More stories

  • in

    Evidence of spatial genetic structure in a snow leopard population from Gansu, China

    Alexander JS, Zhang C, Shi K, Riordan P (2016) A granular view of a snow leopard population using camera traps in Central China. Biol Conserv 197:27–31
    Google Scholar 
    Aryal A, Brunton D, Ji W, Karmacharya D, McCarthy T, Bencini R et al. (2014) Multipronged strategy including genetic analysis for assessing conservation options for the snow leopard in the central Himalaya. J Mammal 95:871–881
    Google Scholar 
    Atzeni L, Cushman SA, Bai D, Wang J, Chen P, Shi K et al. (2020) Meta-replication, sampling bias, and multi-scale model selection: a case study on snow leopard (Panthera uncia) in western China. Ecol Evol 10:7686–7712PubMed 
    PubMed Central 

    Google Scholar 
    Bai D-F, Chen P-J, Atzeni L, Cering L, Li Q, Shi K (2018) Assessment of habitat suitability of the snow leopard (Panthera uncia) in Qomolangma National Nature Reserve based on MaxEnt modeling. Zool Res 39:373–386PubMed 
    PubMed Central 

    Google Scholar 
    Balkenhol N, Cushman SA, Storfer AT, Waits LP, eds (2016) Landscape genetics: concepts, methods, applications, 1st ed. John Wiley and Sons Ltd. Oxford, UKBauman D, Vleminckx J, Hardy OJ, Drouet T (2018c) Testing and interpreting the shared space-environment fraction in variation partitioning analyses of ecological data. Oikos 128:274–285
    Google Scholar 
    Bauman D, Drouet T, Dray S, Vleminckx J (2018b) Disentangling good from bad practices in the selection of spatial or phylogenetic eigenvectors. Ecography 41:1638–1649
    Google Scholar 
    Bauman D, Drouet T, Fortin M, Dray S (2018a) Optimizing the choice of a spatial weighting matrix in eigenvector-based methods. Ecology 99:2159–2166PubMed 

    Google Scholar 
    Benone NL, Soares BE, Lobato CMC, Seabra LB, Bauman D, Montag LF de A (2020) How modified landscapes filter rare species and modulate the regional pool of ecological traits? HydrobiologiaBlair C, Weigel DE, Lazik M, Keeley AT, Walker FM, Landguth E et al. (2012) A simulation-based evaluation of methods for inferring linear barriers to gene flow. Mol Ecol Resour 12:822–833PubMed 

    Google Scholar 
    Blanchet FG, Legendre P, Borcard D (2008) Forward selection of explanatory variables. Ecology 89:2623–2632PubMed 

    Google Scholar 
    Bothwell HM, Cushman SA, Woolbright SA, Hersch-Green EI, Evans LM, Whitham TG et al. (2017) Conserving threatened riparian ecosystems in the American West: Precipitation gradients and river networks drive genetic connectivity and diversity in a foundation riparian tree (Populus angustifolia). Mol Ecol 26:5114–5132PubMed 

    Google Scholar 
    Breyne P, Mergeay J, Casaer J (2014) Roe deer population structure in a highly fragmented landscape. Eur J Wildl Res 60:909–917
    Google Scholar 
    Bruggeman DJ, Wiegand T, Fernández N (2010) The relative effects of habitat loss and fragmentation on population genetic variation in the red-cockaded woodpecker (Picoides borealis). Mol Ecol 19:3679–3691PubMed 

    Google Scholar 
    Burgess SM, Garrick RC (2020) Regional replication of landscape genetics analyses of the Mississippi slimy salamander, Plethodon mississippi. Landsc Ecol 35:337–351
    Google Scholar 
    Castillo JA, Epps CW, Davis AR, Cushman SA (2014) Landscape effects on gene flow for a climate-sensitive montane species, the American pika. Mol Ecol 23:843–856PubMed 

    Google Scholar 
    Chambers SM (1995) Spatial structure, genetic variation, and the neighborhood adjustment to effective population size. Conserv Biol 9:1312–1315PubMed 

    Google Scholar 
    Charlesworth B (2009) Effective population size and patterns of molecular evolution and variation. Nat Rev Genet 10:195–205CAS 
    PubMed 

    Google Scholar 
    Chybicki IJ, Burczyk J (2009) Simultaneous estimation of null alleles and inbreeding coefficients. J Heredity 100:106–113CAS 

    Google Scholar 
    Cushman SA, Landguth EL (2010) Scale dependent inference in landscape genetics. Landsc Ecol 25:967–979
    Google Scholar 
    Cushman SA, Shirk A, Landguth EL (2012) Separating the effects of habitat area, fragmentation and matrix resistance on genetic differentiation in complex landscapes. Landsc Ecol 27:369–380
    Google Scholar 
    Cushman SA, Shirk AJ, Landguth EL (2013) Landscape genetics and limiting factors. Conserv Genet 14:263–274
    Google Scholar 
    Cushman SA, McRae BH, McGarigal K (2015) Basics of landscape ecology: an introduction to landscapes and population processes for landscape geneticists. In: Balkhenol N, Cushman S, Storfer A, Waits L (Eds) Landscape genetics: concepts, methods, applications. Wiley, Ofxord, UK, p 11–34
    Google Scholar 
    Cushman SA, McKelvey KS, Hayden J, Schwartz MK (2006) Gene flow in complex landscapes: testing multiple hypotheses with causal modeling. Am Naturalist 168:486–499
    Google Scholar 
    Dalongeville A, Andrello M, Mouillot D, Lobreaux S, Fortin M, Lasram F et al. (2018) Geographic isolation and larval dispersal shape seascape genetic patterns differently according to spatial scale. Evol Appl 11:1437–1447CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Dharmarajan G, Beasley JC, Fike JA, Rhodes OE (2014) Effects of landscape, demographic and behavioral factors on kin structure: testing ecological predictions in a mesopredator with high dispersal capability. Anim Conserv 17:225–234
    Google Scholar 
    Do C, Waples RS, Peel D, Macbeth GM, Tillett BJ, Ovenden JR (2014) NeEstimator v2: re-implementation of software for the estimation of contemporary effective population size (Ne) from genetic data. Mol Ecol Resour 14:209–214CAS 
    PubMed 

    Google Scholar 
    Dray S, Dufour A (2007) The ade4 package: implementing the duality diagram for ecologists. J Stat Soft 22:1–20
    Google Scholar 
    Dray S, Legendre P, Peres-Neto PR (2006) Spatial modelling: a comprehensive framework for principal coordinate analysis of neighbour matrices (PCNM). Ecol Model 196:483–493
    Google Scholar 
    Dray S, Bauman D, Blanchet G, Borcard D, Clappe S, Guenard G, et al. (2020) adespatial: multivariate multiscale spatial analysis. R package version 0.3-8. https://CRAN.R-project.org/package=adespatialEvans JS (2020) spatialEco. R package version 1.3-1, https://github.com/jeffreyevans/spatialEcoForester BR, Jones MR, Joost S, Landguth EL, Lasky JR (2016) Detecting spatial genetic signatures of local adaptation in heterogeneous landscapes. Mol Ecol 25:104–120CAS 
    PubMed 

    Google Scholar 
    François O, Durand E (2010) Spatially explicit Bayesian clustering models in population genetics. Mol Ecol Resour 10:773–784PubMed 

    Google Scholar 
    Frankham R (1996) Relationship of genetic variation to population size in wildlife. Conserv Biol 10:1500–1508
    Google Scholar 
    Frankham R (2005) Genetics and extinction. Biol Conserv 126:131–140
    Google Scholar 
    Galpern P, Peres-Neto PR, Polfus J, Manseau M (2014) MEMGENE: Spatial pattern detection in genetic distance data. Methods Ecol Evolution 5:1116–1120
    Google Scholar 
    Galpern P, Manseau M, Hettinga P, Smith K, Wilson P (2012) Allelematch: an R package for identifying unique multilocus genotypes where genotyping error and missing data may be present. Mol Ecol Resour 12:771–778PubMed 

    Google Scholar 
    Guerrero J, Byrne AW, Lavery J, Presho E, Kelly G, Courcier EA et al. (2018) The population and landscape genetics of the European badger (Meles meles) in Ireland. Ecol Evol 8:10233–10246PubMed 
    PubMed Central 

    Google Scholar 
    Guillot G, Leblois R, Coulon A, Frantz AC (2009) Statistical methods in spatial genetics. Mol Ecol 18:4734–4756PubMed 

    Google Scholar 
    Hearn AJ, Cushman SA, Goossens B, Ross J, Macdonald EA, Hunter LTB et al. (2019) Predicting connectivity, population size and genetic diversity of Sunda clouded leopards across Sabah, Borneo. Landsc Ecol 34:275–290
    Google Scholar 
    Hein C, Moniem HEA, Wagner HH (2021) Can we compare effect size of spatial genetic structure between studies and species using moran eigenvector maps? Frontiers. Ecol Evol 9:612718
    Google Scholar 
    Jackson DA (1993) Stopping rules in principal components analysis: a comparison of heuristical and statistical approaches. Ecology 74:2204–2214
    Google Scholar 
    Jackson ND, Fahrig L (2016) Habitat amount, not habitat configuration, best predicts population genetic structure in fragmented landscapes. Landsc Ecol 31:951–968
    Google Scholar 
    Janecka J, Jackson R, Yuquang Z et al. (2008) Population monitoring of snow leopards using noninvasive collection of scat samples: a pilot study. Anim Conserv 11:401–411
    Google Scholar 
    Janecka JE, Janecka JE, Yu-Guang Z, Di-Qiang L, Munkhtsog B, Bayaraa M et al. (2017) Range-wide snow leopard phylogeography supports three subspecies. J Hered 108:597–607PubMed 

    Google Scholar 
    Johansson Ö, Rauset G, Samelius G, McCarthy T, Andrén H, Tumursukh L et al. (2016) Land sharing is essential for snow leopard conservation. Biol Conserv 203:1–7
    Google Scholar 
    Johansson Ö, Koehler G, Rauset G, Samelius G, Andrén H, Mishra C et al. (2018) Sex-specific seasonal variation in puma and snow leopard home range utilization. Ecosphere 9(8):e02371. https://doi.org/10.1002/ecs2.2371.Article 

    Google Scholar 
    Johansson Ö, Ausilio G, Low M, Lkhagvajav P, Weckworth B, Sharma K (2021) The timing of breeding and independence for snow leopard females and their cubs. Mamm Biol 101:173–180
    Google Scholar 
    Jombart T (2008b) adegenet: a R package for the multivariate analysis of genetic markers. Bioinformatics 24:1403–1405CAS 
    PubMed 

    Google Scholar 
    Jombart T, Pontier D, Dufour AB (2009) Genetic markers in the playground of multivariate analysis. Heredity 102:330–341CAS 
    PubMed 

    Google Scholar 
    Jombart T, Devillard S, Dufour A-B, Pontier D (2008a) Revealing cryptic spatial patterns in genetic variability by a new multivariate method. Heredity 101:hdy200834
    Google Scholar 
    Jombart T (2017) An introduction to adegenet 2.1.0. https://github.com/thibautjombart/adegenet/wiki/TutorialsKarmacharya DB, Thapa K, Shrestha R, Dhakal M, Janecka JE (2011) Noninvasive genetic population survey of snow leopards (Panthera uncia) in Kangchenjunga conservation area, Shey Phoksundo National Park and surrounding buffer zones of Nepal. BMC Research Notes 4:516PubMed 
    PubMed Central 

    Google Scholar 
    Kaszta Ż, Cushman SA, Hearn AJ, Burnham D, Macdonald EA, Goossens B et al. (2019) Integrating Sunda clouded leopard (Neofelis diardi) conservation into development and restoration planning in Sabah (Borneo). Biol Conserv 235:63–76
    Google Scholar 
    Kaszta Ż, Cushman SA, Htun S, Naing H, Burnham D, Macdonald DW (2020) Simulating the impact of Belt and Road initiative and other major developments in Myanmar on an ambassador felid, the clouded leopard, Neofelis nebulosa. Landsc Ecol 35:727–746
    Google Scholar 
    Kindt R, Coe R (2005) Tree diversity analysis: a manual and software for common statistical methods for ecological and biodiversity studies. World Agroforestry Centre (ICRAF), Nairobi (Kenya). http://www.worldagroforestry.org/output/tree-diversity-analysisKorablev MP, Poyarkov AD, Karnaukhov AS, Zvychaynaya EYU, Kuksin AN, Malykh SV et al. (1776) Large-scale and fine-grain population structure and genetic diversity of snow leopards (Panthera uncia Schreber, 1776) from the northern and western parts of the range with an emphasis on the Russian population. Conserv Genet 22:397–410
    Google Scholar 
    Kuhn A, Bauman D, Darras H, Aron S (2017) Sex-biased dispersal creates spatial genetic structure in a parthenogenetic ant with a dependent-lineage reproductive system. Heredity 119:207–213CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Landguth E, Cushman S, Schwz M, Mckelvey K, Mury M, Luikart G (2010) Quantifying the lag time to detect barriers in landscape genetics. Mol Ecol 19:4179–4191CAS 
    PubMed 

    Google Scholar 
    Landguth EL, Cushman SA (2010) cdpop: A spatially explicit cost distance population genetics program. Mol Ecol Resour 10:156–161CAS 
    PubMed 

    Google Scholar 
    Landguth EL, Fedy BC, Oyler-Mccance SJ, Garey AL, Emel SL, Mumma M et al. (2012) Effects of sample size, number of markers, and allelic richness on the detection of spatial genetic pattern. Mol Ecol Resour 12:276–284
    Google Scholar 
    Legendre P, Oksanen J, ter Braak CJF (2011) Testing the significance of canonical axes in redundancy analysis. Methods Ecol Evol 2:269–277
    Google Scholar 
    Legendre P, Fortin M-J, Borcard D (2015) Should the Mantel test be used in spatial analysis? Methods Ecol Evol 6:1239–1247
    Google Scholar 
    Li J, Weckworth BV, McCarthy TM, Liang X, Liu Y, Xing R et al. (2020) Defining priorities for global snow leopard conservation landscapes. Biol Conserv 241:108387
    Google Scholar 
    Macdonald EA, Cushman SA, Landguth EL, Hearn AJ, Malhi Y, Macdonald DW (2018) Simulating impacts of rapid forest loss on population size, connectivity and genetic diversity of Sunda clouded leopards (Neofelis diardi) in Borneo. Plos One 13:e0196974PubMed 
    PubMed Central 

    Google Scholar 
    Manel S, Poncet BN, Legendre P, Gugerli F, Holderegger R (2010) Common factors drive adaptive genetic variation at different spatial scales in Arabis alpina. Mol Ecol 19:3824–35PubMed 

    Google Scholar 
    Mateo-Sánchez MC, Balkenhol N, Cushman S, Pérez T, Domínguez A, Saura S (2015) A comparative framework to infer landscape effects on population genetic structure: are habitat suitability models effective in explaining gene flow? Landscape Ecol 30:1405–1420
    Google Scholar 
    McCarthy TM, Fuller TK, Munkhtsog B (2005) Movements and activities of snow leopards in Southwestern Mongolia. Biol Conserv 124:527–537
    Google Scholar 
    McCarthy T, Mallon D, Jackson R, Zahler P, McCarthy K (2017) Panthera uncia. The IUCN red list of threatened species 2017: e.T22732A50664030. https://doi.org/10.2305/IUCN.UK.2017-2.RLTS.T22732A50664030.en. Accessed 29 June 2019Miquel C, Bellemain E, Poillot C, Bessiere J, Durand A, Taberlet P (2006) Quality indexes to assess the reliability of genotypes in studies using noninvasive sampling and multiple-tube approach. Mol Ecol Notes 6:985–988
    Google Scholar 
    Neel MC, McKelvey K, Ryman N, Lloyd MW, Bull RS, Allendorf FW et al. (2013) Estimation of effective population size in continuously distributed populations: there goes the neighborhood. Heredity 111:189–199CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Oksanen J, Blanchet FG, Friendly M, Kindt R, Legendre P, McGlinn D, et al. (2019) vegan: community ecology package. R package version 2.5-6. https://CRAN.R-project.org/package=vegancitatOyler-McCance SJ, Fedy BC, Landguth EL (2013) Sample design effects in landscape genetics. Conserv Genet 14:275–285
    Google Scholar 
    Paradis E (2010) pegas: an R package for population genetics with an integrated–modular approach. Bioinformatics 26:419–420CAS 
    PubMed 

    Google Scholar 
    Patterson N, Price AL, Reich D (2006) Population structure and eigenanalysis. Plos Genet 2:e190PubMed 
    PubMed Central 

    Google Scholar 
    Peakall R, Smouse PE (2006) GENALEX 6: genetic analysis in Excel. Population genetic software for teaching and research. Mol Ecol Notes 6:288–295
    Google Scholar 
    Peakall R, Smouse PE (2012) GenAlEx 6.5: genetic analysis in Excel. Population genetic software for teaching and research-an update. Bioinformatics 28:2537–2539CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Pearson K (1901) On lines and planes of closest fit to systems of points in space. Philos Mag 2:559–572
    Google Scholar 
    Peres-Neto PR, Legendre P, Dray S, Borcard D (2006) Variation partitioning of species data matrices: estimation and comparison of fractions. Ecology 87:2614–2625PubMed 

    Google Scholar 
    Phillips SJ, Anderson RP, Schapire RE (2006) Maximum entropy modeling of species geographic distributions. Ecol Model 190:231–259
    Google Scholar 
    Riordan P, Cushman SA, Mallon D, Shi K, Hughes J (2016) Predicting global population connectivity and targeting conservation action for snow leopard across its range. Ecography 39:419–426
    Google Scholar 
    Rousset F (2008) genepop’007: a complete re-implementation of the genepop software for Windows and Linux. Mol Ecol Resour 8:103–106PubMed 

    Google Scholar 
    Ruiz-Gonzalez A, Cushman SA, Madeira MJ, Randi E, Gómez-Moliner BJ (2015) Isolation by distance, resistance and/or clusters? Lessons learned from a forest-dwelling carnivore inhabiting a heterogeneous landscape. Mol Ecol 24:5110–5129PubMed 

    Google Scholar 
    Savary P, Foltête J, Moal H, Vuidel G, Garnier S (2021) Analysing landscape effects on dispersal networks and gene flow with genetic graphs. Mol Ecol Resour 21:1167–1185PubMed 

    Google Scholar 
    Schwartz MK, McKelvey KS (2008) Why sampling scheme matters: the effect of sampling scheme on landscape genetic results. Conserv Genet 10:441
    Google Scholar 
    Shirk AJ, Cushman SA (2011) sGD: software for estimating spatially explicit indices of genetic diversity. Mol Ecol Resour 11:922–934CAS 
    PubMed 

    Google Scholar 
    Shirk AJ, Cushman SA (2014) Spatially-explicit estimation of Wright’s neighborhood size in continuous populations. Front Ecol Evolution 2:62
    Google Scholar 
    Shirk AJ, Landguth EL, Cushman SA (2020) The effect of gene flow from unsampled demes in landscape genetic analysis. Mol Ecol Resour 21:394–403PubMed 

    Google Scholar 
    Shirk AJ, Wallin DO, Cushman SA, Rice CG, Warheit KI (2010) Inferring landscape effects on gene flow: a new model selection framework. Mol Ecol 19:3603–3619CAS 
    PubMed 

    Google Scholar 
    Shirk AJ, Landguth EL, Cushman SA (2017a) A comparison of regression methods for model selection in individual-based landscape genetic analysis. Mol Ecol Resour 18:55–67Shirk AJ, Landguth EL, Cushman SA (2017b) A comparison of individual-based genetic distance metrics for landscape genetics. Mol Ecol Resour 17:1308–1317Short-Bull RA, Cushman S, Mace R, Chilton T, Kendall K, Landguth E et al. (2011) Why replication is important in landscape genetics: American black bear in the Rocky Mountains. Mol Ecol 20:1092–1107
    Google Scholar 
    Shrestha B, Kindlmann P (2020) Implications of landscape genetics and connectivity of snow leopard in the Nepalese Himalayas for its conservation. Sci Rep 10:19853CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Stekhoven DJ (2013) missForest: nonparametric missing value imputation using random forest. R package version 1.4Storfer A, Murphy MA, Spear SF, Holderegger R, Waits LP (2010) Landscape genetics: where are we now? Mol Ecol 19:3496–3514PubMed 

    Google Scholar 
    Wagner HH, Fortin M-J (2012) A conceptual framework for the spatial analysis of landscape genetic data. Conserv Genet 14:253–261
    Google Scholar 
    Wagner HH, Fortin MJ (2016) Basics of spatial data analysis: linking landscape and genetic data for landscape genetic studies. In: Balkenhol N, Cushman SA, Storfer AT, Waits LP, eds. Landscape genetics: concepts, methods, applications, 1st ed. John Wiley and Sons Ltd. Oxford, UK. pp. 77–98Wahlund S (1928) Composition of populations and correlation appearances viewed in relation to the studies of inheritance. Hereditas 11:65–106
    Google Scholar 
    Wan HY, Cushman SA, Ganey JL (2019) Improving habitat and connectivity model predictions with multi-scale resource selection functions from two geographic areas. Landsc Ecol 34:503–519
    Google Scholar 
    Wasserman TN, Cushman SA, Schwartz MK, Wallin DO (2010) Spatial scaling and multi-model inference in landscape genetics: Martes americana in northern Idaho. Landsc Ecol 25:1601–1612
    Google Scholar 
    Weckworth B (2021) Snow leopard (Panthera uncia) genetics: the knowledge gaps, needs, and implications for conservation. J Indian I Sci 101:279–290
    Google Scholar 
    Wollenberg AL, van den (1977) Redundancy analysis. An alternative for canonical correlation analysis. Psychometrika 42:207–219
    Google Scholar 
    Wright S (1943) Isolation by distance. Genetics 28:114–138CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wright S (1946) Isolation by distance under diverse systems of mating. Genetics 31:39–59CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wu S, Qinye Y, Du Z (2003) Delineation of eco-geographic regional system of China. J Geogr Sci 13:309
    Google Scholar 
    Zeller KA, Jennings MK, Vickers TW, Ernest HB, Cushman SA, Boyce WM (2018) Are all data types and connectivity models created equal? Validating common connectivity approaches with dispersal data. Divers Distrib 24:868–879
    Google Scholar 
    Zhang Y, Hacker C, Zhang Y, Xue Y, Wu L, Dai Y et al. (2019) An analysis of genetic structure of snow leopard populations in Sanjiang-Yuan and Qilianshan National Parks. Acta Theriologica Sin 39:442–449
    Google Scholar  More

  • in

    Root-associated fungal community reflects host spatial co-occurrence patterns in a subtropical forest

    1.Bever JD, Mangan SA, Alexander HM. Maintenance of plant species diversity by pathogens. Annu Rev Ecol Evol Syst. 2015;46:305–25.
    Google Scholar 
    2.Peay KG. The mutualistic niche: mycorrhizal symbiosis and community dynamics. Annu Rev Ecol Evol Syst. 2016;47:143–64.
    Google Scholar 
    3.Tedersoo L, Bahram M, Zobel M. How mycorrhizal associations drive plant population and community biology. Science. 2020;367:eaba1223.CAS 
    PubMed 

    Google Scholar 
    4.Bever JD, Dickie IA, Facelli E, Facelli JM, Klironomos J, Moora M, et al. Rooting theories of plant community ecology in microbial interactions. Trends Ecol Evol. 2010;25:468–78.PubMed 
    PubMed Central 

    Google Scholar 
    5.Bever JD, Platt TG, Morton ER. Microbial population and community dynamics on plant roots and their feedbacks on plant communities. Annu Rev Microbiol. 2012;66:265–83.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    6.Ke PJ, Miki T. Incorporating the soil environment and microbial community into plant competition theory. Front Microbiol. 2015;6:1066.PubMed 
    PubMed Central 

    Google Scholar 
    7.Mangan SA, Schnitzer SA, Herre EA, Mack KM, Valencia MC, Sanchez EI, et al. Negative plant-soil feedback predicts tree-species relative abundance in a tropical forest. Nature. 2010;466:752–5.CAS 
    PubMed 

    Google Scholar 
    8.Bennett JA, Maherali H, Reinhart KO, Lekberg Y, Hart MM, Klironomos J. Plant-soil feedbacks and mycorrhizal type influence temperate forest population dynamics. Science. 2017;355:181–4.CAS 
    PubMed 

    Google Scholar 
    9.Teste FP, Kardol P, Turner BL, Wardle DA, Zemunik G, Renton M, et al. Plant-soil feedback and the maintenance of diversity in Mediterranean-climate shrublands. Science. 2017;355:173–6.CAS 
    PubMed 

    Google Scholar 
    10.Semchenko M, Leff JW, Lozano YM, Saar S, Davison J, Wilkinson A, et al. Fungal diversity regulates plant-soil feedbacks in temperate grassland. Sci Adv. 2018;4:eaau4578.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    11.Chen L, Swenson NG, Ji N, Mi X, Ren H, Guo L, et al. Differential soil fungus accumulation and density dependence of trees in a subtropical forest. Science. 2019;366:124–8.CAS 
    PubMed 

    Google Scholar 
    12.LaManna JA, Walton ML, Turner BL, Myers JA. Negative density dependence is stronger in resource-rich environments and diversifies communities when stronger for common but not rare species. Ecol Lett. 2016;19:657–67.PubMed 

    Google Scholar 
    13.Eppinga MB, Baudena M, Johnson DJ, Jiang J, Mack KM, Strand AE, et al. Frequency-dependent feedback constrains plant community coexistence. Nat Ecol Evol. 2018;2:1403–7.PubMed 

    Google Scholar 
    14.Brundrett MC. Coevolution of roots and mycorrhizas of land plants. New Phytol. 2002;154:275–304.PubMed 

    Google Scholar 
    15.van der Linde S, Suz LM, Orme CDL, Cox F, Andreae H, Asi E, et al. Environment and host as large-scale controls of ectomycorrhizal fungi. Nature. 2018;558:243–8.PubMed 

    Google Scholar 
    16.Schroeder JW, Martin JT, Angulo DF, Razo IAD, Barbosa JM, Perea R, et al. Host plant phylogeny and abundance predict root‐associated fungal community composition and diversity of mutualists and pathogens. J Ecol. 2019;107:1557–66.
    Google Scholar 
    17.Jiang J, Karen CA, Mara B, Maarten BE, James AE, James DB. Pathogens and mutualists as joint drivers of host species coexistence and turnover: implications for plant competition and succession. Am Nat. 2020;195:591–602.
    Google Scholar 
    18.Schroeder JW, Dobson A, Mangan SA, Petticord DF, Herre EA. Mutualist and pathogen traits interact to affect plant community structure in a spatially explicit model. Nat Commun. 2020;11:2204.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    19.Gilbert GS, Webb CO. Phylogenetic signal in plant pathogen‐host range. Proc Natl Acad Sci USA. 2007;104:4979–83.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    20.Liu X, Liang M, Etienne RS, Wang Y, Staehelin C, Yu S. Experimental evidence for a phylogenetic Janzen‐Connell effect in a subtropical forest. Ecol Lett. 2012;15:111–8.PubMed 

    Google Scholar 
    21.Liang M, Liu X, Etienne RS, Huang F, Wang Y, Yu S. Arbuscular mycorrhizal fungi counteract the Janzen‐Connell effect of soil pathogens. Ecology. 2015;96:562–74.PubMed 

    Google Scholar 
    22.Benítez MS, Hersh MH, Vilgalys R, Clark JS. Pathogen regulation of plant diversity via effective specialization. Trends Ecol Evol. 2013;28:705–11.PubMed 

    Google Scholar 
    23.Klironomos J, Zobel M, Tibbett M. Forces that structure plant communities: quantifying the importance of the mycorrhizal symbiosis. New Phytol. 2011;189:366–70.PubMed 

    Google Scholar 
    24.van der Heijden MGA, Bardgett RD, van Straalen NM. The unseen majority: soil microbes as drivers of plant diversity and productivity in terrestrial ecosystems. Eco Lett. 2008;11:296–310.
    Google Scholar 
    25.Wiegand T, Moloney KA. Rings, circles, and null‐models for point pattern analysis in ecology. Oikos. 2004;104:209–29.
    Google Scholar 
    26.Perry GL, Miller BP, Enright NJ. A comparison of methods for the statistical analysis of spatial point patterns in plant ecology. Plant Ecol. 2006;187:59–82.
    Google Scholar 
    27.Law R, Illian J, Burslem DF, Gratzer G, Gunatilleke CV, Gunatilleke IA. Ecological information from spatial patterns of plants: insights from point process theory. J Ecol. 2009;97:616–28.
    Google Scholar 
    28.Liang M, Liu X, Parker IM, Johnson D, Zheng Y, Luo S, et al. Soil microbes drive phylogenetic diversity-productivity relationships in a subtropical forest. Sci Adv. 2019;5:eaax5088.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    29.Chen Y, Jia P, Cadotte MW, Wang P, Liu X, Qi Y, et al. Rare and phylogenetically distinct plant species exhibit less diverse root-associated pathogen communities. J Ecol. 2019;107:1226–37.
    Google Scholar 
    30.Peters HA. Neighbour‐regulated mortality: the influence of positive and negative density dependence on tree populations in species‐rich tropical forests. Ecol Lett. 2003;6:757–65.
    Google Scholar 
    31.Cutler DR, Edwards TC Jr, Beard KH, Cutler A, Hess KT, Gibson J, et al. Random forests for classification in ecology. Ecology. 2007;88:2783–92.PubMed 

    Google Scholar 
    32.Kattge J, Diaz S, Lavorel S, Prentice IC, Leadley P, Bönisch G, et al. TRY – a global database of plant traits. Glob Chang Biol. 2011;17:2905–35.PubMed Central 

    Google Scholar 
    33.Davey ML, Heegaard E, Halvorsen R, Ohlson M, Kauserud H. Seasonal trends in the biomass and structure of bryophyte-associated fungal communities explored by 454 pyrosequencing. New Phytol. 2012;195:844–56.CAS 
    PubMed 

    Google Scholar 
    34.Nguyen NH, Song Z, Bates ST, Branco S, Tedersoo L, Menke J, et al. FUNGuild: an open annotation tool for parsing fungal community datasets by ecological guild. Fungal Ecol. 2016;20:241–8.
    Google Scholar 
    35.Leff JW, Bardgett RD, Wilkinson A, Jackson BG, Pritchard WJ, Jonathan R, et al. Predicting the structure of soil communities from plant community taxonomy, phylogeny, and traits. ISME J. 2018;12:1794–805.PubMed 
    PubMed Central 

    Google Scholar 
    36.Wang Z, Jiang Y, Deane DC, He F, Shu W, Liu Y. Effects of host phylogeny, habitat and spatial proximity on host specificity and diversity of pathogenic and mycorrhizal fungi in a subtropical forest. New Phytol. 2019;223:462–74.PubMed 

    Google Scholar 
    37.Zhao Z, Li X, Liu MF, Merckx VS, Saunders RM, Zhang D. Specificity of assemblage, not fungal partner species, explains mycorrhizal partnerships of mycoheterotrophic Burmannia plants. ISME J. 2021;15:1614–27.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    38.Peay KG, Baraloto C, Fine PV. Strong coupling of plant and fungal community structure across western Amazonian rainforests. ISME J. 2013;7:1852–61.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    39.Barberán A, McGuire KL, Wolf JA, Jones FA, Wright SJ, Turner BL, et al. Relating belowground microbial composition to the taxonomic, phylogenetic, and functional trait distributions of trees in a tropical forest. Ecol Lett. 2015;18:1397–405.PubMed 

    Google Scholar 
    40.LaManna JA, Belote RT, Burkle LA, Catano CP, Myers JA. Negative density dependence mediates biodiversity-productivity relationships across scales. Nat Ecol Evol. 2017;1:1107–15.PubMed 

    Google Scholar 
    41.Peh KS, Lewis SL, Lloyd J. Mechanisms of monodominance in diverse tropical tree‐dominated systems. J Ecol. 2011;99:891–8.
    Google Scholar 
    42.Johnson DJ, Clay K, Phillips RP. Mycorrhizal associations and the spatial structure of an old-growth forest community. Oecologia. 2018;186:195–204.PubMed 

    Google Scholar 
    43.Waud M, Busschaert P, Lievens B, Jacquemyn H. Specificity and localised distribution of mycorrhizal fungi in the soil may contribute to co-existence of orchid species. Fungal Ecol. 2016;20:155–65.
    Google Scholar 
    44.Põlme S, Bahram M, Jacquemyn H, Kennedy P, Kohout P, Moora M, et al. Host preference and network properties in biotrophic plant–fungal associations. New Phytol. 2018;217:1230–9.PubMed 

    Google Scholar 
    45.Simard SW, Beiler KJ, Bingham MA, Deslippe JR, Philip LJ, Teste FP. Mycorrhizal networks: mechanisms, ecology and modelling. Fungal Biol Rev. 2012;26:39–60.
    Google Scholar 
    46.Bever JD, Westover KM, Antonovics J. Incorporating the soil community into plant population dynamics: the utility of the feedback approach. J Ecol. 1997;85:561–73.
    Google Scholar 
    47.Bardgett RD, Wardle DA. Aboveground-belowground linkages: biotic interactions, ecosystem processes, and global change. New York: Oxford University Press; 2010.48.Kandlikar GS, Johnson CA, Yan X, Kraft NJ, Levine JM. Winning and losing with microbes: how microbially mediated fitness differences influence plant diversity. Ecol Lett. 2019;22:1178–91.PubMed 

    Google Scholar 
    49.Swenson NG, Iida Y, Howe R, Wolf A, Umaña MN, Petprakob K, et al. Tree co-occurrence and transcriptomic response to drought. Nat Commun. 2017;8:1996.PubMed 
    PubMed Central 

    Google Scholar 
    50.Řezáčová V, Gryndler M, Bukovská P, Šmilauer P, Jansa J. Molecular community analysis of arbuscular mycorrhizal fungi—contributions of PCR primer and host plant selectivity to the detected community profiles. Pedobiologia. 2016;59:179–87.
    Google Scholar 
    51.Hart MM, Reader RJ, Klironomos JN. Plant coexistence mediated by arbuscular mycorrhizal fungi. Trends Ecol Evol. 2003;18:418–23.
    Google Scholar 
    52.Taylor DL, Walters WA, Lennon NJ, Bochicchio J, Krohn A, Caporaso JG, et al. Accurate estimation of fungal diversity and abundance through improved lineage-specific primers optimized for Illumina amplicon sequencing. Appl Environ Microbiol. 2016;82:7217–26.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    53.Lekberg Y, Vasar M, Bullington LS, Sepp SK, Antunes PM, Bunn R, et al. More bang for the buck? Can arbuscular mycorrhizal fungal communities be characterized adequately alongside other fungi using general fungal primers? New Phytol. 2018;220:971–6.PubMed 

    Google Scholar 
    54.Egan CP, Rummel A, Kokkoris V, Klironomos J, Lekberg Y, Hart MM. Using mock communities of arbuscular mycorrhizal fungi to evaluate fidelity associated with Illumina sequencing. Fungal Ecol. 2018;33:52–64.
    Google Scholar  More

  • in

    Viral community analysis in a marine oxygen minimum zone indicates increased potential for viral manipulation of microbial physiological state

    1.Fuhrman JA. Marine viruses and their biogeochemical and ecological effects. Nature. 1999;399:541–8.CAS 
    PubMed 

    Google Scholar 
    2.Suttle CA. Viruses in the sea. Nature. 2005;437:356–61.CAS 
    PubMed 

    Google Scholar 
    3.Breitbart M. Marine viruses: truth or dare. Ann Rev Mar Sci. 2012;4:425–48.PubMed 

    Google Scholar 
    4.Brum JR, Sullivan MB. Rising to the challenge: accelerated pace of discovery transforms marine virology. Nat Rev Microbiol. 2015;13:147–59.5.Breitbart M, Thompson L, Suttle C, Sullivan M. Exploring the vast diversity of marine viruses. Oceanography. 2007;20:135–9.
    Google Scholar 
    6.Puxty RJ, Millard AD, Evans DJ, Scanlan DJ. Shedding new light on viral photosynthesis. Photosynth Res. 2015;126:71–97.7.Sharon I, Tzahor S, Williamson S, Shmoish M, Man-Aharonovich D, Rusch DB, et al. Viral photosynthetic reaction center genes and transcripts in the marine environment. ISME J. 2007;1:492–501.CAS 
    PubMed 

    Google Scholar 
    8.Roux S, Hawley AK, Torres Beltran M, Scofield M, Schwientek P, Stepanauskas R, et al. Ecology and evolution of viruses infecting uncultivated SUP05 bacteria as revealed by single-cell- and meta-genomics. Elife. 2014;3:e03125.9.Trubl G, Jang H Bin, Roux S, Emerson JB, Solonenko N, Vik DR, et al. Soil viruses are underexplored players in ecosystem carbon processing. mSystems. 2018;3:e00076–18.10.Roux S, Brum JR, Dutilh BE, Sunagawa S, Duhaime MB, Loy A, et al. Ecogenomics and potential biogeochemical impacts of globally abundant ocean viruses. Nature. 2016;537:689–93.CAS 
    PubMed 

    Google Scholar 
    11.Hurwitz BL, Brum JR, Sullivan MB. Depth-stratified functional and taxonomic niche specialization in the ‘core’ and ‘flexible’ Pacific Ocean Virome. ISME J. 2015;9:472–84.CAS 
    PubMed 

    Google Scholar 
    12.Gazitúa MC, Vik DR, Roux S, Gregory AC, Bolduc B, Widner B, et al. Potential virus-mediated nitrogen cycling in oxygen-depleted oceanic waters. ISME J. 2021;15:981–98.PubMed 

    Google Scholar 
    13.Brum JR, Ignacio-Espinoza JC, Roux S, Doulcier G, Acinas SG, Alberti A, et al. Ocean plankton. Patterns and ecological drivers of ocean viral communities. Science. 2015;348:1261498.PubMed 

    Google Scholar 
    14.Cassman N, Prieto-Davó A, Walsh K, Silva GGZ, Angly F, Akhter S, et al. Oxygen minimum zones harbour novel viral communities with low diversity. Environ Microbiol. 2012;14:3043–65.CAS 
    PubMed 

    Google Scholar 
    15.Vik D, Gazitúa MC, Sun CL, Zayed AA, Aldunate M, Mulholland MR, et al. Genome-resolved viral ecology in a marine oxygen minimum zone. Environ Microbiol. 2021;23:2858–74.16.Mara P, Vik D, Pachiadaki MG, Suter EA, Poulos B, Taylor GT, et al. Viral elements and their potential influence on microbial processes along the permanently stratified Cariaco Basin redoxcline. ISME J. 2020;14:3079–92.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    17.Tiano L, Garcia-Robledo E, Dalsgaard T, Devol AH, Ward BB, Ulloa O, et al. Oxygen distribution and aerobic respiration in the north and south eastern tropical Pacific oxygen minimum zones. Deep Res Part I Oceanogr Res Pap. 2014;94:173–83.CAS 

    Google Scholar 
    18.Schmidtko S, Stramma L, Visbeck M. Decline in global oceanic oxygen content during the past five decades. Nature. 2017;542:335–9.CAS 
    PubMed 

    Google Scholar 
    19.Paulmier A, Ruiz-Pino D. Oxygen minimum zones (OMZs) in the modern ocean. Prog Oceanogr. 2009;80:113–28.
    Google Scholar 
    20.Wright JJ, Konwar KM, Hallam SJ. Microbial ecology of expanding oxygen minimum zones. Nat Rev Microbiol. 2012;10:381–94.CAS 
    PubMed 

    Google Scholar 
    21.Bertagnolli AD, Stewart FJ. Microbial niches in marine oxygen minimum zones. Nat Rev Microbiol. 2018;1:723–9.22.Codispoti LA, Friedrich GE, Packard TT, Glover HE, Kelly PJ, Spinrad RW, et al. High nitrite levels off northern Peru: a signal of instability in the marine denitrification rate. Science. 1986;233:1200 LP–1202.
    Google Scholar 
    23.Canfield DE, Stewart FJ, Thamdrup B, De Brabandere L, Dalsgaard T, Delong EF, et al. A cryptic sulfur cycle in oxygen-minimum-zone waters off the Chilean coast. Science. 2010;330:1375–8.CAS 
    PubMed 

    Google Scholar 
    24.Hurwitz BL, Westveld AH, Brum JR, Sullivan MB. Modeling ecological drivers in marine viral communities using comparative metagenomics and network analyses. Proc Natl Acad Sci USA. 2014;111:10714 LP–10719.
    Google Scholar 
    25.Garcia-Robledo E, Padilla CC, Aldunate M, Stewart FJ, Ulloa O, Paulmier A, et al. Cryptic oxygen cycling in anoxic marine zones. Proc Natl Acad Sci USA. 2017;114:8319–24.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    26.Lavin P, González B, Santibáñez JF, Scanlan DJ, Ulloa O. Novel lineages of Prochlorococcus thrive within the oxygen minimum zone of the eastern tropical South Pacific. Environ Microbiol Rep. 2010;2:728–38.CAS 
    PubMed 

    Google Scholar 
    27.Ulloa O, Canfield DE, DeLong EF, Letelier RM, Stewart FJ. Microbial oceanography of anoxic oxygen minimum zones. Proc Natl Acad Sci USA. 2012;109:15996–6003.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    28.Bettarel Y, Sime-Ngando T, Amblard C, Dolan J. Viral activity in two contrasting lake ecosystems. Appl Environ Microbiol. 2004;70:2941–51.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    29.Weinbauer MG, Brettar I, Höfle MG. Lysogeny and virus-induced mortality of bacterioplankton in surface, deep, and anoxic marine waters. Limnol Oceanogr. 2003;48:1457–65.
    Google Scholar 
    30.Heldal M, Bratbak G. Production and decay of viruses in aquatic environments. Mar Ecol Prog Ser. 1991;72:205–12.
    Google Scholar 
    31.Proctor LM, Fuhrman JA. Viral mortality of marine bacteria and cyanobacteria. Nature. 1990;343:60–62.
    Google Scholar 
    32.Brum JR, Morris J, Décima M, Stukel M. Mortality in the oceans: causes and consequences. In Eco-DAS IX Symposium Proceedings. Association for the Sciences of Limnology and Oceanography; 2014.33.Colombet J, Sime-Ngando T. Seasonal depth-related gradients in virioplankton: lytic activity and comparison with protistan grazing potential in Lake Pavin (France). Micro Ecol. 2012;64:67–78.
    Google Scholar 
    34.Colombet J, Sime-Ngando T, Cauchie HM, Fonty G, Hoffmann L, Demeure G. Depth-related gradients of viral activity in Lake Pavin. Appl Environ Microbiol. 2006;72:4440–5.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    35.Brum J, Steward G, Jiang S, Jellison R. Spatial and temporal variability of prokaryotes, viruses, and viral infections of prokaryotes in an alkaline, hypersaline lake. Aquat Micro Ecol. 2005;41:247–60.
    Google Scholar 
    36.Brum JR, Schenck RO, Sullivan MB. Global morphological analysis of marine viruses shows minimal regional variation and dominance of non-tailed viruses. ISME J. 2013;7:1738–51.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    37.Kauffman KM, Hussain FA, Yang J, Arevalo P, Brown JM, Chang WK, et al. A major lineage of non-tailed dsDNA viruses as unrecognized killers of marine bacteria. Nature. 2018;554:118–22.CAS 
    PubMed 

    Google Scholar 
    38.Székely AJ, Breitbart M. Single-stranded DNA phages: from early molecular biology tools to recent revolutions in environmental microbiology. FEMS Microbiol Lett. 2016;363:27.
    Google Scholar 
    39.Roux S, Enault F, Hurwitz BL, Sullivan MB. VirSorter: Mining viral signal from microbial genomic data. PeerJ. 2015;2015:e985.
    Google Scholar 
    40.Hurwitz BL, Sullivan MB. The Pacific Ocean Virome (POV): a marine viral metagenomic dataset and associated protein clusters for quantitative viral ecology. PLoS One. 2013;8:e57355.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    41.Aldunate M, Henríquez-Castillo C, Ji Q, Lueders-Dumont J, Mulholland MR, Ward BB, et al. Nitrogen assimilation in picocyanobacteria inhabiting the oxygen-deficient waters of the eastern tropical North and South Pacific. Limnol Oceanogr. 2020;65:437–53.CAS 

    Google Scholar 
    42.Solonenko SA, Ignacio-Espinoza JC, Alberti A, Cruaud C, Hallam S, Konstantinidis K, et al. Sequencing platform and library preparation choices impact viral metagenomes. BMC Genom. 2013;14:320.CAS 

    Google Scholar 
    43.Duhaime MB, Deng L, Poulos BT, Sullivan MB. Towards quantitative metagenomics of wild viruses and other ultra-low concentration DNA samples: a rigorous assessment and optimization of the linker amplification method. Environ Microbiol. 2012;14:2526–37.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    44.Ganesh S, Bristow LA, Larsen M, Sarode N, Thamdrup B, Stewart FJ. Size-fraction partitioning of community gene transcription and nitrogen metabolism in a marine oxygen minimum zone. ISME J. 2015;9:2682–96.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    45.Allen LZ, Allen EE, Badger JH, McCrow JP, Paulsen IT, Elbourne LD, et al. Influence of nutrients and currents on the genomic composition of microbes across an upwelling mosaic. ISME J. 2012;6:1403–14.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    46.Hurwitz BL, U’Ren JM. Viral metabolic reprogramming in marine ecosystems. Curr Opin Microbiol. 2016;31:161–8.47.Breitbart M, Bonnain C, Malki K, Sawaya NA. Phage puppet masters of the marine microbial realm. Nat Microbiol. 2018;3:754–66.CAS 
    PubMed 

    Google Scholar 
    48.Ignacio-Espinoza JC, Sullivan MB. Phylogenomics of T4 cyanophages: lateral gene transfer in the ‘core’ and origins of host genes. Environ Microbiol. 2012;14:2113–26.CAS 
    PubMed 

    Google Scholar 
    49.Crummett LT, Puxty RJ, Weihe C, Marston MF, Martiny JBH. The genomic content and context of auxiliary metabolic genes in marine cyanomyoviruses. Virology. 2016;499:219–29.CAS 
    PubMed 

    Google Scholar 
    50.Sullivan MB, Lindell D, Lee JA, Thompson LR, Bielawski JP, Chisholm SW. Prevalence and evolution of core photosystem II genes in marine cyanobacterial viruses and their hosts. PLoS Biol. 2006;4:e234.PubMed 
    PubMed Central 

    Google Scholar 
    51.Bragg JG, Chisholm SW. Modeling the fitness consequences of a cyanophage-encoded photosynthesis gene. PLoS One. 2008;3:e3550.PubMed 
    PubMed Central 

    Google Scholar 
    52.Puxty RJ, Evans DJ, Millard AD, Scanlan DJ. Energy limitation of cyanophage development: implications for marine carbon cycling. ISME J. 2018;12:1273–86.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    53.White AE, Foster RA, Benitez-Nelson CR, Masqué P, Verdeny E, Popp BN, et al. Nitrogen fixation in the Gulf of California and the Eastern Tropical North Pacific. Prog Oceanogr. 2013;109:1–17.
    Google Scholar 
    54.Jayakumar A, Chang BX, Widner B, Bernhardt P, Mulholland MR, Ward BB. Biological nitrogen fixation in the oxygen-minimum region of the eastern tropical North Pacific ocean. ISME J. 2017;11:2356–67.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    55.Fuchsman CA, Devol AH, Saunders JK, McKay C, Rocap G. Niche partitioning of the N cycling microbial community of an offshore oxygen deficient zone. Front Microbiol. 2017;8:2384.PubMed 
    PubMed Central 

    Google Scholar 
    56.Zhang Y, Pohlmann EL, Halbleib CM, Ludden PW, Roberts GP. Effect of P(II) and its homolog GlnK on reversible ADP-ribosylation of dinitrogenase reductase by heterologous expression of the Rhodospirillum rubrum dinitrogenase reductase ADP-ribosyl transferase-dinitrogenase reductase-activating glycohydrolase regula. J Bacteriol. 2001;183:1610–20.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    57.Tong W-H. Distinct iron-sulfur cluster assembly complexes exist in the cytosol and mitochondria of human cells. EMBO J. 2000;19:5692–5700.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    58.Py B, Barras F. Building Feg-S proteins: bacterial strategies. Nat Rev Microbiol. 2010;8:436–46.59.Eichhorn E, van der Ploeg JR, Kertesz MA, Leisinger T. Characterization of alpha-ketoglutarate-dependent taurine dioxygenase from Escherichia coli. J Biol Chem. 1997;272:23031–6.CAS 
    PubMed 

    Google Scholar 
    60.Friedrich CG, Bardischewsky F, Rother D, Quentmeier A, Fischer J. Prokaryotic sulfur oxidation. Curr Opin Microbiol. 2005;8:253–9.61.Anantharaman K, Duhaime MB, Breier JA, Wendt KA, Toner BM, Dick GJ. Sulfur oxidation genes in diverse deep-sea viruses. Science. 2014;344:757 LP–760.
    Google Scholar 
    62.Callbeck CM, Lavik G, Ferdelman TG, Fuchs B, Gruber-Vodicka HR, Hach PF, et al. Oxygen minimum zone cryptic sulfur cycling sustained by offshore transport of key sulfur oxidizing bacteria. Nat Commun. 2018;9:1729.PubMed 
    PubMed Central 

    Google Scholar 
    63.Carolan MT, Smith JM, Beman JM. Transcriptomic evidence for microbial sulfur cycling in the eastern tropical North Pacific oxygen minimum zone. Front Microbiol. 2015;6:334.PubMed 
    PubMed Central 

    Google Scholar 
    64.Ganesh S, Bertagnolli AD, Bristow LA, Padilla CC, Blackwood N, Aldunate M, et al. Single cell genomic and transcriptomic evidence for the use of alternative nitrogen substrates by anammox bacteria. ISME J. 2018;1:2706–22.65.Howard-Varona C, Hargreaves KR, Abedon ST, Sullivan MB. Lysogeny in nature: mechanisms, impact and ecology of temperate phages. ISME J. 2017;11:1511–20.PubMed 
    PubMed Central 

    Google Scholar 
    66.Lill R, Dutkiewicz R, Elsässer HP, Hausmann A, Netz DJA, Pierik AJ, et al. Mechanisms of iron-sulfur protein maturation in mitochondria, cytosol and nucleus of eukaryotes. Biochim Biophys Acta Mol Cell Res. 2006;1763:652–67.67.Fontecave M. Iron-sulfur clusters: ever-expanding roles. Nat Chem Biol. 2006;2:171–4.CAS 
    PubMed 

    Google Scholar 
    68.Roche B, Aussel L, Ezraty B, Mandin P, Py B, Barras F. Iron/sulfur proteins biogenesis in prokaryotes: formation, regulation and diversity. Biochim Biophys Acta Bioenerg. 2013;1827:455–69.CAS 

    Google Scholar 
    69.Xu XM, Møller SG. Iron-sulfur clusters: Biogenesis, molecular mechanisms, and their functional significance. Antioxid Redox Signal. 2011;15:271–307.PubMed 

    Google Scholar 
    70.Miller HK, Auerbuch V. Bacterial iron-sulfur cluster sensors in mammalian pathogens. Metallomics. 2015;7:943–56.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    71.Sharon I, Battchikova N, Aro E-M, Giglione C, Meinnel T, Glaser F, et al. Comparative metagenomics of microbial traits within oceanic viral communities. ISME J. 2011;5:1178–90.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    72.Loiseau L, Ollagnier-de-Choudens S, Nachin L, Fontecave M, Barras F. Biogenesis of Fe-S cluster by the bacterial suf system. SufS and SufE form a new type of cysteine desulfurase. J Biol Chem. 2003;278:38352–9.CAS 
    PubMed 

    Google Scholar 
    73.Outten FW, Wood MJ, Muñoz FM, Storz G. The SufE protein and the SufBCD complex enhance SufS cysteine desulfurase activity as part of a sulfur transfer pathway for Fe-S cluster assembly in Escherichia coli. J Biol Chem. 2003;278:45713–9.CAS 
    PubMed 

    Google Scholar 
    74.Ayala-Castro C, Saini A, Outten FW. Fe-S cluster assembly pathways in bacteria. Microbiol Mol Biol Rev. 2008;72:110–25.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    75.Shepard EM, Boyd ES, Broderick JB, Peters JW. Biosynthesis of complex iron-sulfur enzymes. Curr Opin Chem Biol. 2011;15:319–27.76.Lill R. Function and biogenesis of iron–sulphur proteins. Nature. 2009;460:831–8.CAS 
    PubMed 

    Google Scholar 
    77.Seidler A, Jaschkowitz K, Wollenberg M. Incorporation of iron-sulphur clusters in membrane-bound proteins. Biochem Soc Trans. 2001;29:418–21.CAS 
    PubMed 

    Google Scholar 
    78.Buchanan BB, Schürmann P, Wolosiuk RA, Jacquot J-P. The ferredoxin/thioredoxin system: from discovery to molecular structures and beyond. Photosynth Res. 2002;73:215–22.CAS 
    PubMed 

    Google Scholar 
    79.Dubnau D, Losick R. Bistability in bacteria. Mol Microbiol. 2006;61:564–72.CAS 
    PubMed 

    Google Scholar 
    80.Resnekov O, Driks A, Losick R. Identification and characterization of sporulation gene spoVS from Bacillus subtilis. J Bacteriol. 1995;177:5628–35.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    81.Sonenshein AL. Bacteriophages: how bacterial spores capture and protect phage DNA. Curr Biol. 2006;16:R14–R16.82.Sullivan MB, Coleman ML, Weigele P, Rohwer F, Chisholm SW. Three Prochlorococcus cyanophage genomes: signature features and ecological interpretations. PLoS Biol. 2005;3:0790–806.CAS 

    Google Scholar 
    83.Fortier L-C, Sekulovic O. Importance of prophages to evolution and virulence of bacterial pathogens. Virulence. 2013;4:354–65.PubMed 
    PubMed Central 

    Google Scholar 
    84.Mobberley J, Nathan Authement R, Segall AM, Edwards RA, Slepecky RA, Paul JH. Lysogeny and sporulation in Bacillus isolates from the Gulf of Mexico. Appl Environ Microbiol. 2010;76:829–42.CAS 
    PubMed 

    Google Scholar 
    85.Brüssow H, Canchaya C, Hardt W-D. Phages and the evolution of bacterial pathogens: from genomic rearrangements to lysogenic conversion. Microbiol Mol Biol Rev. 2004;68:560–602.PubMed 
    PubMed Central 

    Google Scholar 
    86.Meinhart A, Alonso JC, Strater N, Saenger W. Crystal structure of the plasmid maintenance system /: functional mechanism of toxin and inactivation by 2 2 complex formation. Proc Natl Acad Sci. 2003;100:1661–6.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    87.Schuster CF, Bertram R. Toxin-antitoxin systems are ubiquitous and versatile modulators of prokaryotic cell fate. FEMS Microbiol Lett. 2013;340:73–85.88.Kawano M. Divergently overlapping cis -encoded antisense RNA regulating toxin-antitoxin systems from E. coli. RNA Biol. 2012;9:1520–7.CAS 
    PubMed 

    Google Scholar 
    89.Smith MA, Bidochka MJ. Bacterial fitness and plasmid-loss: the importance of culture conditions and plasmid size. Can J Microbiol. 1998;44:351–5.CAS 
    PubMed 

    Google Scholar 
    90.Summers DK. The kinetics of plasmid loss. Trends Biotechnol. 1991;9: 273–8.91.Persad AK, Williams ML, LeJeune JT. Rapid loss of a green fluorescent plasmid in Escherichia coli O157:H7. AIMS Microbiol. 2017;3:872–84.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    92.Modi SR, Lee HH, Spina CS, Collins JJ. Antibiotic treatment expands the resistance reservoir and ecological network of the phage metagenome. Nature. 2013;499:219–22.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    93.Hargreaves KR, Kropinski AM, Clokie MR. Bacteriophage behavioral ecology. Bacteriophage. 2014;4:e29866.PubMed 
    PubMed Central 

    Google Scholar 
    94.Naught LE, Gilbert S, Imhoff R, Snook C, Beamer L, Tipton P. Allosterism and cooperativity in Pseudomonas aeruginosa GDP-mannose dehydrogenase. Biochemistry. 2002;41:9637–45.CAS 
    PubMed 

    Google Scholar 
    95.Dong C, Flecks S, Unversucht S, Haupt C, van Pee K-H, Naismith JH. Tryptophan 7-halogenase (PrnA) structure suggests a mechanism for regioselective chlorination. Science. 2005;309:2216–9.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    96.Fouces R, Mellado E, Diez B, Barredo JL. The tylosin biosynthetic cluster from Streptomyces fradiae: genetic organization of the left region. Microbiology. 1999;145:855–68.CAS 
    PubMed 

    Google Scholar 
    97.Heacock-Kang Y, Zarzycki-Siek J, Sun Z, Poonsuk K, Bluhm AP, Cabanas D, et al. Novel dual regulators of Pseudomonas aeruginosa essential for productive biofilms and virulence. Mol Microbiol. 2018;109:401–14.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    98.Kurtov D, Kinghorn JR, Unkles SE. The Aspergillus nidulans panB gene encodes ketopantoate hydroxymethyltransferase, required for biosynthesis of pantothenate and Coenzyme A. Mol Gen Genet. 1999;262:115–20.CAS 
    PubMed 

    Google Scholar 
    99.Huisjes R, Card DJ. Methods for assessment of pantothenic acid (Vitamin B5). In: Harrington D, editor. Laboratory assessment of vitamin status. London, UK; San Diego, CA, USA; Cambridge, MA, USA; Oxford, UK : Elsevier Inc.; 2019. p. 265–299. https://doi.org/10.1038/s41396-021-01143-1.100.Leonardi R, Jackowski S. Biosynthesis of pantothenic acid and coenzyme A. EcoSal Plus. 2007;2:2.101.Begley TP, Kinsland C, Strauss E. The biosynthesis of coenzyme a in bacteria. Vitam Horm. 2001;61:157–71.CAS 
    PubMed 

    Google Scholar 
    102.Cameron B, Guilhot C, Blanche F, Cauchois L, Rouyez MC, Rigault S, et al. Genetic and sequence analyses of a Pseudomonas denitrificans DNA fragment containing two cob genes. J Bacteriol. 1991;173:6058–65.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    103.Doxey AC, Kurtz DA, Lynch MD, Sauder LA, Neufeld JD. Aquatic metagenomes implicate Thaumarchaeota in global cobalamin production. ISME J. 2015;9:461–71.CAS 
    PubMed 

    Google Scholar 
    104.Heal KR, Qin W, Amin SA, Devol AH, Moffett JW, Armbrust EV, et al. Accumulation of NO2-cobalamin in nutrient-stressed ammonia-oxidizing archaea and in the oxygen deficient zone of the eastern tropical North Pacific. Environ Microbiol Rep. 2018;10:453–7.CAS 
    PubMed 

    Google Scholar 
    105.Vik DR, Roux S, Brum JR, Bolduc B, Emerson JB, Padilla CC, et al. Putative archaeal viruses from the mesopelagic ocean. PeerJ. 2017;5:e3428.PubMed 
    PubMed Central 

    Google Scholar 
    106.Streisinger G, Emrich J, Stahl MM. Chromosome structure in phage T4, III. Terminal redundancy and length determination. Proc Natl Acad Sci USA. 1967;57:292–5.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    107.Mahmoudabadi G, Milo R, Phillips R. Energetic cost of building a virus. Proc Natl Acad Sci USA. 2017;114:E4324–E4333.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    108.Brum J. 5m intervals of CTD profiles from R/V New Horizon cruise NH1315 in the Eastern Tropical North Pacific (ETNP) during June 2013. Biological and Chemical Oceanography Data Management Office (BCO-DMO). (Version 1) Version Date 2020-08-31 (2020). https://doi.org/10.26008/1912/bco-dmo.822818.1.109.Noble RT, Fuhrman JA. Use of SYBR Green I for rapid epifluorescence counts of marine viruses and bacteria. Aquat Micro Ecol. 1998;14:113–8.
    Google Scholar 
    110.Brum J. Estimated abundances of viruses and bacteria determined in samples collected in the Eastern Tropical North Pacific (ETNP) on R/V New Horizon cruise NH1315 during June 2013. Biological and Chemical Oceanography Data Management Office (BCO-DMO). (Version 1) Version Date 2020-09-02 (2020). https://doi.org/10.26008/1912/bco-dmo.823094.1.111.Binder B. Reconsidering the relationship between vitally induced bacterial mortality and frequency of infected cells. Aquat Micro Ecol. 1999;18:207–15.
    Google Scholar 
    112.Brum J. Estimated frequency of lytic viral infection from samples collected in the Eastern Tropical North Pacific oxygen minimum zone region (ETNP OMZ) on R/V New Horizon cruise NH1315 during June 2013. Biological and Chemical Oceanography Data Management Office (BCO-DMO). (Version 1) Version Date 2020-09-01 (2020). https://doi.org/10.26008/1912/bco-dmo.822914.1.113.Abramoff MD, Magalhaes PJ, Ram SJ. Image processing with ImageJ. Biophotonics Int 2004;11:36–42.114.Brum J. Morphotypes, capsid widths, and tail lengths of viruses from samples collected in the Eastern Tropical North Pacific oxygen minimum zone region (ETNP OMZ) on R/V New Horizon cruise NH1315 during June 2013. Biological and Chemical Oceanography Data Management Office (BCO-DMO). (Version 1) Version Date 2020-09-02 (2020). https://doi.org/10.26008/1912/bco-dmo.823131.1.115.John SG, Mendez CB, Deng L, Poulos B, Kauffman AKM, Kern S, et al. A simple and efficient method for concentration of ocean viruses by chemical flocculation. Environ Microbiol Rep. 2011;3:195–202.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    116.Duhaime MB, Sullivan MB. Ocean viruses: Rigorously evaluating the metagenomic sample-to-sequence pipeline. Virology. 2012;434:181–6.CAS 
    PubMed 

    Google Scholar 
    117.Brum J. Accession numbers of viral metagenomes from samples collected in the Eastern Tropical North Pacific oxygen minimum zone region (ETNP OMZ) on R/V New Horizon cruise NH1315 during June 2013. Biological and Chemical Oceanography Data Management Office (BCO-DMO). (Version 1) Version Date 2020-09-04 (2020) https://doi.org/10.26008/1912/bco-dmo.823295.1.118.Peng Y, Leung HCM, Yiu SM, Chin FYL. IDBA-UD: A de novo assembler for single-cell and metagenomic sequencing data with highly uneven depth. Bioinformatics. 2012;28:1420–8.CAS 
    PubMed 

    Google Scholar 
    119.Delcher AL, Salzberg SL, Phillippy AM. Using MUMmer to identify similar regions in large sequence sets. Curr Protoc Bioinforma. 2003; Chapter 10: Unit 10.3.120.Finn RD, Bateman A, Clements J, Coggill P, Eberhardt RY, Eddy SR, et al. Pfam: the protein families database. Nucleic Acids Res. 2014;42:D222–D230.121.Eddy SR. Accelerated profile HMM searches. PLoS Comput Biol. 2011;7:1002195.
    Google Scholar 
    122.Team RCR. A language and environment for statistical computing. Vienna, Austria: R Foundation for Statistical Computing; 2018.123.Wickham H. ggplot2: elegant graphics for data analysis. Springer-Verlag, New York: 2016.124.Oksanen J, Blanchet FG, Kindt R, Legendre P, Minchin PR, O’Hara RB et al. vegan: Community Ecology Package. R package version 2.5-2. 2013 http://RForge.R-project.org/projects/vegan/.125.Wilkinson L. venneuler: Venn and Euler diagrams. R package version 1.1-0. 2011 https://CRAN.Rproject.org/package=venneuler.126.Harrell FE, With contributions from Charles Dupont and many others. Hmisc: Harrell Miscellaneous. R package version 4.3-0. 2019 https://CRAN.R-project.org/package=Hmisc.127.Wei T, Simko V. R package “corrplot”: Visualization of a Correlation Matrix (Version 0.84). 2017. Available from https://github.com/taiyun/corrplot.128.Emerson JB, Roux S, Brum JR, Bolduc B, Woodcroft BJ, Jang H Bin, et al. Host-linked soil viral ecology along a permafrost thaw gradient. Nat Microbiol. 2018;3:870–80.129.Sturges HA. The choice of a class interval. J Am Stat Assoc. 1926;21:65–66.130.Suzuki R, Shimodaira H. Pvclust: an R package for assessing the uncertainty in hierarchical clustering. Bioinformatics. 2006;22:1540–2.CAS 
    PubMed 

    Google Scholar  More

  • in

    Whales’ gigantic appetites, climate fears — the week in infographics

    NEWS
    05 November 2021

    Whales’ gigantic appetites, climate fears — the week in infographics

    Nature highlights three key infographics from the week in science and research.

    Share on Twitter
    Share on Twitter

    Share on Facebook
    Share on Facebook

    Share via E-Mail
    Share via E-Mail

    Climate scientists are scepticalThe momentous COP26 climate summit now under way in Glasgow, UK, represents one final opportunity for the governments of the world to craft a plan to meet their most ambitious goals for curbing climate change. Pledges are already flowing in, but the meeting has another week to run and much is still to be decided. Ahead of the summit, Nature conducted an anonymous survey of the 233 living authors of a climate-science report published in August by the Intergovernmental Panel on Climate Change, and received responses from 92 scientists — about 40% of the group. Their answers suggest strong scepticism that governments will markedly slow the pace of global warming, despite political promises made by international leaders as part of the landmark 2015 Paris climate agreement. Six in ten of the respondents, for example, said that they expect the world to warm by at least 3 °C by the end of the century, compared with conditions before the Industrial Revolution. That is far beyond the Paris agreement’s goal to limit warming to 1.5–2 °C.

    Source: Nature analysis

    Africa’s clinical trialsA shocking lack of COVID-19 vaccines in Africa, and the cost of existing treatments, means the continent really needs affordable, readily available COVID-19 drugs. These could reduce COVID-19 symptoms, lower the burden of disease on health-care systems and reduce deaths. The pandemic has given clinical research in Africa a boost: the Pan African Clinical Trials Registry recorded more clinical trials in 2020 than in 2019, and the number for 2021 is also on track to exceed 2019. But trials of COVID-19 drugs are still lacking in Africa, where they face infrastructure and recruitment challenges. One solution could be to establish a body to coordinate treatment trials on the continent.

    Source: https://pactr.samrc.ac.za

    The gluttony of whalesHow much do baleen whales, the largest known animals that have ever lived, eat? Three times as much as previously thought, report researchers who used cameras to study seven species of baleen whale. Writing in Nature, the researchers also suggest a feeding cycle involving iron and whale poo that could explain how such gluttony is possible. When whales eat iron-rich prey such as krill, they use the prey’s protein to make blubber — and defecate the iron-rich remains. Whale faeces might then provide a source of iron for microscopic marine algae called phytoplankton, and drive blooms of a type of plankton called diatoms. Diatoms, in turn, can move iron along the food chain when they are eaten by krill, which also excrete iron. Whales can further aid iron availability by mixing ocean waters through their vigorous tail movements.

    doi: https://doi.org/10.1038/d41586-021-03066-5

    Related Articles

    African scientists race to test COVID drugs — but face major hurdles

    Top climate scientists are sceptical that nations will rein in global warming

    A whale of an appetite revealed by analysis of prey consumption

    Subjects

    Ecology

    Climate change

    Public health

    Latest on:

    Ecology

    Baleen whale prey consumption based on high-resolution foraging measurements
    Article 03 NOV 21

    A whale of an appetite revealed by analysis of prey consumption
    News & Views 03 NOV 21

    A seagrass harbours a nitrogen-fixing bacterial partner
    News & Views 03 NOV 21

    Climate change

    Scientists cheer India’s ambitious carbon-zero climate pledge
    News 05 NOV 21

    Carbon emissions rapidly rebounded following COVID pandemic dip
    News 04 NOV 21

    Glass is the hidden gem in a carbon-neutral future
    Editorial 03 NOV 21

    Public health

    When are masks most useful? COVID cases offer hints
    News 04 NOV 21

    Why scientists worldwide are watching UK COVID infections
    News Explainer 02 NOV 21

    A reconstruction of early cryptic COVID spread
    News & Views 01 NOV 21

    Jobs

    Associate or Senior Editor (Genome engineering), Nature Communications

    Springer Nature
    London, Greater London, United Kingdom

    Senior Research Associate / Senior Scientist / Clinical Epidemiologist /Epidemiologist

    Blind Veterans UK
    London, United Kingdom

    Department Manager in the area of Microbial Physiology/Microbial Systems Biology

    Chr. Hansen A/S
    Hørsholm, Denmark

    Senior Scientist / Discovery project Leader, infectious diseases

    Vaccibody AS
    Oslo, Norway More

  • in

    Applications of unmanned aerial vehicles in Antarctic environmental research

    Identification and characterization of biotic and abiotic components in a penguin colony using RGB and multispectral camerasFigure 1 shows two image mosaics of a Chinstrap penguin (Pygoscelis antarcticus) colony, composed of 3800 pictures taken during a 29-min flight at 100 m altitude with a multispectral camera (MicaSense RedEdge-MX) using RGB bands (i.e., Red-668, Green-560 and Blue-475) (Fig. 1A) and the 10 wavelength bands covering the spectrum from visible to near-infrared light (Fig. 1B). With a resolution of 6 cm/pixel, penguin nests are clearly visible in the RGB mosaic, which are characterized by the absence of vegetation and with a predominant pink/brown color due to the abundance of guano deposition. The RGB mosaic also shows snow patches (white color), moss beds (green color) and one small lagoon with a bloom of red-pigmented greenalgae (Chlorophyceae) (Fig. 1A, upper right corner). Red algae (Chlamydomonas nivalis) patches on snow and ice are visible by zooming into a region of ice (Fig. 1A). More detailed information is obtained when the light spectrum from visible to near-infrared is used. Using the 10 wavelength bands, a thematic map was generated with the QGIS software and using a non-supervised classification method (Fig. 1B). Here it is possible to distinguish up to four species of mosses and three types of penguin guano that was verified with field observations.Figure 1Photomosaics of Vapour Col Chinstrap penguin colony on Deception Island composed of 3800 pictures taken at 100 m altitude with a 10 bands multispectral camera onboard a hexacopter, achieving 6 cm/pixel size. Panel (A): visible RGB mosaic (Red-668, Green-560 and Blue-475) with a zoom capture showing red snow patch; Panel (B): thematic map generated through non-supervised classification method.Full size imageDeception Island harbors up to 54 species of mosses, of which 13 species (including two endemics) have not been recorded elsewhere in the Antarctic. This, together with eight species of liverwort and 75 species of lichen, makes Deception Island an exceptional and unique place in Antarctica with legal protection under the Antarctic Treaty3. The use of a multispectral sensor onboard the UAV provides unique information to detect, classify and monitor moss beds without anthropogenic impacts. Antarctic moss bed health has already been assessed using multispectral sensors onboard UAVs12. Taxonomic identification would be feasible by indentifying previously each species in the field and later assigning the spectral signature using the UAV, as recently suggested by Miranda et al. (2020), who monitored lichens and mosses in the Antarctic using a combination of satellite imagery and UAVs13.Penguin guano has been suggested to be an important source of bioactive metals (e.g. Cu, Fe, Mn, Zn) for the sea surface waters, potentially fueling primary production of the Southern Ocean14. It has been suggested that the penguin species that feed mainly on Antarctic krill (Euphausia superba) (i.e., Chinstrap: Pygoscelis antarcticus, Adélie: Pygoscelis adeliae and Gentoo: Pygoscelis papua) excrete the highest concentrations of these bioactive metals15. Guano from these three congeneric penguin species has revealed the presence of microplastics across the Antarctic5. However, in order to estimate the magnitude of penguin fecal products that reach the sea, it is necessary to quantify the amount of guano excreted by the penguin colonies on land. This is possible with the multispectral reflectance data obtained from the UAV, which not only identify the guano coverage but also distinguishes different types of guano. Guano color is the result of diet, which, in turn, is related to the phase of the breeding cycle; therefore, a diet rich in krill is characterized by an excretion of pink guano, while a diet predominantly based on fish implies white guano16. Dark guano is the result of the mixture of guano with the soils that produce mud during wet precipitation.It is increasingly common in the Arctic and Antarctic to find well-developed algae blooms as highly visible red patches on the snow surface caused by red-pigmented green algae (Chlorophyceae), and that produce the phenomenon commonly-known as red snow17. These algal blooms play a crucial role in decreasing the snow-surface albedo and, consequently, accelerating the melt rate, as well as in nutrient and carbon cycling18,19. Mapping and monitoring the extent of snow algal blooms have so far been focused on satellite remote sensing; however, the spectral, temporal and spatial resolution of multi-spectral satellite imagery limits the study of most snow and ice algae18. Images taken from our UAV can enable the detection of patches of red snow on the surface snow with centimetric resolution (Fig. 1A). In addition, the image mosaic reveals the existence of a red snow bloom in a small pond located in a valley inside the colony (Supplementary Fig. S1). To the best of our knowledge, the existence of this bloom has not been previously reported and its monitoring could provide relevant information about the formation and proliferation of this bloom and its impact on cryospheric environments.As a whole, the image mosaic of the Chinstrap penguin colony in Vapour Col (the second largest breeding colony in the island with about 12,000 pairs of penguins20) may provide unique information about the different ecological niches linked to a penguin colony and their interactions. For example, the distribution and type of guano as nutrient and metal sources could be influencing the distribution and speciation of the flora in the area.3D geological formation using RGB cameraDeception Island is a complex volcanic system formed as a result of the explosive eruption of basaltic-to-andesitic magmas21. Among its multiple structures and stratigraphy, we surveyed the Murature formation, a consolidated andesitic lapilli tuff22. Using the quadcopter with a RGB camera and the software Pix4D we created a 3D photogrammetry of the Murature formation (Fig. 2; Supplementary Movie S1). The software uses a Structure from Motion photogrammetry algorithm, where obtained 3D points are interpolated to form a triangulated irregular network in order to obtain digital Surface model (DSM). This DSM is then used to project every image pixel and to calculate the georeferenced orthomosaic. For the Murature formation, the photogrammetry was generated with 843 pictures obtained from three 20-min flights at an altitude of 40 meters, taking pictures from two different angles to obtain the heights of the features (60° and 90°). With 1.4 cm/pixel resolution the resulting mosaic provides a unique view of the geological formation that will support the study of how the rocks were formed and its evolution in relation to the various geological processes that occurred on the island. 3D photogrammetry is also useful in geomorphological research. Specifically, in Deception Island morphometrics studies of landform (e.g. Crater and cone diameters, depths, slopes, heights, etc.) are useful to estimate the eruptive recurrence of the island, and in turn, for advising volcanic hazards23.Figure 23D photogrammetry of the Murature formation built with 843 RGB pictures taken from the RGB Hasselblad camera quadcopter DJI Mavic 2 Zoom at 40-m altitude, achieving 1.4 cm/pixel size.Full size imageThermal imagery to estimate animal abundance and to detect thermal anomaliesThe combination of UAV technology with a thermal-imaging camera is very useful for studying and monitoring wildlife and thermal anomalies on Deception Island. Chinstrap penguin and fur seal (Arctocephalus gazella) heat signatures were detected at Vapour Col and Baily Head, respectively (Fig. 3A, E). Figure 3A shows a mosaic from a Vapour Col section composed of 336 images taken with a thermal camera (FLIR Vue Pro R) onboard the hexacopter during a 29-min flight at 100 m altitude, whereas Fig. 3C shows one thermal picture of fur seals at Baily Head. Penguins and fur seals, with a thermal signature of 15 °C and 26 °C, respectively, are clearly identified. Penguins are highly sensitive to climate change and are considered “marine sentinels” for quantifying environmental change in the Southern Ocean24. However, the distribution and population dynamics of species such as the Chinstrap penguin are not well understood, mainly because they nest in remote and rugged areas, on-the-ground census work is difficult and sporadic25. As demonstrated for Adelia penguins26 the use of thermal imagery would allow reliable population estimates of Chinstrap penguins. Even, the use of RGB aerial images for animal counting would be far more accurate than from land-based surveys. Nevertheless, the scientific challenge is to develop a machine learning algorithm that can distinguish between animal species, based on their morphology and unique thermal fingerprint, which is only feasible using the high resolution provided by UAVs.Figure 3Thermal imagery. Panel (A): thermal mosaic of a section of Vapour Col (8.5 cm/pixel). Penguins are distinguished throughout the colony as small dots around 15 °C; Panel (B) and (D): RGB (Red-668, Green-560 and Blue-475 bands) and thermal picture of fumarole at Fumarole Bay (5.4 cm/pixel), respectively; Panel (C) and (E): RGB and Thermal image of Fur seals at Baily Head (5.4 cm/pixel), respectively.Full size imageOther useful application of thermal cameras onboard UAVs on Deception Island is the easy and precise detection and monitoring of thermal anomalies. Figure 3B–D shows a thermal picture of one of the multiple fumaroles on the island, reaching temperatures above 90 °C. Seismic monitoring of volcanos on Deception Island has being ongoing since 1986, including many recorded volcano-tectonic earthquakes, long-period events and volcanic tremor27. There have been six documented volcanic eruptions on the island between 1841 and 197128, nowadays volcanic and geothermal activities are limited to fumaroles and hot sands. Monitoring of these fumaroles using UAVs can provide a key in surveillance for early warming systems alerting of volcano activity on the island. UAVs not only accurately detect changes in temperature but also allow the increase in monitoring frequency when required.Surface water samplingUAVs provide unique opportunities for remote sample collection from surface waters, particularly in harsh or dangerous environments. Using a surface water sampling device described in the sampling and method sections we collected filtered fresh and saline surface waters at: (1) Three locations in Crater Lake (Fig. 4A). Crater Lake is part of the Antarctic Specially Protected Area (ASPA 140) due to its exceptional botanic and ecological value3. The use of drones for water sampling avoids human disturbance through the transportation and use of infrastructure, such as inflatable boats, and the risk that they pose to the natural ecological system. (2) One and six coastal locations in the Vapour Col and Baily Head penguin colonies, respectively (Fig. 4B, C). Access to the coastal zone inhabited by penguins requires approaches by boat (often assisted by an oceanographic vessel). The approaches do not only disturb the penguins that enter and exit the colony but, due to the coastal orography and waves, also dangerously hinders such an operation. The surface water sampling device onboard the UAV allowed in-situ water collection, minimizing the risk of impact on flora and fauna, limiting water disturbance and preventing contamination in the trace metal analysis. Attached to the sampling system we included a small multiparametric instrument referenced with time and GPS position to measure ancillary parameters, such as conductivity, temperature and depth (CastAway-CTD®) (Fig. 4D). The aerial water sampling has been validated for trace metal analysis using ICP-MS by comparing metal concentrations of samples collected in a saline pond with the surface water sampling device onboard the UAV (i.e. average ± SD, n = 3; Ti: 0.20 ± 0.09; V: 1.92 ± 0.07; Cr: 1.5 ± 0.1; Mn: 19.4 ± 0.4; Fe: 11.6 ± 0.5; Cu: 1.9 ± 0.2; Zn: 0.5 ± 0.3; all values in ppb) and the traditional peristaltic pump system used from land or on boats29 (i.e. average ± SD, n = 3; Ti: 0.20 ± 0.06; V: 1.93 ± 0.09; Cr: 1.3 ± 0.1; Mn: 19.1 ± 0.3; Fe: 11.8 ± 0.3; Cu: 2.1 ± 0.4; Zn: 0.4 ± 0.3; all values in ppb).Figure 4Locations of surface water samples collected in Crater lake (A), Vapour Col (B), and Baily Head (C) using aerial water sampling device, and picture of the UAV (hexacopter) carrying, at 100 m altitude, the water sampling device and the multiparametric instrument (D). Stations at Crater lake are plotted on a mosaic composed of 3096 pictures taken during three flights of 14 min each at 120 m altitude using a quadcopter with an integrated RGB camera and a multispectral camera array with 5 bands, achieving 6.5 cm/pixel size.Full size imageDeception Island is an example of the complexity of Antarctic environments, where environmental research studies need to deal with the inter- and multi-disciplinary analysis of processes, such as volcanic and geothermal activities, limnological process from its multiple lakes and ponds, sparse and exceptional flora and diverse fauna, among other. UAV surveys on Deception Island have demonstrated that this technology may substantially contribute to the progress in environmental biological, geological and chemical studies. UAVs permit researchers to study environmental processes at smaller spatial and temporal scales compared to other remote platforms (e.g. satellites), in a more cost-effective and safer way than on foot studies. Furthermore, they are less invasive and less disturbing to wildlife and the ecosystem. The simultaneous use of multi-sensors for multiple applications and the development of algorithms based on images obtained from the drone to detect, classify and count animals in real time are the new challenges that would significantly contribute to the study of the functioning of the Antarctic ecosystem and its ongoing environmental processes. More

  • in

    Modelling the growth, development and yield of Triticum durum Desf under the changes of climatic conditions in north-eastern Europe

    Climatic conditions and phenologyThe growth and development of T. durum plants was moderately differentiated by weather conditions in the analyzed years (Table 1).Table 1 The duration of growing seasons (Days), sum of temperatures (Temp.) and sum of precipitation (Prec.) during the growth and development of T. durum Desf. in the analyzed years.Full size tableThe growing seasons of 2015, 2016 and 2017 lasted 136, 132 and 145 days, respectively; the sum of temperatures was determined at 2011.3, 1895.6 and 2069.9 °C, respectively, and the sum of precipitation was determined at 366.7, 360.1 and 350.5 mm, respectively. However, a comparison of cumulative temperatures and precipitation in the phenological phases of T. durum in each year of the study indicates that temperature and precipitation could have influenced the duration of the examined phases and plant growth indicators (Fig. 1). Weather conditions were generally favorable for the growth and development of T. durum in 2015 and 2016. Cumulative temperatures and precipitation were quite similar in 2015 and 2016 up to the booting stage, but precipitation levels in successive stages were higher in 2016 than in 2015. The growing season was shortest in 2016 and longest in 2017, mainly due to low temperatures during sowing and seed germination, and high precipitation during tillering, grain formation and ripening.Figure 1Cumulative temperatures and precipitation in the phenological phases of T. durum in 2015–2017.Full size imageBiophysical parameters: LAI and SPADThe LAI denotes the area of photosynthetic tissue per unit ground surface area (m2 m−2). The LAI is directly associated with plant canopy, and it is an indicator of net primary production, water and nutrient use, and the carbon balance. SPAD is a measure of leaf greenness that is directly associated with chlorophyll content and nitrogen sufficiency.The main effects of LAI and SPAD were analyzed separately in the framework of the Zadoks scale to reveal the significant effects of years, nitrogen rates and sowing density, and an absence of significant effects associated with the application of the growth regulator (see Tables 1.1–1.6 in the Supplementary Information). In the analyzed years, LAI and SPAD were similar in the 2nd node detectable stage (Z32), but they differed in the stem elongation stage (Z45) and the ear emergence and heading stage (Z59), when LAI values were higher and leaf greenness values were lower in 2016 and 2017 than in 2015. These findings can be attributed to moderate temperatures and precipitation in 2015, and high precipitation in the critical growth stages in the remaining years. The general trend associated with the nitrogen rate was similar across the examined growth stages, i.e. a significant increase in LAI and SPAD values with nearly identical effects were noted in treatments with nitrogen rates of 80 and 120 kg ha−1. A similar trend was observed in sowing density. In treatments with a sowing density of 450 and 550 germinating seeds per m2, LAI values continued to increase, whereas SPAD values were below those noted in the treatment with a sowing density of 350 germinating seeds per m2. The only significant interaction was observed between years and nitrogen rates.The average values of LAI continued to increase in successive growth stages and were determined at 1.30 at Z32, 1.75 at Z45, and 1.99 at Z59. In turn, leaf greenness was significantly lower in the stem elongation stage (Z45) than in the preceding (Z32) and subsequent (Z59) stages.The significant effect of the years × growth stages interaction for LAI and SPAD values resulted from similar means in stage Z32 in all years, as well as higher LAI values and lower SPAD values in subsequent growth stages in 2016 and 2017 than in 2015. In 2015, the increase in the nitrogen rate induced only a rising trend in LAI and SPAD values, whereas significant differences were observed in 2016 and 2017. To summarize, it should be noted that in successive Zadoks growth stages, the interactions between years, nitrogen rates and sowing density exerted significant effects on LAI and SPAD values, whereas the effects of year × nitrogen rate interactions were significant only in selected growth stages.Contribution of different sources of variation to physiological and biophysical parameters of plant growthThe calculated eta-squares η2 provide information about the contribution of different sources of variation to physiological variables (Table 2). The experimental years and agronomic factors (33.1% and 38.6%), growth stages, and interactions with other factors (32.5% and 39.3%) and random factors (34.4% and 22.1%) made similar contributions to the variation in the LAI and chlorophyll content. The variation in the net photosynthetic rate was related mostly to variations across years (32.6%) and the interactions between growth stages and other factors (24.3%). The variation in the transpiration rate was attributed mostly to the main effects of growth stages (45.8%) and the year × growth stage interaction (16.1%). Instantaneous WUE was strongly determined by variation in agronomic factors and growth stages (22.3% and 21.1%, respectively).Table 2 Eta-square (η2) values for the sources of variation in the leaf area index (LAI), chlorophyll content (SPAD), net photosynthetic ratio (Pn), transpiration rate (E) and instantaneous water use efficiency (WUE).Full size tableIt is worth noting that the variation in agronomic factors made a considerable contribution to the total variation in LAI (22.3%) and SPAD (11.8%), but only a marginal contribution to the net photosynthetic rate (0.4%) and transpiration (2.0%).Photosynthetic indicators— net photosynthetic rate, transpiration rate, and instantaneous water use efficiencyThe effects of the net photosynthetic rate (Pn), transpiration rate (E) and instantaneous WUE were highly differentiated in successive growth stages, and relatively small differences were noted for agronomic factors (see Tables 2.1–2.9 in the Supplementary Information). At the same time, the analyzed photosynthetic indicators differed in successive stages of growth. The net photosynthetic rate was similar in the 2nd node detectable stage (Z32) and the stem elongation stage (Z45) at 29.7 μmol CO2 m–2 s–1, and it was 15% higher at the end of the heading stage (Z59) than in the preceding stages. The transpiration rate continued to increase by 60% on average in successive stages of growth and development, from 1.59 H2O m–2 s–1 in stage Z32, to 2.52 mmol H2O m–2 s–1 in stage Z45, and 4.06 mmol H2O m–2 s–1 in stage Z59.An analysis of the results noted in different growth stages across years revealed significant year × growth stage and growth regulator × growth stage interactions (Fig. 2).Figure 2Mean values and standard error of photosynthesis indicators for year × growth stage (upper) and growth regulator × growth stage interactions (GR 0—without growth regulator, GR 1—with growth regulator).Full size imageThe year × growth stage interaction resulted from differences in the rates of photosynthesis and transpiration in the analyzed growth stages across years. In 2015, the net photosynthetic rate was similar in the first two growth stages, and it increased by around 30% at the end of the heading stage (Z59). In 2016, the photosynthetic rate continued to increase in successive growth stages. In 2017, the net photosynthetic rate was around 10% higher in the 2nd node detectable stage (Z32) than in the stem elongation stage (Z45) and at the end of the heading stage (Z59). The transpiration rate increased significantly in successive stages of plant growth and development, and the only exception was noted in 2015, when the analyzed parameter was similar in stages Z32 and Z45. The WUE index was highest in stage Z32, and a significant interaction was noted due to the correlation between the net photosynthetic rate and the transpiration rate in the remaining stages. Water use efficiency was similar in stages Z32 and Z45 in 2015, and in stages Z45 and Z59 in 2016, whereas significantly lower values in successive stages of plant growth were noted in 2017.The growth regulator was the only agronomic factor that induced significant differences in the net photosynthetic rate across the examined growth stages. Photosynthesis indicators were similar regardless of the application of the growth regulator, and significant interactions resulted mainly from varied disproportions between the end of heading and the stem elongation stage in treatments with and without the application of the growth regulator.It should be noted that the interactions between growth stages and nitrogen rates and sowing density were not significant, which implies that the effects of the interactions between increasing nitrogen rates and sowing density on photosynthetic indicators in successive growth stages were similar to the average values of photosynthetic indicators in the corresponding growth stages (Supplementary Information).Agronomic traitsThe means for yield components and yield are presented in Tables 3.1–3.8 of the Supplementary Information. Stem length was differentiated by the nitrogen rate and nitrogen rate × year interaction. Nitrogen rates of 80 and 120 kg N per ha increased stem length by 11% and 13%, respectively, relative to the unfertilized control. The significant year × nitrogen rate interaction resulted from the fact that the nitrogen-induced increase in stem length was smaller in 2015 (0.07 cm per 1 kg of nitrogen) than in 2016 and 2017 (0.09 cm per 1 kg of nitrogen). In 2015, ear length was similar to that noted in the remaining years, and only in 2017, ear length was 7% higher than in 2016. Ear length and the number of kernels per ear increased with a rise in nitrogen rate and decreased with a rise in sowing density.Grain weight per ear and 1000 kernel weight were highest in 2015 and significantly lower in the following years. Grain weight per ear increased only in response to the nitrogen rate of 120 kg ha−1, but 1000 kernel weight was not affected. Both traits decreased with a rise in sowing density. The significant year × nitrogen rate and year × sowing density interactions for both traits can be largely attributed to the magnitude of differences between years, rather than an increase or a decrease in this trend.The biological yield (grain and straw) differed across years and nitrogen rates. In 2016, the biological yield was similar to that noted in 2015 and significantly higher (by 30%) than that noted in 2017. The biological yield increased by 28% and 35% in response to nitrogen rates of 80 and 120 kg ha−1, respectively, relative to the unfertilized control. The significant year × nitrogen rate interaction was associated with variations in nitrogen use efficiency, and the difference between maximal biological yield was determined at 0.5 t ha−1 in 2015, 2.3 t ha−1 in 2016, and 2.8 t ha−1 in 2017.Grain yield was similar in 2015 (4.94 t ha−1) and 2016 (5.38 t ha−1), and it was significantly lowest in 2017 (3.87 t ha−1). Straw yield was highest in 2016 (2.86 t ha−1), and it exceeded the values noted in the remaining years by 16%. The harvest index was similar in 2015 and 2016, and it was 9% lower in 2017. Grain yield increased by 30% and 36%, whereas straw yield increased by 20% and 35% in response to the nitrogen rates of 80 and 120 kg ha−1, respectively. A minor increase in grain yield (3%) was observed in treatments with a sowing density of 550 seeds m−2 relative to the remaining sowing densities.Path modellingA simple correlation analysis of manifest variables in all phenological stages revealed significant correlations between the LAI and leaf greenness (SPAD) only in stage Z32, as well as a very strong correlation between the net photosynthetic rate and the transpiration rate, which was positive in stages Z32 and Z45 and negative in stage Z59. No simple correlations were noted between the indicators of physiological processes (Pn, E, WUE) and biophysical parameters (LAI, SPAD). AAll correlations between the manifest variables of yield components and biological yield were statistically significant, excluding the correlation between stem length and ear length (Supplementary information).All correlations between the manifest variables of yield components and biological yield were statistically significant, excluding the correlation between stem length and ear length (Supplementary Information). The outer and inner PLS-PM models well fit the data, and their goodness of fit was determined at 0.973 and 0.786, respectively. The outer weights provide information about the relative importance of a manifest variable for the corresponding latent variable (for details please see the Supplementary Information). Outer weights that exceed 0.3 are considered meaningful. By the same token, loading estimates represent the correlations between a latent variable and the corresponding manifest variables. Loadings higher than 0.7 capture more than 50% of the variability contributed by a latent variable to the corresponding manifest variable. In general, both indicators in the outer model, i.e. outer weights and loadings, exceeded the thresholds, which indicates that manifest variables were strongly related with latent variables. Growth regulators (({w}_{GR}) = − 0.007) and the length of the growing season (({w}_{DAYS})=0.197) provided the only evidence for the low explanatory value of latent variable A (agronomic factors).In the inner model, all equations that regressed latent variables well fit the data and were statistically significant (Table 3). The latent variables expressed by the value of R2 increased in successive stages of T. durum growth and development, from 0.218 in physiological processes in stage Z32 (Table 3, Eq. 1) to 0.698 and 0.708 in yield components and Biological Yield, respectively (Table 3, Eqs. 7 and 8). It is worth noting that in successive stages of growth, the value of physiological processes was relatively lower in comparison with biophysical parameters.Table 3 Parameters of regression models for latent variables.Full size tableThe analysis of path coefficients (βi) revealed that agronomic factors (A) and climate conditions (CC) in stages Z32, Z45 and Z59 exerted a specific influence on physiological processes (PP) and biophysical parameters (BP) of T. durum plants. Agronomic factors directly determined physiological processes in all stages and biophysical parameters in stages Z32 and Z59. At the same time, climate conditions did not exert a direct influence on physiological processes in any stage, but directly affected biophysical parameters in all stages. All of the modeled parameters, i.e. agronomic factors, climate conditions and physiological processes, significantly influenced biophysical parameters in stages Z32 and Z59, but not Z45. Consequently, it can be stated that agronomic factors were the main determinant of variability in physiological processes (photosynthesis, transpiration) in a model evaluating the impact of agricultural practices on yield and the manifest variables associated with T. durum growth and development. At the same time, physiological processes made a significant but negative contribution to biophysical parameters. A one unit increase in photosynthesis processes with constant values of agronomic factors and climate conditions implies a decrease of − 0.382, − 0.065 and − 0.395 in biophysical parameters in stages Z32, Z45 and Z59, respectively.The performance of every preceding latent variable in terms of its total impact on the target latent variable, i.e. the biological yield of T. durum (IPMA – Importance-Performance Map Analysis), was analyzed to highlight latent variables associated with agricultural practices that improve biological yield. The total effect (importance) of preceding latent variables (A, CC32, PP32, BP32, CC45, PP45, BP45, CC59, PP59, BP59, YC and CC) on the anticipated performance of the specific target (Biological Yield) is presented in Fig. 3.Figure 3Importance-Performance Map Analysis presenting the impact of latent variables on biological yield (A—agronomic factors, YC—yield components, CC32, CC45, CC59—climate conditions in growth stages, PP32, PP45, PP59—physiological processes, BP32, BP45, BP59—biophysical parameters in the phenological stages of plant growth and development Z32, Z45 and Z59, CC—climate conditions for the entire growing season).Full size imageThe importance and performance of latent variables that influenced the biological yield of T. durum varied. The biological yield of T. durum was affected mostly by agronomic factors (A), followed by yield components (YC) and biophysical parameters (BP) in growth stages Z59 (BP59) and Z32 (BP32), climate conditions in stage Z59 (CC59), and climate conditions in stage Z32 (CC32). A one unit increase in the above latent variables led to an increase of 0.575, 0.422, 0.234, 0.203 and 0.109 units in biological yield, respectively. At the same time, the performance scores of these latent variables were determined at 53.1, 53.5, 67.1, 59.6 and 61.8, respectively (scores closer to 100 denote higher performance). The remaining latent variables, in particular climate conditions for the entire growing season and physiological processes in stage Z32, were characterized by low importance and exerted a relatively small effect on biological yield performance.The results of the importance-performance analysis clearly indicate that latent variables have considerable potential to optimize the agricultural conditions for the growth and development of T. durum plants. More

  • in

    Detection of heteroplasmy and nuclear mitochondrial pseudogenes in the Japanese spiny lobster Panulirus japonicus

    Direct nucleotide sequencingReadable electropherograms were obtained from both direction in COI fragments of all three individuals of the Japanese spiny lobster. COI sequences determined by direct nucleotide sequencing ranged from 807 to 864 bp and have been deposited in International Nucleotide Sequence Database Collection (INSDC) under accession numbers of LC571524‒LC571526. No stop codon was observed in these sequences (designated by PJK1-direct, PJK2-direct, and PJK3-direct). No indel was observed between these sequences. All nucleotide substitutions at 19 variable sites observed between these sequences were transition at the 3rd position of a codon, and all substitutions were synonymous. The mean Kimura two parameter (K2P) distance between these three haplotypes was 1.510 ± 0.352% SE and that between these sequences and a reference sequence of P. japonicus (NC_004251) was 1.087 ± 0.270%, which were all well within the range reported for Japanese spiny lobster samples collected in Japan and Taiwan9,10.Electropherograms obtained by forward primer for 12S fragments were not readable, while those by reverse primer were readable in all individuals. 12S sequences determined by direct nucleotide sequencing using reverse primer alone ranged from 551 to 570 bp and have been deposited in INSDC under accession numbers of LC605705‒LC605707. Of nine variable sites, eight were transition and one was indel. The mean K2P distance between these three haplotypes (designated by PJK1-12Sdirect, PJK2-12Sdirect, and PJK3-12Sdirect) was 0.970 ± 0.338%, and that between these sequences and a reference sequence of P. japonicus was 0.835 ± 0.282%.Electropherograms obtained by both primers for Dloop fragments were readable only in one individual (PJK2). This Dloop sequence determined by direct nucleotide sequencing was 762 bp and deposited in INSDC under accession number of LC605749. K2P distance between this haplotype (designated by PJK2-Dloopdirect) and a reference sequence of P. japonicus was 3.666%. No indel was observed between the two sequences, and 25 of 27 variable sites were transition.Phylogenetic analysis of clones, heteroplasmy and NUMTsAmong the 36–42 positive COI clones examined per individual, sequences (809–892 bp) of 22–31 clones per individual (75 clones in total) were successfully determined. After alignment, both ends of all sequences were trimmed to fit the shortest sequence obtained by direct nucleotide sequencing, yielding 774–810 bp sequences. Eleven clones of PJK1 were identical to PJK1-direct, as well as seven of PJK2 to PJK2-direct and three of PJK3 to PJK3-direct. These dominant haplotypes (807 bp) were determined to be genuine COI haplotypes of each individual, and representative sequences of these three genuine haplotypes were deposited in INSDC (LC 571527, LC571533 and LC571538). Nucleotide sequences of the remaining 54 clones were all different one another, in which 20 haplotypes were observed in PJK1, 14 in PJK2, and 20 in PJK3 (LC571541–LC571577, OK429332–OK429343, LC654683-LC654687).Phylogenetic tree constructed using three genuine COI haplotypes, 57 unique haplotypes and eight sequences of reference lobster species is shown in Fig. 1. Haplotypes detected from P. japonicus were segregated into four groups (designated by A, B, C and D). Among the outgroup species used, Australian rock lobster (P. cygnus) that morphologically and genetically belongs to the P. japonicus group11,12, appeared to be the closest kin to all haplotypes detected from P. japonicus. All haplotypes in group A were of the same length (807 bp), and no indel was observed. Three distinct clades (designated by c-I to c-III) were observed in group A, in which 14 haplotypes from PJK1, 11 from PJK2 and 11 from PJK3 were cohesively clustered together with their corresponding genuine haplotypes (bold italic). PJK1-C25 was outlier, having 10 nucleotide differences from the genuine COI sequence. The numbers of variable nucleotide sites between haplotypes within c-I, c-II and c-III were 20, 15 and 26, respectively, of which nonsynonymous nucleotide substitutions were observed at 11, 13 and 10 sites. Stop codon was observed only in one haplotype (PJK3-C1). The mean K2P distance between different haplotypes within these clades ranged from 0.320 ± 0.075 to 0.561 ± 0.103%. The mean K2P distances between three clades ranged from 1.343 ± 0.339 to 2.178 ± 0.464%. Although group A must be composed of sequences containing those caused by Taq polymerase error or true heteroplasmic sequences as well as genuine haplotypes, it is difficult to determine the former two categories. All of the non-genuine haplotypes in group A had singleton difference one another, supporting the occurrence of Taq polymerase error. We determined haplotypes (marked with dagger in Fig. 1) differed by less than two substitutions from the genuine haplotype to be due to Taq polymerase error. This criterion may be reasonable, since Taq polymerase-mediated errors were estimated to occur approximately at a frequency of 7.2 × 10−5 per bp per cycle13 to one mutation per 10,000 nucleotides per cycle14. When Taq polymerase error is taken into account, these K2P distances within and between clades and number of haplotypes are likely to be somewhat overestimated. PJK1-C25, two (PJK1-C5 and PJK1-C60) in c-I clade, one (PJK2-C26) in c-II, and five (PJK3-C1, PJK3-C5, PJK3-C26, PJK3-C31, PJK3-C34) in c-III differed by 3 to 10 nucleotides from their genuine haplotypes, which were determined to be heteroplasmic haplotypes.Figure 1Neighbor-joining phylogenetic (NJ) tree showing relationships among 57 different haplotypes of cytochrome oxidase subunit I (COI) or COI-like sequences obtained from the Japanese spiny lobster (Panulirus japonicus), and COI sequences of eight congeneric species derived from the GenBank database. Haplotypes detected from the same individual of the Japanese spiny lobster share the same color. Genuine mtDNA haplotype is shown in bold italic and number of clones examined is shown in parenthesis. Stop codons were observed in haplotypes carrying asterisk. Haplotypes carrying dagger differ from the corresponding genuine mtDNA haplotype by less than two nucleotides (including indel). The bootstrap values greater than 60% (out of 1000 replicates) are shown at the nodes.Full size imageSequence size of haplotypes in groups B to D ranged from 774 to 810 bp. K2P distance between haplotypes of groups A and B ranged from 7.169 to 8.177% with a mean of 7.754 ± 0.973%, that between A and C ranged from 12.073 to 17.392% with a mean of 14.521 ± 1.151%, and that between A and D ranged from 17.472 to 23.880% with a mean of 21.042 ± 1.600%. Multiple stop codons were observed in a haplotype of group B, in five of eight haplotypes of group C, and all haplotypes of group D. Three haplotypes in group C had no stop codon but differed in four to 10 deduced amino acids from the genuine haplotypes. BLAST homology search revealed no identical sequence for haplotypes in groups B to D but indicated that the closest species were P. japonicus or P. cygnus with moderate similarity (83–89% homology). Therefore, all haplotypes of groups B to D (LC571565–LC571570, LC571572–LC571577, LC654683-LC654687) were determined to be NUMTs.Among the 30–35 positive 12S clones examined per individual, sequences (772–806 bp) of 25–27 clones per individual (77 clones in total) were successfully determined. After alignment, primer sequences were trimmed, yielding 731–765 bp sequences. Thirteen clones of PJK1 were identical one another, as well as 12 of PJK2 and three of PJK3, and these were identical to PJK1-12Sdirect, PJK2-12Sdirect and PJK3-12Sdirect, respectively. These dominant haplotypes ranging from 761 to 762 bp in size were determined to be genuine 12S haplotypes of the individual, and representative sequences of these three genuine haplotypes were deposited in INSDC (LC605708‒LC605710). Nucleotide sequences of the remaining 49 clones were all different one another, in which 12 haplotypes were observed in PJK1, 23 in PJK2, and 14 in PJK3 (LC605711‒LC605748, OK429126–OK429131, LC654678-LC654682).Since incorporation of all eight Panulirus species sequences made sequence alignment ambiguous because of multiple indels, reference sequences of P. japonicus and of closely related P. cygnus were used for constructing phylogenetic tree (Fig. 2). Haplotypes detected from P. japonicus were segregated into three groups (designated by A to C). Sequence size of haplotypes in group A ranged from 760 to 762 bp. Three distinct clades (s-I to s-III) were observed in group A, in which 12 haplotypes each from PJK1, PJK2 and PJK3 were cohesively clustered together with their corresponding genuine haplotypes (bold italic). The numbers of variable nucleotide sites between haplotypes within s-I, s-II and s-III were 24, 17 and 16, respectively. Of these variable sites, transversion was observed at five, one and three sites, and indel was observed at one, zero and one sites, respectively. The mean K2P distances between different haplotypes within these clades ranged from 0.345 ± 0.081 to 0.519 ± 0.101%. The mean K2P distances between three clades ranged from 0.936 ± 0.275 to 1.371 ± 0.359%. Haplotypes differed by less than two substitutions (including indel) from the genuine haplotypes are marked with dagger. Five haplotypes in s-I clade and two haplotypes in s-III clade differed by three to six nucleotides from their genuine haplotypes, which were determined to be heteroplasmic copies.Figure 2Neighbor-joining phylogenetic (NJ) tree showing relationships among 52 different haplotypes of clones of 12S rDNA (12S) or 12S-like sequences obtained from the Japanese spiny lobster (Panulirus japonicus), and 12S rDNA sequences of P. japonicus and P. cygnus derived from the GenBank database. Haplotypes detected from the same lobster individual share the same color. Genuine mtDNA haplotype is shown in bold italic and number of clones examined is shown in parenthesis. Haplotypes carrying dagger differ from corresponding genuine mtDNA haplotype by less than two nucleotides (including indel). The bootstrap values greater than 60% (out of 1000 replicates) are shown at the nodes.Full size imageSequence size of haplotypes in group B varied from 731 to 762 bp. K2P distance between groups A and B ranged from 1.336 to 7.445% with a mean of 3.449 ± 0.398%, and those between a reference sequence of P. japonicus and groups A and B were 0.864 ± 0.236% and 3.189 ± 0.410%, respectively. Sequence size of haplotypes in group C varied from 744 to 765 bp. K2P distance between groups A and C ranged from 3.104 to 22.434% with a mean of 12.049 ± 0.901%, and those between a reference sequence of P. japonicus and group C ranged from 3.951 to 21.287% with a mean of 11.764 ± 0.901%. BLAST homology search indicated that the closest species for haplotypes in groups B and C was P. japonicus or P. cygnus with moderate to high similarity (84–98% homology). Therefore, all 13 haplotypes (LC605741‒LC605748, LC654678-LC654682) in groups B and C were determined to be NUMTs.Among the 36–49 positive Dloop clones examined per individual, sequences (777–893 bp) of 26–38 clones per individual (92 clones in total) were successfully determined. After alignment, primer sequences were trimmed, yielding 736–853 bp sequences. Three clones (821 bp) of PJK1 were identical one another and determined to be genuine haplotype of this individual. Nine clones (813 bp) of PJK2 were identical to PJK2-Dloopdirect and determined to be genuine haplotype of this individual. Three clones (821 bp) of PJK3 were identical one another and determined to be genuine haplotype of this individual. Representative sequences of these three genuine haplotypes were deposited in INSDC (LC605750‒LC605752). Nucleotide sequences of the remaining 78 clones were all different one another, in which 25 haplotypes were observed in PJK1, 17 in PJK2, and 36 in PJK3 (LC605753‒LC605815, LC654419-LC654430, LC654675-LC654677).Incorporation of all eight Panulirus species sequences made sequence alignment considerably unreliable because of multiple indels, reference sequences of P. japonicus and of closely related P. cygnus were used for constructing phylogenetic tree (Fig. 3). Haplotypes detected from P. japonicus were segregated into four groups (designated by A to D). Sequence size of haplotypes in group A ranged from 812 to 822 bp. Three distinct clades (d-I to d-III) were observed in group A, in which 17 haplotypes from PJK1, 13 from PJK2 and 15 from PJK3 were cohesively clustered together with their corresponding genuine haplotypes (bold italic). The numbers of variable nucleotide sites between haplotypes within d-I, d-II and d-III were 27, 61 and 28, respectively, of which indels were observed at five, two and four sites and transversion was observed at 0, six and six sites. The mean K2P distance between different haplotypes within these clades ranged from 0.340 ± 0.067 to 1.097 ± 0.139%. The mean K2P distance between these three clades ranged from 7.577 ± 0.951 to 8.770 ± 0.984%. Haplotypes differed by less than two substitutions (including indel) from the genuine haplotypes are marked with dagger. Eight haplotypes in d-I clade, three in d-II clade, and four in d-III clade differed by three to five nucleotides from the genuine haplotype were determined to be heteroplasmic copies.Figure 3Neighbor-joining phylogenetic (NJ) tree showing relationships among 80 different haplotypes of control region (Dloop) or Dloop-like sequences obtained from the Japanese spiny lobster (Panulirus japonicus), and control region sequences of P. japonicus and P. cygnus derived from the GenBank database. Haplotypes detected from the same lobster individual share the same color. Genuine mtDNA haplotype is shown in bold italic and number of clones examined is shown in parenthesis. Haplotypes carrying dagger differ from corresponding genuine mtDNA haplotype by less than two nucleotides (including indel). The bootstrap values greater than 60% (out of 1000 replicates) are shown at the nodes.Full size imageSequence size of haplotypes in groups B to D largely varied from 736 to 853 bp. K2P distances between group A and others ranged from 14.748 ± 1.030% (A vs B) to 61.619 ± 3.045% (A vs D), whereas that between haplotypes of group A and a reference sequence of P. japonicus was much smaller (6.333 ± 0.663%). BLAST homology search revealed no identical sequence for haplotypes in groups B to D and indicated that the closest species for haplotypes in groups B and C was P. japonicus with low to moderate similarity (74–88% homology). On the other hand, no significantly similar sequence was found for haplotypes in group D. Therefore, all 31 haplotypes (LC605788‒LC605815, LC654675-LC654677) in groups B to D were determined to be NUMTs.Impact of heteroplasmy and NUMTs for direct nucleotide sequencingPartial electropherogram obtained by direct nucleotide sequencing for COI amplicon of PJK3 is shown in Fig. 4 (top). Peak signals of this electropherogram are readable, but there are a number of sites where two (asterisk) or three (dagger) signals overlap. Alignment of a genuin haplotype (PJK-C7) and nine NUMTs sequences, corresponding to this partial electropherogram, is shown in Fig. 4 (bottom). At the sites where plural peaks overlap, different NUMT haplotypes were observed to share the same nucleotide different from the PJK3-direct. Heteroplasmic copies in COI determined in this study may have little negative impact on direct nucleotide sequencing, since nucleotides different from the genuine haplotypes were all unique to each heteroplasmic haplotype. Thus, the plural peaks at a site were composed of signals from genuine plus NUMT haplotypes, and the intensity of each peak was positively related to the copy numbers of these haplotypes. Frequent failure to obtain readable electropherograms in 12S and Dloop regions by direct sequencing may be due to extensive indels observed in the NUMT haplotypes.Figure 4A part of electropherogram obtained by direct nucleotide sequencing for COI region of PJK3 (top), and corresponding sequences from genuine haplotype (PJK3-C7) and nine NUMT haplotypes (see Fig. 1) are aligned (bottom). Apparent double (asterisk) and triple (dagger) peaks are observed at seven and five sites, respectively, which are comprised of signals from genuine and NUMT haplotypes.Full size image More

  • in

    Marine phytoplankton functional types exhibit diverse responses to thermal change

    1.Field, C. B., Behrenfeld, M. J., Randerson, J. T. & Falkowski, P. Primary production of the biosphere: Integrating terrestrial and cceanic components. Science 281, 237–240 (1998).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    2.Falkowski, P. G., Barber, R. T. & Smetacek, V. Biogeochemical controls and feedbacks on ocean primary production. Science 281, 200–206 (1998).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    3.Deutsch, C. A. et al. Impacts of climate warming on terrestrial ectotherms across latitude. Proc. Natl Acad. Sci. USA 105, 6668–6672 (2008).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    4.Comte, L. & Olden, J. D. Climatic vulnerability of the world’s freshwater and marine fishes. Nat. Clim. Chang. 7, 718–722 (2017).ADS 
    Article 

    Google Scholar 
    5.Dutkiewicz, S., Scott, J. R. & Follows, M. J. Winners and losers: ecological and biogeochemical changes in a warming ocean. Glob. Biogeochem. Cycles 27, 463–477 (2013).ADS 
    CAS 
    Article 

    Google Scholar 
    6.Sarmiento, J. L. et al. Response of ocean ecosystems to climate warming. Glob. Biogeochem. Cycles 18, GB3003 (2004).ADS 
    Article 
    CAS 

    Google Scholar 
    7.Taucher, J. & Oschlies, A. Can we predict the direction of marine primary production change under global warming? Geophys. Res. Lett. 38, 1–6 (2011).Article 
    CAS 

    Google Scholar 
    8.Vallina, S. M., Cermeno, P., Dutkiewicz, S., Loreau, M. & Montoya, J. M. Phytoplankton functional diversity increases ecosystem productivity and stability. Ecol. Modell. 361, 184–196 (2017).Article 

    Google Scholar 
    9.Dutkiewicz, S. et al. Impact of ocean acidification on the structure of future phytoplankton communities. Nat. Clim. Chang. 5, 1002–1006 (2015).ADS 
    CAS 
    Article 

    Google Scholar 
    10.Laufkotter, C. et al. Drivers and uncertainties of future global marine primary production in marine ecosystem models. Biogeosciences 12, 6955–6984 (2015).ADS 
    Article 

    Google Scholar 
    11.Behrenfeld, M. J., Boss, E., Siegel, D. A. & Shea, D. M. Carbon-based ocean productivity and phytoplankton physiology from space. Glob. Biogeochem. Cycles 19, 1–14 (2005).Article 
    CAS 

    Google Scholar 
    12.Anderson, S. I. & Rynearson, T. A. Variability approaching the thermal limits can drive diatom community dynamics. Limnol. Oceanogr. 65, 1961–1973 (2020).ADS 
    CAS 
    Article 

    Google Scholar 
    13.Boyd, P. W. Physiology and iron modulate diverse responses of diatoms to a warming Southern Ocean. Nat. Clim. Chang. 9, 148–152 (2019).ADS 
    CAS 
    Article 

    Google Scholar 
    14.Thomas, M. K. & Litchman, E. Effects of temperature and nitrogen availability on the growth of invasive and native cyanobacteria. Hydrobiologia 763, 357–369 (2016).Article 

    Google Scholar 
    15.Kremer, C. T., Thomas, M. K. & Litchman, E. Temperature- and size-scaling of phytoplankton population growth rates: Reconciling the Eppley curve and the metabolic theory of ecology. Limnol. Oceanogr. 62, 1658–1670 (2017).ADS 
    Article 

    Google Scholar 
    16.Edwards, K. F., Thomas, M. K., Klausmeier, C. A. & Litchman, E. Allometric scaling and taxonomic variation in nutrient utilization traits and maximum growth rate of phytoplankton. Limnol. Oceanogr. 57, 554–566 (2012).ADS 
    Article 

    Google Scholar 
    17.Poloczanska, E. S. et al. Global imprint of climate change on marine life. Nat. Clim. Chang. 3, 919–925 (2013).ADS 
    Article 

    Google Scholar 
    18.Thomas, M. K., Kremer, C. T., Klausmeier, C. A. & Litchman, E. A global pattern of thermal adaptation in marine phytoplankton. Science 338, 1085–1088 (2012).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    19.Righetti, D., Vogt, M., Gruber, N., Psomas, A. & Zimmermann, N. E. Global pattern of phytoplankton diversity driven by temperature and environmental variability. Sci. Adv. 5, 1–11 (2019).Article 

    Google Scholar 
    20.Barton, A. D., Irwin, A. J., Finkel, Z. V. & Stock, C. A. Anthropogenic climate change drives shift and shuffle in North Atlantic phytoplankton communities. Proc. Natl Acad. Sci. USA 113, 2964–2969 (2016).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    21.García Molinos, J. et al. Climate velocity and the future global redistribution of marine biodiversity. Nat. Clim. Chang. 6, 4–11 (2015).
    Google Scholar 
    22.Uitz, J., Claustre, H., Gentili, B. & Stramski, D. Phytoplankton class-specific primary production in the world’s oceans: Seasonal and interannual variability from satellite observations. Glob. Biogeochem. Cycles 24, 1–19 (2010).Article 
    CAS 

    Google Scholar 
    23.Toseland, A. et al. The impact of temperature on marine phytoplankton resource allocation and metabolism. Nat. Clim. Chang. 3, 979–984 (2013).ADS 
    CAS 
    Article 

    Google Scholar 
    24.Boyd, P. W. & Hutchins, D. A. Understanding the responses of ocean biota to a complex matrix of cumulative anthropogenic change. Mar. Ecol. Prog. Ser. 470, 125–135 (2012).ADS 
    Article 

    Google Scholar 
    25.Bopp, L. et al. Multiple stressors of ocean ecosystems in the 21st century: Projections with CMIP5 models. Biogeosciences 10, 6225–6245 (2013).ADS 
    Article 

    Google Scholar 
    26.Thomas, M. K., Kremer, C. T. & Litchman, E. Environment and evolutionary history determine the global biogeography of phytoplankton temperature traits. Glob. Ecol. Biogeogr. 25, 75–86 (2016).Article 

    Google Scholar 
    27.Angilletta, M. J. Thermal Adaptation: A Theoretical and Empirical Synthesis (Oxford University Press, 2009).28.Eppley, R. W. Temperature and phytoplankton growth in the sea. Fish. Bull. 70, 1063–1085 (1972).
    Google Scholar 
    29.Bissinger, J. E., Montagnes, D. J. S., Sharples, J. & Atkinson, D. Predicting marine phytoplankton maximum growth rates from temperature: Improving on the Eppley curve using quantile regression. Limnol. Oceanogr. 53, 487–493 (2008).ADS 
    Article 

    Google Scholar 
    30.Prowe, A. E. F., Pahlow, M., Dutkiewicz, S. & Oschlies, A. How important is diversity for capturing environmental-change responses in ecosystem models? Biogeosciences 11, 3397–3407 (2014).ADS 
    Article 

    Google Scholar 
    31.Chen, B. & Liu, H. Relationships between phytoplankton growth and cell size in surface oceans: Interactive effects of temperature, nutrients, and grazing. Limnol. Oceanogr. 55, 965–972 (2010).ADS 
    CAS 
    Article 

    Google Scholar 
    32.Barton, S. & Yvon‐Durocher, G. Quantifying the temperature dependence of growth rate in marine phytoplankton within and across species. Limnol. Oceanogr. 64, 2081–2091 (2019).ADS 
    Article 

    Google Scholar 
    33.Sherman, E., Moore, J. K., Primeau, F. & Tanouye, D. Temperature influence on phytoplankton community growth rates. Glob. Biogeochem. Cycles 30, 550–559 (2016).ADS 
    CAS 
    Article 

    Google Scholar 
    34.Alexander, H. et al. Functional group-specific traits drive phytoplankton dynamics in the oligotrophic ocean. Proc. Natl Acad. Sci. USA 112, E5972–E5979 (2015).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    35.Cermeño, P. et al. The role of nutricline depth in regulating the ocean carbon cycle. Proc. Natl Acad. Sci. USA 105, 20344–20349 (2008).ADS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    36.Calvo, E., Pelejero, C., Pena, L. D., Cacho, I. & Logan, G. A. Eastern Equatorial Pacific productivity and related-CO2 changes since the last glacial period. Proc. Natl Acad. Sci. USA 108, 5537–5541 (2011).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    37.McCabe, R. M. et al. An unprecedented coastwide toxic algal bloom linked to anomalous ocean conditions. Geophys. Res. Lett. 43, 10,366–10,376 (2016).Article 

    Google Scholar 
    38.Roberts, S. D., Van Ruth, P. D., Wilkinson, C., Bastianello, S. S. & Bansemer, M. S. Marine heatwave, harmful algae blooms and an extensive fish kill event during 2013 in South Australia. Front. Mar. Sci. 6, 1–20 (2019).CAS 
    Article 

    Google Scholar 
    39.Oliver, E. C. J. et al. Longer and more frequent marine heatwaves over the past century. Nat. Commun. 9, 1–12 (2018).CAS 
    Article 

    Google Scholar 
    40.Oliver, E. C. J. et al. Projected marine heatwaves in the 21st century and the potential for ecological impact. Front. Mar. Sci. 6, 1–12 (2019).MathSciNet 
    Article 

    Google Scholar 
    41.Keeling, P. J. The endosymbiotic origin, diversification and fate of plastids. Philos. Trans. R. Soc. B Biol. Sci. 365, 729–748 (2010).CAS 
    Article 

    Google Scholar 
    42.Yoon, H. S., Hackett, J. D., Pinto, G. & Bhattacharya, D. The single, ancient origin of chromist plastids. Proc. Natl Acad. Sci. USA 99, 15507–15512 (2002).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    43.Pinsky, M. L., Eikeset, A. M., McCauley, D. J., Payne, J. L. & Sunday, J. M. Greater vulnerability to warming of marine versus terrestrial ectotherms. Nature 569, 108–111 (2019).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    44.Sunday, J. M. et al. Thermal-safety margins and the necessity of thermoregulatory behavior across latitude and elevation. Proc. Natl Acad. Sci. USA 111, 5610–5615 (2014).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    45.Jönsson, B. F. & Watson, J. R. The timescales of global surface-ocean connectivity. Nat. Commun. 7, 1–6 (2016).Article 
    CAS 

    Google Scholar 
    46.Doblin, M. A. & van Sebille, E. Drift in ocean currents impacts intergenerational microbial exposure to temperature. Proc. Natl Acad. Sci. USA 113, 5700–5705 (2016).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    47.Whittaker, K. & Rynearson, T. Evidence for environmental and ecological selection in a microbe with no geographic limits to gene flow. Proc. Natl Acad. Sci. USA 114, 2651–2656 (2017).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    48.Ward, B. A., Cael, B. B., Collins, S. & Robert Young, C. Selective constraints on global plankton dispersal. Proc. Natl Acad. Sci. USA 118, 1–7 (2021).
    Google Scholar 
    49.Huey, R. B. & Stevenson, R. D. Integrating thermal physiology and ecology of ectotherms: A discussion of approaches. Integr. Comp. Biol. 19, 357–366 (1979).
    Google Scholar 
    50.Collins, M. et al. in Climate change 2013: The physical science basis. Contribution of working group I to the fifth assessment report of the Intergovernmental Panel on Climate Change (eds. Stocker, T. F. et al.) 1029–1136 (Cambridge University Press, 2013).51.Bopp, L., Aumont, O., Cadule, P., Alvain, S. & Gehlen, M. Response of diatoms distribution to global warming and potential implications: A global model study. Geophys. Res. Lett. 32, L19606 (2005).ADS 
    Article 
    CAS 

    Google Scholar 
    52.Ward, B. A. Temperature-correlated changes in phytoplankton community structure are restricted to polar waters. PLoS ONE 10, 1–15 (2015).
    Google Scholar 
    53.Winter, A., Henderiks, J., Beaufort, L., Rickaby, R. E. M. & Brown, C. W. Poleward expansion of the coccolithophore Emiliania huxleyi. J. Plankton Res. 36, 316–325 (2014).CAS 
    Article 

    Google Scholar 
    54.Rivero-Calle, S., Gnanadesikan, A., Del Castillo, C. E., Balch, W. M. & Guikema, S. D. Multidecadal increase in North Atlantic coccolithophores and the potential role of rising CO2. Science 350, 1533–1537 (2015).ADS 
    CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    55.Steinacher, M. et al. Projected 21st century decrease in marine productivity: a multi-model analysis. Biogeosciences Discuss. 7, 979–1005 (2010).ADS 
    CAS 
    Article 

    Google Scholar 
    56.Arrigo, K. R., van Dijken, G. L. & Strong, A. L. Environmental controls of marine productivity hot spots around Antarctica. J. Geophys. Res. Ocean. 120, 2813–2825 (2015).Article 

    Google Scholar 
    57.Aranguren-Gassis, M., Kremer, C. T., Klausmeier, C. A. & Litchman, E. Nitrogen limitation inhibits marine diatom adaptation to high temperatures. Ecol. Lett. 22, 1860–1869 (2019).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    58.Edwards, K. F., Thomas, M. K., Klausmeier, C. A. & Litchman, E. Phytoplankton growth and the interaction of light and temperature: A synthesis at the species and community level. Limnol. Oceanogr. 61, 1232–1244 (2016).ADS 
    Article 

    Google Scholar 
    59.Ibarbalz, F. M. et al. Global trends in marine plankton diversity across kingdoms of life. Cell 179, 1084–1097.e21 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    60.Allen, A. P., Gillooly, J. F., Savage, V. M. & Brown, J. H. Kinetic effects of temperature on rates of genetic divergence and speciation. Proc. Natl Acad. Sci. USA 103, 9130–9135 (2006).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    61.Padfield, D., Yvon-Durocher, G., Buckling, A., Jennings, S. & Yvon-Durocher, G. Rapid evolution of metabolic traits explains thermal adaptation in phytoplankton. Ecol. Lett. 19, 133–142 (2016).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    62.Baker, K. G. et al. Thermal niche evolution of functional traits in a tropical marine phototroph. J. Phycol. 54, 799–810 (2018).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    63.O’Donnell, D. R. et al. Rapid thermal adaptation in a marine diatom reveals constraints and trade-offs. Glob. Chang. Biol. 24, 4554–4565 (2018).ADS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    64.Seong, K. A., Jeong, H. J., Kim, S., Kim, G. H. & Kang, J. H. Bacterivory by co-occurring red-tide algae, heterotrophic nanoflagellates, and ciliates. Mar. Ecol. Prog. Ser. 322, 85–97 (2006).ADS 
    Article 

    Google Scholar 
    65.Arizona Software Inc. GraphClick 3.0.2. http://www.arizona-software.ch/graphclick/ (2010).66.Norberg, J. Biodiversity and ecosystem functioning: a complex adaptive systems approach. Limnol. Oceanogr. 49, 1269–1277 (2004).ADS 
    Article 

    Google Scholar 
    67.Bolker, B. & Team, R. D. C. bbmle: Tools for general maximum likelihood estimation. https://github.com/bbolker/bbmle (2017).68.R Core Team. R: A language and environment for statistical computing. https://www.R-project.org/ (2020).69.Riahi, K. et al. RCP 8.5-A scenario of comparatively high greenhouse gas emissions. Clim. Change 109, 33–57 (2011).ADS 
    CAS 
    Article 

    Google Scholar 
    70.Koenker, R. quantreg: Quantile regression. https://cran.r-project.org/package=quantreg (2019).71.Chen, B. & Laws, E. A. Is there a difference of temperature sensitivity between marine phytoplankton and heterotrophs? Limnol. Oceanogr. 62, 806–817 (2017).ADS 
    Article 

    Google Scholar 
    72.Sal, S., Alonso-Saez, L., Bueno, J., Garcıa, F. C. & Lopez-Urrutia, A. Thermal adaptation, phylogeny, and the unimodal size scaling of marine phytoplankton growth. Limnol. Oceanogr. 60, 1212–1221 (2015).ADS 
    Article 

    Google Scholar 
    73.Koenker, R. Quantile Regression, https://doi.org/10.1017/CBO9780511754098 (Cambridge University Press, 2005).74.Tomas, C. R. et al. Identifying Marine Phytoplankton. (Academic Press, 1997).75.He, X. & Hu, F. Markov chain marginal bootstrap. J. Am. Stat. Assoc. 97, 783–795 (2002).MathSciNet 
    MATH 
    Article 

    Google Scholar 
    76.Rynearson, T. A. Literature compilation of thermal growth rates from four phytoplankton functional types. Biological and Chemical Oceanography Data Management Office (BCO-DMO), (2021). https://doi.org/10.26008/1912/bco-dmo.839696.177.Rynearson, T. A. Estimated thermal capacities for phytoplankton strains. Biological and Chemical Oceanography Data Management Office (BCO-DMO), https://doi.org/10.26008/1912/bco-dmo.839713.1 (2021).78.Rynearson, T. A. Estimated thermal traits for phytoplankton. Biological and Chemical Oceanography Data Management Office (BCO-DMO), https://doi.org/10.26008/1912/bco-dmo.839689.1 (2021).79.Anderson, S. I. sianderson/PFT_thermal_response: Marine Phytoplankton Functional Types Exhibit Diverse Responses to Thermal Change. zenodo. https://doi.org/10.5281/zenodo.5507532 (2021).80.Buitenhuis, E. T., Pangerc, T., Franklin, D. J., Le Quéré, C. & Malin, G. Growth rates of six coccolithophorid strains as a function of temperature. Limnol. Oceanogr. 53, 1181–1185 (2008).ADS 
    Article 

    Google Scholar 
    81.Stawiarski, B., Buitenhuis, E. T. & Le Quéré, C. The physiological response of picophytoplankton to temperature and its model representation. Front. Mar. Sci. 3, 1–13 (2016).Article 

    Google Scholar  More