMed  Article  Google Scholar 

    11.
    Heilmann S, Sneppen K, Krishna S. Coexistence of phage and bacteria on the boundary of self-organized refuges. Proc Natl Acad Sci USA. 2012;109:12828–33.
    CAS  PubMed  Article  Google Scholar 

    12.
    Patterson AG, Jackson SA, Taylor C, Evans GB, Salmond GPC, Przybilski R, et al. Quorum sensing controls adaptive immunity through the regulation of multiple CRISPR-Cas systems. Mol Cell. 2016;64:1102–8.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    13.
    Alseth EO, Pursey E, Luján AM, McLeod I, Rollie C, Westra ER. Bacterial biodiversity drives the evolution of CRISPR-based phage resistance. Nature. 2019;574:549–52.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    14.
    Sousa JAM, de, Rocha EPC. Environmental structure drives resistance to phages and antibiotics during phage therapy and to invading lysogens during colonisation. Sci Rep. 2019;9:1–13.
    Article  CAS  Google Scholar 

    15.
    Scanlan PD. Bacteria–bacteriophage coevolution in the human gut: implications for microbial diversity and functionality. Trends Microbiol. 2017;25:614–23.
    CAS  PubMed  Article  Google Scholar 

    16.
    Dang VT, Sullivan MB. Emerging methods to study bacteriophage infection at the single-cell level. Front Microbiol. 2014;5:724.
    PubMed  PubMed Central  Article  Google Scholar 

    17.
    Edwards RA, McNair K, Faust K, Raes J, Dutilh BE. Computational approaches to predict bacteriophage-host relationships. FEMS Microbiol Rev. 2016;40:258–72.
    CAS  PubMed  Article  Google Scholar 

    18.
    Barrangou R, Fremaux C, Deveau H, Richards M, Boyaval P, Moineau S, et al. CRISPR provides acquired resistance against viruses in prokaryotes. Science. 2007;315:1709–12.
    CAS  PubMed  Article  Google Scholar 

    19.
    McGinn J, Marraffini LA. Molecular mechanisms of CRISPR–Cas spacer acquisition. Nat Rev Microbiol. 2019;17:7–12.
    CAS  PubMed  Article  Google Scholar 

    20.
    Andersson AF, Banfield JF. Virus population dynamics and acquired virus resistance in natural microbial communities. Science. 2008;320:1047–50.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    21.
    Sun CL, Thomas BC, Barrangou R, Banfield JF. Metagenomic reconstructions of bacterial CRISPR loci constrain population histories. ISME J. 2016;10:858–70.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    22.
    Stern A, Mick E, Tirosh I, Sagy O, Sorek R. CRISPR targeting reveals a reservoir of common phages associated with the human gut microbiome. Genome Res. 2012;22:1985–94.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    23.
    Emerson JB, Andrade K, Thomas BC, Norman A, Allen EE, Heidelberg KB, et al. Virus-host and CRISPR dynamics in archaea-dominated hypersaline Lake Tyrrell, Victoria, Australia. Archaea. 2013;2013:370871.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    24.
    Laanto E, Hoikkala V, Ravantti J, Sundberg LR. Long-term genomic coevolution of host-parasite interaction in the natural environment. Nat Commun. 2017;8:111.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    25.
    Held NL, Herrera A, Quiroz HC, Whitaker RJ. CRISPR associated diversity within a population of Sulfolobus islandicus. PLoS ONE. 2010;5:e12988.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    26.
    Tomida J, Morita Y, Shibayama K, Kikuchi K, Sawa T, Akaike T, et al. Diversity and microevolution of CRISPR loci in Helicobacter cinaedi. PLoS ONE. 2017;12:e0186241.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    27.
    Pauly MD, Bautista MA, Black JA, Whitaker RJ. Diversified local CRISPR-Cas immunity to viruses of Sulfolobus islandicus. Philos Trans R Soc B Biol Sci. 2019;374:20180093.
    CAS  Article  Google Scholar 

    28.
    Richter C, Dy RL, McKenzie RE, Watson BNJ, Taylor C, Chang JT, et al. Priming in the Type I-F CRISPR-Cas system triggers strand-independent spacer acquisition, bi-directionally from the primed protospacer. Nucleic Acids Res. 2014;42:8516–26.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    29.
    Watson BNJ, Vercoe RB, Salmond GPC, Westra ER, Staals RHJ, Fineran PC. Type I-F CRISPR-Cas resistance against virulent phages results in abortive infection and provides population-level immunity. Nat Commun. 2019;10:1–8.
    CAS  Article  Google Scholar 

    30.
    Childs LM, England WE, Young MJ, Weitz JS, Whitaker RJ. CRISPR-induced distributed immunity in microbial populations. PLoS ONE. 2014;9:e101710.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    31.
    Common J, Walker-Sünderhauf D, van Houte S, Westra ER. Diversity in CRISPR-based immunity protects susceptible genotypes by restricting phage spread and evolution. J Evol Biol. 2020;33:1097–108.
    CAS  Article  Google Scholar 

    32.
    Payne P, Geyrhofer L, Barton NH, Bollback JP. CRISPR-based herd immunity can limit phage epidemics in bacterial populations. Elife. 2018;7:e32035.
    PubMed  PubMed Central  Article  Google Scholar 

    33.
    Van Houte S, Ekroth AKE, Broniewski JM, Chabas H, Ashby B, Bondy-Denomy J, et al. The diversity-generating benefits of a prokaryotic adaptive immune system. Nature. 2016;532:385–8.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    34.
    Chabas H, Lion S, Nicot A, Meaden S, van Houte S, Moineau S, et al. Evolutionary emergence of infectious diseases in heterogeneous host populations. PLoS Biol. 2018;16:e2006738.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    35.
    Morley D, Broniewski JM, Westra ER, Buckling A, van Houte S. Host diversity limits the evolution of parasite local adaptation. Mol Ecol. 2017;26:1756–63.
    PubMed  Article  PubMed Central  Google Scholar 

    36.
    Westra ER, Van Houte S, Gandon S, Whitaker R. The ecology and evolution of microbial CRISPR-Cas adaptive immune systems. Philos Trans R Soc B Biol Sci. 2019;374:20190101.
    CAS  Article  Google Scholar 

    37.
    Fernández L, Rodríguez A, García P. Phage or foe: an insight into the impact of viral predation on microbial communities. ISME J. 2018;12:1171–9.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    38.
    Pérez MV, Guerrero LD, Orellana E, Figuerola EL, Erijman L. Time Series Genome-Centric Analysis Unveils Bacterial Response to Operational Disturbance in Activated Sludge. mSystems. 2019;4:e00169–19.
    PubMed  PubMed Central  Article  Google Scholar 

    39.
    Goodfellow M, Kumar Y, Maldonado LA. Bergey’s manual of systematics of archaea and bacteria. Hoboken, New Jersey: John Wiley & Sons, Inc; 2015.
    Google Scholar 

    40.
    Drzyzga O. The strengths and weaknesses of Gordonia: a review of an emerging genus with increasing biotechnological potential. Crit Rev Microbiol. 2012;38:300–16.
    CAS  PubMed  Article  Google Scholar 

    41.
    Arenskötter M, Bröker D, Steinbüchel A. Biology of the metabolically diverse genus Gordonia. Appl Environ Microbiol. 2004;70:3195–204.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    42.
    de los Reyes MF, de los Reyes FL, Hernandez M, Raskin L. Quantification of Gordona amarae strains in foaming activated sludge and anaerobic digester systems with oligonucleotide hybridization probes. Appl Environ Microbiol. 1998;64:2503–12.
    Article  Google Scholar 

    43.
    Kragelund C, Remesova Z, Nielsen JL, Thomsen TR, Eales K, Seviour R, et al. Ecophysiology of mycolic acid-containing Actinobacteria (Mycolata) in activated sludge foams. FEMS Microbiol Ecol. 2007;61:174–84.
    CAS  PubMed  Article  Google Scholar 

    44.
    Russell DA, Hatfull GF. PhagesDB: The actinobacteriophage database. Bioinformatics. 2017;33:784–6.
    CAS  PubMed  Article  Google Scholar 

    45.
    Pope WH, Mavrich TN, Garlena RA, Guerrero-Bustamante CA, Jacobs-Sera D, Montgomery MT, et al. Bacteriophages of Gordonia spp. Display a spectrum of diversity and genetic relationships. mBio. 2017;8:e01069–17.
    PubMed  PubMed Central  Article  Google Scholar 

    46.
    Montgomery MT, Guerrero Bustamante CA, Dedrick RM, Jacobs-Sera D, Hatfull GF. Yet more evidence of collusion: a new viral defense system encoded by Gordonia phage CarolAnn. mBio. 2019;10:e02417–18.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    47.
    Rodriguez-R LM, Konstantinidis KT. Bypassing cultivation to identify bacterial species, culture-independent genomic approaches identify credibly distinct clusters, avoid cultivation bias, and provide true insights into microbial species. Microbe. 2014;9:111–7.
    Google Scholar 

    48.
    Quince C, Delmont TO, Raguideau S, Alneberg J, Darling AE, Collins G, et al. DESMAN: a new tool for de novo extraction of strains from metagenomes. Genome Biol. 2017;18:181.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    49.
    Couvin D, Bernheim A, Toffano-Nioche C, Touchon M, Michalik J, Néron B, et al. CRISPRCasFinder, an update of CRISRFinder, includes a portable version, enhanced performance and integrates search for Cas proteins. Nucleic Acids Res. 2018;46:W246–51.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    50.
    Biswas A, Staals RHJ, Morales SE, Fineran PC, Brown CM. CRISPRDetect: a flexible algorithm to define CRISPR arrays. BMC Genomics. 2016;17:356.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    51.
    Schloss PD, Westcott SL, Ryabin T, Hall JR, Hartmann M, Hollister EB, et al. Introducing mothur: open-source, platform-independent, community-supported software for describing and comparing microbial communities. Appl Environ Microbiol. 2009;75:7537–41.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    52.
    Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS, et al. SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J Comput Biol. 2012;19:455–77.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    53.
    Langmead B, Salzberg SL. Fast gapped-read alignment with Bowtie 2. Nat Methods. 2012;9:357–9.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    54.
    Kang DD, Froula J, Egan R, Wang Z. MetaBAT, an efficient tool for accurately reconstructing single genomes from complex microbial communities. PeerJ. 2015;3:e1165.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    55.
    Pope WH, Jacobs-Sera D. Annotation of bacteriophage genome sequences using DNA master: an overview. Methods Mol Biol. 2018;1681:217–29.
    CAS  PubMed  Article  Google Scholar 

    56.
    Crooks GE, Hon G, Chandonia JM, Brenner SE. WebLogo: a sequence logo generator. Genome Res. 2004;14:1188–90.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    57.
    Skennerton CT, Imelfort M, Tyson GW. Crass: identification and reconstruction of CRISPR from unassembled metagenomic data. Nucleic Acids Res. 2013;41:e105.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    58.
    Rho M, Wu YW, Tang H, Doak TG, Ye Y. Diverse CRISPRs evolving in human microbiomes. PLoS Genet. 2012;8:e1002441.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    59.
    Li H. A statistical framework for SNP calling, mutation discovery, association mapping and population genetical parameter estimation from sequencing data. Bioinformatics. 2011;27:2987–93.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    60.
    Keck F, Rimet F, Bouchez A, Franc A. Phylosignal: an R package to measure, test, and explore the phylogenetic signal. Ecol Evol. 2016;6:2774–80.
    PubMed  PubMed Central  Article  Google Scholar 

    61.
    Parks DH, Imelfort M, Skennerton CT, Hugenholtz P, Tyson GW. CheckM: assessing the quality of microbial genomes recovered from isolates, single cells, and metagenomes. Genome Res. 2015;25:1043–55.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    62.
    Bowers RM, Kyrpides NC, Stepanauskas R, Harmon-Smith M, Doud D, Reddy TBK, et al. Minimum information about a single amplified genome (MISAG) and a metagenome-assembled genome (MIMAG) of bacteria and archaea. Nat Biotechnol. 2017;35:725–31.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    63.
    Lei J, Sun Y. Assemble CRISPRs from metagenomic sequencing data. Bioinformatics. 2016;32:i520–8.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    64.
    Lam TJ, Ye Y. Long reads reveal the diversification and dynamics of CRISPR reservoir in microbiomes. BMC Genomics. 2019;20:567.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    65.
    Leenay RT, Maksimchuk KR, Slotkowski RA, Agrawal RN, Gomaa AA, Briner AE, et al. Identifying and visualizing functional PAM diversity across CRISPR-Cas systems. Mol Cell. 2016;62:137–47.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    66.
    Jackson SA, McKenzie RE, Fagerlund RD, Kieper SN, Fineran PC, Brouns SJJ. CRISPR-Cas: adapting to change. Science. 2017;356:eaal5056.
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    67.
    Levy A, Goren MG, Yosef I, Auster O, Manor M, Amitai G, et al. CRISPR adaptation biases explain preference for acquisition of foreign DNA. Nature. 2015;520:505–10.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    68.
    Ellington MJ, Heinz E, Wailan AM, Dorman MJ, de Goffau M, Cain AK, et al. Contrasting patterns of longitudinal population dynamics and antimicrobial resistance mechanisms in two priority bacterial pathogens over 7 years in a single center. Genome Biol. 2019;20:184.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    69.
    Jackson SA, Birkholz N, Malone LM, Fineran PC. Imprecise spacer acquisition generates CRISPR-Cas immune diversity through primed adaptation. Cell Host Microbe. 2019;25:250–60.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    70.
    Deveau H, Barrangou R, Garneau JE, Labonté J, Fremaux C, Boyaval P, et al. Phage response to CRISPR-encoded resistance in Streptococcus thermophilus. J Bacteriol. 2008;190:1390–400.
    CAS  PubMed  Article  Google Scholar 

    71.
    Semenova E, Jore MM, Datsenko KA, Semenova A, Westra ER, Wanner B, et al. Interference by clustered regularly interspaced short palindromic repeat (CRISPR) RNA is governed by a seed sequence. Proc Natl Acad Sci USA. 2011;108:10098–103.
    CAS  PubMed  Article  Google Scholar 

    72.
    Iranzo J, Lobkovsky AE, Wolf YI, Koonin EV. Evolutionary dynamics of the prokaryotic adaptive immunity system CRISPR-Cas in an explicit ecological context. J Bacteriol. 2013;195:3834–44.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    73.
    Lenski RE. Coevolution of bacteria and phage: Are there endless cycles of bacterial defenses and phage counterdefenses? J Theor Biol. 1984;108:319–25.
    CAS  PubMed  Article  Google Scholar 

    74.
    Bull JJ, Christensen KA, Scott C, Jack BR, Crandall CJ, Krone SM. Phage-bacterial dynamics with spatial structure: Self organization around phage sinks can promote increased cell densities. Antibiotics. 2018;7:8.
    PubMed Central  Article  CAS  PubMed  Google Scholar 

    75.
    Hall AR, Scanlan PD, Morgan AD, Buckling A. Host-parasite coevolutionary arms races give way to fluctuating selection. Ecol Lett. 2011;14:635–42.
    PubMed  Article  PubMed Central  Google Scholar 

    76.
    Best A, Ashby B, White A, Bowers R, Buckling A, Koskella B, et al. Host-parasite fluctuating selection in the absence of specificity. Proc R Soc B Biol Sci. 2017;284:20171615.
    Article  Google Scholar 

    77.
    Gómez P, Buckling A. Bacteria-phage antagonistic coevolution in soil. Science. 2011;332:106–9.
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    78.
    Ignacio-Espinoza JC, Ahlgren NA, Fuhrman JA. Long-term stability and Red Queen-like strain dynamics in marine viruses. Nat Microbiol. 2020;5:265–71.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    79.
    Lopez Pascua L, Hall AR, Best A, Morgan AD, Boots M, Buckling A. Higher resources decrease fluctuating selection during host-parasite coevolution. Ecol Lett. 2014;17:1380–8.
    PubMed  PubMed Central  Article  Google Scholar 

    80.
    Hampton HG, Watson BNJ, Fineran PC. The arms race between bacteria and their phage foes. Nature. 2020;577:327–36.
    CAS  PubMed  Article  Google Scholar 

    81.
    Brockhurst MA, Chapman T, King KC, Mank JE, Paterson S, Hurst GDD. Running with the Red Queen: The role of biotic conflicts in evolution. Proc R Soc B Biol Sci. 2014;281:20141382.
    Article  Google Scholar 

    82.
    Vuono DC, Benecke J, Henkel J, Navidi WC, Cath TY, Munakata-Marr J, et al. Disturbance and temporal partitioning of the activated sludge metacommunity. ISME J. 2015;9:425–35.
    CAS  PubMed  Article  Google Scholar 

    83.
    Kuai L, Verstraete W. Ammonium removal by the oxygen-limited autotrophic nitrification- denitrification system. Appl Environ Microbiol. 1998;64:4500–6.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    84.
    Hosseinidoust Z, Tufenkji N, van de Ven TGM. Formation of biofilms under phage predation: considerations concerning a biofilm increase. Biofouling. 2013;29:457–68.
    CAS  PubMed  Article  Google Scholar 

    85.
    Doron S, Melamed S, Ofir G, Leavitt A, Lopatina A, Keren M, et al. Systematic discovery of antiphage defense systems in the microbial pangenome. Science. 2018;359:eaar4120.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    86.
    Winter C, Bouvier T, Weinbauer MG, Thingstad TF. Trade-Offs between competition and defense specialists among unicellular planktonic organisms: the ‘Killing the Winner’ hypothesis revisited. Microbiol Mol Biol Rev. 2010;74:42–57.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    87.
    Tyson GW, Banfield JF. Rapidly evolving CRISPRs implicated in acquired resistance of microorganisms to viruses. Environ Microbiol. 2008;10:200–7.
    CAS  PubMed  Google Scholar 

    88.
    Weinberger AD, Wolf YI, Lobkovsky AE, Gilmore MS, Koonin EV. Viral diversity threshold for adaptive immunity in prokaryotes. mBio. 2012;3:e00456–12.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    89.
    Cohan FM, Perry EB. A systematics for discovering the fundamental units of bacterial diversity. Curr Biol. 2007;17:R373–86.
    CAS  PubMed  Article  Google Scholar 

    90.
    Bendall ML, Stevens SLR, Chan LK, Malfatti S, Schwientek P, Tremblay J, et al. Genome-wide selective sweeps and gene-specific sweeps in natural bacterial populations. ISME J. 2016;10:1589–601.
    PubMed  PubMed Central  Article  Google Scholar 

    91.
    Sloan WT, Lunn M, Woodcock S, Head IM, Nee S, Curtis TP. Quantifying the roles of immigration and chance in shaping prokaryote community structure. Environ Microbiol. 2006;8:732–40.
    PubMed  Article  Google Scholar 

    92.
    Ayarza JM, Erijman L. Balance of neutral and deterministic components in the dynamics of activated sludge floc assembly. Micro Ecol. 2011;61:486–95.
    Article  Google Scholar 

    93.
    de los Reyes FL. Foam in wastewater treatment facilities. In: Handbook of hydrocarbon and lipid microbiology. Berlin Heidelberg: Springer; 2010. pp. 2401–11.

    94.
    Petrovski S, Seviour RJ, Tillett D. Prevention of Gordonia and Nocardia stabilized foam formation by using bacteriophage GTE7. Appl Environ Microbiol. 2011;77:7864–7.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    95.
    Liu M, Gill JJ, Young R, Summer EJ. Bacteriophages of wastewater foaming-associated filamentous Gordonia reduce host levels in raw activated sludge. Sci Rep. 2015;5:13754.
    PubMed  PubMed Central  Article  Google Scholar 

    96.
    Curtis TP, Head IM, Graham DW. Peer reviewed: theoretical ecology for engineering biology. Environ Sci Technol. 2003;37:64A–70A.
    PubMed  Article  Google Scholar 

    97.
    Daims H, Taylor MW, Wagner M. Wastewater treatment: a model system for microbial ecology. Trends Biotechnol. 2006;24:483–9.
    CAS  PubMed  Article  Google Scholar  More

  • in

    Flowering season of vernal herbs is shortened at elevated temperatures with reduced precipitation in early spring

    1.
    Walther, G. et al. Ecological responses to recent climate change. Nature 416, 389–395. https://doi.org/10.1038/416389a (2002).
    ADS  CAS  Article  PubMed  Google Scholar 
    2.
    Pereira, H. M. et al. Scenarios for global biodiversity in the 21st century. Science 330, 1496–1501. https://doi.org/10.1126/science.1196624 (2010).
    ADS  CAS  Article  PubMed  Google Scholar 

    3.
    Carter, S. K., Saenz, D. & Rudolf, V. H. W. Shifts in phenological distributions reshape interaction potential in natural communities. Ecol. Lett. 21, 1143–1151. https://doi.org/10.1111/ele.13081 (2018).
    Article  PubMed  Google Scholar 

    4.
    Kahl, S. M., Lenhard, M. & Joshi, J. Compensatory mechanisms to climate change in the widely distributed species Silene vulgaris. J. Ecol. 107, 1918–1930. https://doi.org/10.1111/1365-2745.13133 (2019).
    Article  Google Scholar 

    5.
    Easterling, D. R. et al. Climate extremes: observations, modeling, and impacts. Science 289, 2068–2074. https://doi.org/10.1126/science.289.5487.2068 (2000).
    ADS  CAS  Article  PubMed  Google Scholar 

    6.
    IPCC. Global Warming of 1.5°C: An IPCC Special Report on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse gas emission pathways, in the context of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty https://www.ipcc.ch/sr15/ (2018).

    7.
    Wolkovich, et al. Warming experiments underpredict plant phenological responses to climate change. Nature 485, 494–497. https://doi.org/10.1038/nature11014 (2012).
    ADS  CAS  Article  PubMed  Google Scholar 

    8.
    Ahammed, G. J., Li, X., Wan, H., Zhou, G. & Cheng, Y. SlWRKY81 reduces drought tolerance by attenuating proline biosynthesis in tomato. Sci. Hortic. 270, 109444. https://doi.org/10.1016/j.scienta.2020.109444 (2020).
    CAS  Article  Google Scholar 

    9.
    Dorji, T. et al. Impacts of climate change on flowering phenology and production in alpine plants: the importance of end of flowering. Agric. Ecosyst. Environ. 291, 106795. https://doi.org/10.1016/j.agee.2019.106795 (2020).
    Article  Google Scholar 

    10.
    Bertin, R. I. Plant phenology and distribution in relation to recent climate change. J. Torrey Bot. Soc. 135, 126–146. https://doi.org/10.3159/07-RP-035R.1 (2008).
    Article  Google Scholar 

    11.
    Lawson, C. R., Vindenes, Y., Bailey, L. & van de Poll, M. Environmental variation and population responses to global change. Ecol. Lett. 18, 724–736. https://doi.org/10.1111/ele.12437 (2015).
    Article  PubMed  Google Scholar 

    12.
    Sherry, R. A. et al. Divergence of reproductive phenology under climate warming. Proc. Nat. Acad. Sci. USA 104, 198–202. https://doi.org/10.1073/pnas.0605642104 (2007).
    ADS  CAS  Article  PubMed  Google Scholar 

    13.
    Nicotra, A. B. et al. Plant phenotypic plasticity in a changing climate. Trends Plant Sci. 15, 684–692. https://doi.org/10.1016/j.tplants.2010.09.008 (2010).
    CAS  Article  PubMed  Google Scholar 

    14.
    Prevéy, J. S. et al. Warming shortens flowering seasons of tundra plant communities. Nat. Ecol. Evol. 3, 45–52. https://doi.org/10.1038/s41559-018-0745-6 (2019).
    Article  PubMed  Google Scholar 

    15.
    Ahammed, G. J., Li, X., Liu, A. & Chen, S. Physiological and defense responses of tea plants to elevated CO2: a review. Front. Plant Sci. 11, 305. https://doi.org/10.3389/fpls.2020.00305 (2020).
    Article  PubMed  PubMed Central  Google Scholar 

    16.
    Fogelström, E. & Ehrlén, J. Phenotypic but not genotypic selection for earlier flowering in a perennial herb. J. Ecol. 107, 2650–2659. https://doi.org/10.1111/1365-2745.13240 (2019).
    Article  Google Scholar 

    17.
    Badeck, F. et al. Responses of spring phenology to climate change. New Phytol. 162, 295–309. https://doi.org/10.1111/j.1469-8137.2004.01059.x (2004).
    Article  Google Scholar 

    18.
    Ehrlén, J., Raabova, J. & Dahlgren, J. P. Flowering schedule in a perennial plant: life-history trade-offs, seed predation, and total offspring fitness. Ecology 96, 2280–2288. https://doi.org/10.1890/14-1860.1 (2015).
    Article  PubMed  Google Scholar 

    19.
    Körner, C. & Basler, D. Phenology under global warming. Science 327, 1461–1462. https://doi.org/10.1126/science.1186473 (2010).
    ADS  Article  PubMed  Google Scholar 

    20.
    Gerst, K. L., Rossington, N. L. & Mazer, S. J. Phenological responsiveness to climate differs among four species of Quercus in North America. J. Ecol. 105, 1610–1622. https://doi.org/10.1111/1365-2745.12774 (2017).
    Article  Google Scholar 

    21.
    Grossiord, C. et al. Precipitation, not air temperature, drives functional responses of trees in semi-arid ecosystems. J. Ecol. 105, 163–175. https://doi.org/10.1111/1365-2745.12662 (2017).
    Article  Google Scholar 

    22.
    Crimmins, T. M., Crimmins, M. A. & Bertelsen, C. D. Onset of summer flowering in a ‘Sky Island’ is driven by monsoon moisture. New Phytol. 191, 468–479. https://doi.org/10.1111/j.1469-8137.2011.03705.x (2011).
    Article  PubMed  Google Scholar 

    23.
    Meng, F. D. et al. Changes in flowering functional group affect responses of community phenological sequences to temperature change. Ecology 98, 734–740. https://doi.org/10.1002/ecy.1685 (2017).
    CAS  Article  PubMed  Google Scholar 

    24.
    Dunne, J. A., Harte, J. & Taylor, K. J. Subalpine meadow flowering phenology responses to climate change: integrating experimental and gradient methods. Ecol. Monogr. 73, 69–86. https://doi.org/10.1890/0012-9615(2003)073[0069:SMFPRT]2.0.CO;2 (2003).
    Article  Google Scholar 

    25.
    Gugger, S., Kesselring, H., Stöcklin, J. & Hamann, E. Lower plasticity exhibited by high- versus mid- elevation species in their phenological responses to manipulated temperature and drought. Annu. Bot. 116, 953–962. https://doi.org/10.1093/aob/mcv155 (2015).
    Article  Google Scholar 

    26.
    Richardson, A. D. et al. Ecosystem warming extends vegetation activity but heightens vulnerability to cold temperatures. Nature 560, 368–371. https://doi.org/10.1038/s41586-018-0399-1 (2018).
    ADS  CAS  Article  PubMed  Google Scholar 

    27.
    Fenner, M. The phenology of growth and reproduction in plants. Perspect. Plant Ecol. 1, 78–91. https://doi.org/10.1078/1433-8319-00053 (1998).
    Article  Google Scholar 

    28
    Lee, H. & Kang, H. Temperature-driven changes of pollinator assemblage and activity of Megaleranthis saniculifolia (Ranunculaceae) at high altitudes on Mt. Sobaeksan, South Korea. J. Ecol. Environ. 42, 31. https://doi.org/10.1186/s41610-018-0092-1 (2018).
    Article  Google Scholar 

    29.
    Yu, H., Luedeling, E. & Xu, J. Winter and spring warming result in delayed spring phenology on the Tibetan Plateau. Proc. Nat. Acad. Sci. USA 107, 22151–22156. https://doi.org/10.1073/pnas.1012490107 (2010).
    ADS  Article  PubMed  Google Scholar 

    30.
    Cook, B. I., Wolkovich, E. M. & Parmesan, C. Divergent responses to spring and winter warming drive community level flowering trends. Proc. Nat. Acad. Sci. USA 109, 9000–9005. https://doi.org/10.1073/pnas.1118364109 (2012).
    ADS  Article  PubMed  Google Scholar 

    31.
    Meier, A. J., Bratton, S. P. & Duffy, D. C. Possible ecological mechanisms for loss of vernal-herb diversity in logged eastern deciduous forests. Ecol. Appl. 5, 935–946. https://doi.org/10.2307/2269344 (1995).
    Article  Google Scholar 

    32.
    Sung, J. et al. Growth environment and vegetation structure of native habitat of Corydalis cornupetala. Korean J. Environ. Ecol. 27, 271–279 (2013).
    Google Scholar 

    33.
    Augspurger, C. K. & Salk, C. F. Constraints of cold and shade on the phenology of spring ephemeral herb species. J. Ecol. 105, 246–254. https://doi.org/10.1111/1365-2745.12651 (2017).
    CAS  Article  Google Scholar 

    34.
    Rizhsky, L. et al. When defense pathways collide: the response of Arabidopsis to a combination of drought and heat stress. Plant Physiol. 134, 1683–1696. https://doi.org/10.1104/pp.103.033431 (2004).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    35.
    Su, Z. et al. Flower development under drought stress: morphological and transcriptomic analyses reveal acute response of long-term acclimation in Arabidopsis. Plant Cell 25, 3785–3807. https://doi.org/10.1105/tpc.113.115428 (2013).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    36.
    Vallales, F., Wright, S. J., Lasso, E., Kitajima, K. & Pearcy, R. W. Plastic phenotypic response to light of 16 congeneric shrubs from a Panamanian rainforest. Ecology 81, 1925–1936. https://doi.org/10.1890/0012-9658(2000)081[1925:PPRTLO]2.0.CO;2 (2000).
    Article  Google Scholar 

    37.
    Valladares, F., Sanchez-Gomez, D. & Zavala, M. A. Quantitative estimation of phenotypic plasticity: bridging the gap between the evolutionary concept and its ecological applications. J. Ecol. 94, 1103–1116. https://doi.org/10.1111/j.1365-2745.2006.01176.x (2006).
    Article  Google Scholar 

    38.
    CaraDonna, P. J., Iler, A. M. & Inouye, D. W. Shifts in flowering phenology reshape a subalpine plant community. Proc. Nat. Acad. Sci. USA 111, 13. https://doi.org/10.1073/pnas.1323073111 (2014).
    CAS  Article  Google Scholar 

    39.
    Forrest, J. & Miller-Rushing, A. J. Toward a synthetic understanding of the role of phenology in ecology and evolution. Philos. Trans. Biol. Sci. 365, 3101–3112. https://doi.org/10.1098/rstb.2010.0145 (2010).
    Article  Google Scholar 

    40.
    Richardson, A. D. et al. Climate change, phenology, and phenological control of vegetation feedbacks to the climate system. Agric. For. Meteorol. 169, 156–173. https://doi.org/10.1016/j.agrformet.2012.09.012 (2013).
    ADS  Article  Google Scholar 

    41.
    Richards, F. J. A flexible growth function for empirical use. J. Exp. Bot. 29, 290–300. https://doi.org/10.1093/jxb/10.2.290 (1959).
    Article  Google Scholar 

    42.
    Barnabás, B., Jäger, K. & Fehér, A. The effect of drought and heat stress on reproductive processes in cereals. Plant Cell Environ. 31, 11–38. https://doi.org/10.1111/j.1365-3040.2007.01727.x (2008).
    CAS  Article  PubMed  Google Scholar 

    43.
    Limousin, J.-M. et al. Morphological and phenological shoot plasticity in a Mediterranean evergreen oak facing long-term increased drought. Oecologia 169, 565–577. https://doi.org/10.1007/s00442-011-2221-8 (2012).
    ADS  Article  PubMed  Google Scholar 

    44.
    Li, X. et al. Exogeneous melatonin improves tea quality under moderate high temperatures by increasing epigallacatechin-3-gallate and theanine biosynthesis in Camellia sinensis L. J. Plant Physiol. 253, 153273. https://doi.org/10.1016/j.jplph.2020.153273 (2020).
    CAS  Article  PubMed  Google Scholar 

    45.
    Wheeler, J. A. et al. The snow and the willows: earlier spring snowmelt reduces performance in the low-lying alpine shrub Salix herbacea. J. Ecol. 104, 1041–1050. https://doi.org/10.1111/1365-2745.12579 (2016).
    CAS  Article  Google Scholar 

    46.
    Llorens, L. & Peñuelas, J. Experimental evidence of future drier and warmer conditions affecting flowering of two co-occurring Mediterranean shrubs. Int. J. Plant Sci. 166, 235–245. https://doi.org/10.1086/427480 (2005).
    Article  Google Scholar 

    47.
    Bernal, M., Estiarte, M. & Penuelas, J. Drought advances spring growth phenology of the Mediterranean shrub Erica multiflora. Plant Biol. 13, 252–257. https://doi.org/10.1111/j.1438-8677.2010.00358.x (2011).
    CAS  Article  PubMed  Google Scholar 

    48.
    Shavrukov, Y. et al. Early flowering as a drought escape mechanism in plants: how can it aid wheat production?. Front. Plant Sci. 8, 1950. https://doi.org/10.3389/fpls.2017.01950 (2017).
    Article  PubMed  PubMed Central  Google Scholar 

    49.
    Sherry, R. A. et al. Changes in duration of reproductive phases and lagged phenological response to experimental climate warming. Plant Ecol. Divers. 4, 23–35. https://doi.org/10.1080/17550874.2011.557669 (2011).
    Article  Google Scholar 

    50.
    Prasad, P. V. V., Pisipati, S. R., Momčilović, I. & Ristic, Z. Independent and combined effects of high temperature and drought stress during grain filling on plant yield and chloroplast EF-Tu expression in spring wheat. J. Agric. Crop Sci. 197(430–441), 2011. https://doi.org/10.1111/j.1439-037X.2011.00477.x (2011).
    CAS  Article  Google Scholar 

    51.
    Zong, J.-M. et al. The AaDREB1 transcription factor from the cold-tolerant plant Adonis amurensis enhances abiotic stress tolerance in transgenic plant. Int. J. Mol. Sci. 17, 611. https://doi.org/10.3390/ijms17040611 (2016).
    ADS  CAS  Article  PubMed Central  Google Scholar 

    52.
    Żuraw, B., Rysiak, K. & Szymczak, G. Ecology and morphology of the flowers of Hepatica nobilisSchreb. (Ranunculaceae). Mod. Phytomorphol. 4, 39–43. https://doi.org/10.5281/zenodo.161177 (2013).
    Article  Google Scholar 

    53.
    Kalliovirta, M., Ryttäri, T. & Heikkinen, R. K. Population structure of a threatened plant, Pulsatilla patens, in boreal forests: modeling relationships to overgrowth and site closure. Biodivers. Conserv. 15, 3095–3108. https://doi.org/10.1007/s10531-005-5403-z (2006).
    Article  Google Scholar 

    54
    Inghe, O. & Tamm, C. O. Survival and flowering of perennial herbs. IV. The behavior of Hepatica nobilis and Sanicula europaea on permanent plots during 1943–1981. Oikos 45, 400–420. https://doi.org/10.2307/3565576 (1985).
    Article  Google Scholar 

    55.
    Lee, T. B. Colored Flora of Korea (Hyangmunsa, Seoul, 2003).
    Google Scholar 

    56.
    Kang, H. & Jang, S. Flowering patterns among angiosperm species in Korea: diversity and constraints. J. Plant Biol. 47, 348–355. https://doi.org/10.1007/BF03030550 (2004).
    Article  Google Scholar 

    57.
    Culley, T. M. Reproductive biology and delayed selfing in Viola pubscens (Violaceae), an understory herb with chasmogamous and cleistogamous flowers. Int. J. Plant Sci. 163, 113–122. https://doi.org/10.1086/324180 (2002).
    Article  Google Scholar 

    58.
    R Core Team. R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, https://www.R-project.org (2017). More