Ecology
Subterms
More stories
113 Shares169 Views
in EcologyQuantifying individual influence in leading-following behavior of Bechstein’s bats
Inferring leading-following networks
Defining leading-following events
Unlike studies on collective motion where group movement is tracked continuously5,15, our datasets contain only discrete records of bat appearances at experimental boxes. Quantifying individual influence is, thus, contingent on a rigorous method for inferring leading-following events from discrete recordings of animal occurrences. To denote the information that individuals possess about the location of experimental boxes, we refine the nomenclature used by Kerth and Reckardt3. An individual bat is said to be naïve at time ({{{mathbf {t}}}}_{{{mathbf {1}}}}) regarding a given box, if it has not been recorded by the reading device in that box for all times ({{mathbf {t}}} More113 Shares189 Views
in EcologyCryptic terrestrial fungus-like fossils of the early Ediacaran Period
1.
Kenrick, P., Wellman, C. H., Schneider, H. & Edgecombe, G. D. A timeline for terrestrialization: consequences for the carbon cycle in the Palaeozoic. Philos. Trans. R. Soc. B 367, 519–536 (2012).
Article Google Scholar
2.
Kennedy, M., Droser, M., Mayer, L. M., Pevear, D. & Mrofka, D. Late Precambrian oxygenation; inception of the clay mineral factory. Science 311, 1446–1449 (2006).
ADS CAS PubMed Article Google Scholar3.
Naranjo-Ortiz, M. A. & Gabaldón, T. Fungal evolution: major ecological adaptations and evolutionary transitions. Biol. Rev. 94, 1443–1476 (2019).
PubMed Article Google Scholar4.
Heckman, D. S. et al. Molecular evidence for the early colonization of land by fungi and plants. Science 293, 1129–1133 (2001).
CAS PubMed Article Google Scholar5.
Lutzoni, F. et al. Contemporaneous radiations of fungi and plants linked to symbiosis. Nat. Commun. 9, 5451 (2018).
ADS CAS PubMed PubMed Central Article Google Scholar6.
Chang, Y. et al. Phylogenomic analyses indicate that early fungi evolved digesting cell walls of algal ancestors of land plants. Genome Biol. Evol. 7, 1590–1601 (2015).
CAS PubMed PubMed Central Article Google Scholar7.
Taylor, T. N., Krings, M. & Taylor, E. L. Fossil Fungi. 1st edn (Academic Press, 2015).8.
Berbee, M. L. et al. Genomic and fossil windows into the secret lives of the most ancient fungi. Nat. Rev. Microbiol. 18, 717–730 (2020).
Google Scholar9.
Bengtson, S. et al. Fungus-like mycelial fossils in 2.4-billion-year-old vesicular basalt. Nat. Ecol. Evol. 1, 0141 (2017).
Article Google Scholar10.
Loron, C. C. et al. Early fungi from the Proterozoic era in Arctic Canada. Nature 570, 232–235 (2019).
ADS CAS PubMed Article Google Scholar11.
Bonneville, S. et al. Molecular identification of fungi microfossils in a Neoproterozoic shale rock. Sci. Adv. 6, eaax7599 (2020).
ADS CAS PubMed PubMed Central Article Google Scholar12.
Butterfield, N. J. Probable Proterozoic fungi. Paleobiology 31, 165–182 (2005).
Article Google Scholar13.
Yuan, X., Xiao, S. & Taylor, T. N. Lichen-like symbiosis 600 million years ago. Science 308, 1017–1020 (2005).
ADS CAS PubMed Article Google Scholar14.
Smith, M. R. Cord-forming Palaeozoic fungi in terrestrial assemblages. Bot. J. Linn. Soc. 180, 452–460 (2016).
Article Google Scholar15.
Krings, M., Harper, C. J. & Taylor, E. L. Fungi and fungal interactions in the Rhynie chert: a review of the evidence, with the description of Perexiflasca tayloriana gen. et sp. nov. Philos. Trans. R. Soc. B 373, 20160500 (2018).
Article Google Scholar16.
Condon, D. et al. U-Pb ages from the Neoproterozoic Doushantuo Formation, China. Science 308, 95–98 (2005).
ADS CAS PubMed Article PubMed Central Google Scholar17.
Zhou, C., Huyskens, M. H., Lang, X., Xiao, S. & Yin, Q.-Z. Calibrating the terminations of Cryogenian global glaciations. Geology 47, 251–254 (2019).
ADS CAS Article Google Scholar18.
Jiang, G., Kennedy, M. J., Christie-Blick, N., Wu, H. & Zhang, S. Stratigraphy, sedimentary structures, and textures of the late Neoproterozoic Doushantuo cap carbonate in South China. J. Sediment. Res. 76, 978–995 (2006).
ADS CAS Article Google Scholar19.
Hoffman, P. F. & Macdonald, F. A. Sheet-crack cements and early regression in Marinoan (635 Ma) cap dolostones: regional benchmarks of vanishing ice-sheets? Earth Planet. Sci. Lett. 300, 374–384 (2010).
ADS CAS Article Google Scholar20.
Gan, T. et al. Miniature paleo-speleothems from the earliest Ediacaran (635 Ma) Doushantuo cap dolostone in South China and their implications for terrestrial ecosystems. EarthArXiv, https://doi.org/10.31223/osf.io/srkcp (2019).21.
Zhou, C., Bao, H., Peng, Y. & Yuan, X. Timing the deposition of 17O-depleted barite at the aftermath of Nantuo glacial meltdown in South China. Geology 38, 903–906 (2010).
ADS CAS Article Google Scholar22.
Zhou, G., Luo, T., Zhou, M., Xing, L. & Gan, T. A ubiquitous hydrothermal episode recorded in the sheet-crack cements of a Marinoan cap dolostone of South China: implication for the origin of the extremely 13C-depleted calcite cement. J. Asian Earth Sci. 134, 63–71 (2017).
ADS Article Google Scholar23.
Muscente, A. D., Czaja, A. D., Tuggle, J., Winkler, C. & Xiao, S. Manganese oxides resembling microbial fabrics and their implications for recognizing inorganically preserved microfossils. Astrobiology 18, 249–258 (2018).
ADS CAS PubMed Article Google Scholar24.
García-Ruiz, J. M. et al. Self-assembled silica-carbonate structures and detection of ancient microfossils. Science 302, 1194 (2003).
ADS PubMed Article CAS Google Scholar25.
Rouillard, J., García-Ruiz, J. M., Gong, J. & van Zuilen, M. A. A morphogram for silica-witherite biomorphs and its application to microfossil identification in the early earth rock record. Geobiology 16, 279–296 (2018).
CAS PubMed PubMed Central Article Google Scholar26.
Hofmann, B. A., Farmer, J. D., Blanckenburg, F. V. & Fallick, A. E. Subsurface filamentous fabrics: an evaluation of origins based on morphological and geochemical criteria, with implications for exopaleontology. Astrobiology 8, 87–117 (2008).
ADS CAS PubMed Article Google Scholar27.
Rasmussen, B. Filamentous microfossils in a 3,235-million-year-old volcanogenic massive sulphide deposit. Nature 405, 676–679 (2000).
ADS CAS PubMed Article Google Scholar28.
Schopf, J. W. et al. Sulfur-cycling fossil bacteria from the 1.8-Ga Duck Creek Formation provide promising evidence of evolution’s null hypothesis. Proc. Natl Acad. Sci. USA 112, 2087–2092 (2015).
ADS CAS PubMed Article Google Scholar29.
Teske, A. & Nelson, D. C. in The Prokaryotes: Volume 6: Proteobacteria: Gamma Subclass (eds Dworkin, M. et al.) 784–810 (Springer, 2006).30.
Zhou, X. et al. Biogenic iron-rich filaments in the quartz veins in the uppermost Ediacaran Qigebulake Formation, Aksu area, northwestern Tarim Basin, China: implications for iron oxidizers in subseafloor hydrothermal systems. Astrobiology 15, 523–537 (2015).
ADS CAS PubMed Article Google Scholar31.
Thurston, E. L. & Ingram, L. O. Morphology and fine structure of Fischerella ambigua. J. Phycol. 7, 203–210 (1971).
Google Scholar32.
Iyengar, M. O. P. & Desikachary, T. V. Mastigocladopsis jogensis gen. et sp. nov., a new member of the stigonemataceæ. Proc. Ind. Acad. Sci. B 24, 55–59 (1946).
Google Scholar33.
Komárek, J. Cyanoprokaryota: 3. Teil/Part 3: Heterocytous Genera. (Springer Spektrum, 2013).34.
Castenholz, R. W. in Bergey’s Manual of Systematic Bacteriology (eds Boone, et al.) 473–599 (Springer, 2001).35.
Bartley, J. K. Actualistic taphonomy of cyanobacteria: implications for the Precambrian fossil record. Palaios 11, 571–586 (1996).
ADS Article Google Scholar36.
Bold, H. C. & Wynne, M. J. Introduction to the Algae: Structure and Reproduction. (Prentice-Hall, 1978).37.
Butterfield, N. J. A vaucheriacean alga from the middle Neoproterozoic of Spitsbergen: implications for the evolution of Proterozoic eukaryotes and the Cambrian explosion. Paleobiology 30, 231–252 (2004).
Article Google Scholar38.
Tang, Q., Pang, K., Yuan, X. & Xiao, S. A one-billion-year-old multicellular chlorophyte. Nat. Ecol. Evol. 4, 543–549 (2020).
PubMed Article Google Scholar39.
Leliaert, F. & Coppejans, E. A revision of Cladophoropsis Børgesen (Siphonocladales, Chlorophyta). Phycologia 45, 657–679 (2006).
Article Google Scholar40.
Zhao, Z.-J., Zhu, H., Hu, Z.-Y. & Liu, G.-X. Occurrence of true branches in Rhizoclonium (Cladophorales, Ulvophyceae) and the reinstatement of Rhizoclonium pachydermum Kjellman. Phytotaxa 166, 273–284 (2014).
Article Google Scholar41.
Entwisle, T. J. A monograph of Vaucheria (Vaucheriaceae, Chrysophyta) in south-eastern mainland Australia. Aust. Syst. Bot. 1, 1–77 (1988).
Article Google Scholar42.
Boo, S. M. & Cho, T. O. The Morphology of Griffithsia tomo-yamadae Okamura (Ceramiaceae, Rhodophyta): a little-known species from the northeast Pacific. Bot. Mar. 44, 109–118 (2001).
Article Google Scholar43.
Ferrer, N. C. & Caceres, E. J. Spirogyra salmonispora sp. nov. (Zygnematophyceae, Chiorophyta), a new freshwater species of the section Conjugata. Arch. Protistenk. 146, 101–105 (1995).
Article Google Scholar44.
Li, Q., Chen, X., Jiang, Y. & Jiang, C. in Actinobacteria: Basics and Biotechnological Applications (eds Dhanasekaran, D. & Jiang, Y.) 59–86 (IntechOpen, 2016).45.
Goodfellow, M. et al. Bergey’s Manual of Systematic Bacteriology: Volume Five The Actinobacteria, Part A and B. (Springer-Verlag, 2012).46.
Erikson, D. The morphology, cytology, and taxonomy of the Actinomycetes. Annu. Rev. Microbiol. 3, 23–54 (1949).
Article Google Scholar47.
Gregory, K. F. Hyphal anastomosis and cytological aspects of Streptomyces scabies. Can. J. Microbiol. 2, 649–655 (1956).
Article Google Scholar48.
Higgins, M. L. & Silvey, J. K. G. Slide culture observations of two freshwater Actinomycetes. Trans. Am. Micros. Soc. 85, 390–398 (1966).
CAS Article Google Scholar49.
Spatafora, J. W. et al. A phylum-level phylogenetic classification of zygomycete fungi based on genome-scale data. Mycologia 108, 1028–1046 (2016).
CAS PubMed PubMed Central Article Google Scholar50.
O’Donnell, K. L. Zygomycetes in Culture. (Department of Botany, University of Georgia, 1979).51.
Fischer, M. S. & Glass, N. L. Communicate and fuse: how filamentous fungi establish and maintain an interconnected mycelial network. Front. Microbiol. 10, 619 (2019).
PubMed PubMed Central Article Google Scholar52.
Webster, J. in Introduction to Fungi (3rd Edn.) (eds Webster, J. & Weber, R.) 165–225 (Cambridge University Press, 2007).53.
Drake, H. et al. Anaerobic consortia of fungi and sulfate reducing bacteria in deep granite fractures. Nat. Commun. 8, 55 (2017).
ADS PubMed PubMed Central Article CAS Google Scholar54.
Bengtson, S. et al. Deep-biosphere consortium of fungi and prokaryotes in Eocene subseafloor basalts. Geobiology 12, 489–496 (2014).
CAS PubMed Article Google Scholar55.
Ivarsson, M. et al. Fossilized fungi in subseafloor Eocene basalts. Geology 40, 163–166 (2012).
ADS CAS Article Google Scholar56.
Northup, D. et al. Biological investigations in Lechuguilla Cave. NSS Bull. 56, 54–63 (1994).
Google Scholar57.
Duane, M. J. Unusual preservation of crustaceans and microbial colonies in a vadose zone, northwest Morocco. Lethaia 36, 21–32 (2003).
Article Google Scholar58.
Kretzschmar, M. Fossile pilze in eisen-stromatolithen von warstein (rheinisches schiefergebirge). Facies 7, 237–259 (1982).
Article Google Scholar59.
Nieves-Rivera, Á. M., Santos-Flores, C. J., Dugan, F. M. & Miller, T. E. Guanophilic fungi in three caves of southwestern Puerto Rico. Int. J. Speleol. 38, 61–70 (2009).
Article Google Scholar60.
Nováková, A. Microscopic fungi isolated from the Domica Cave system (Slovak Karst National Park, Slovakia). A review. Int. J. Speleol. 38, 71–82 (2009).
Article Google Scholar61.
Popović, S. et al. Cyanobacteria, algae and microfungi present in biofilm from Božana Cave (Serbia). Int. J. Speleol. 44, 141–149 (2015).
Article Google Scholar62.
Schopf, J. W. Microflora of the Bitter Springs Formation, late Precambrian, central Australia. J. Paleontol. 42, 651–688 (1968).
Google Scholar63.
Strother, P. K. Systematics and evolutionary significance of some new cryptospores from the Cambrian of eastern Tennessee, USA. Rev. Palaeobot. Palynol. 227, 28–41 (2016).
Article Google Scholar64.
Prave, A. R. Life on land in the Proterozoic: evidence from the Torridonian rocks of northwest Scotland. Geology 30, 811–814 (2002).
ADS Article Google Scholar65.
Blank, C. E. Origin and early evolution of photosynthetic eukaryotes in freshwater environments: reinterpreting Proterozoic paleobiology and biogeochemical processes in light of trait evolution. J. Phycol. 49, 1040–1055 (2013).
CAS PubMed Article Google Scholar66.
Sánchez-Baracaldo, P., Raven, J. A., Pisani, D. & Knoll, A. H. Early photosynthetic eukaryotes inhabited low-salinity habitats. Proc. Natl. Acad. Sci. USA 114, E7737–E7745 (2017).
PubMed Article CAS Google Scholar67.
Föllmi, K. B. The phosphorus cycle, phosphogenesis and marine phosphate-rich deposits. Earth Sci. Rev. 40, 55–124 (1996).
ADS Article Google Scholar68.
Sahoo, S. K. et al. Ocean oxygenation in the wake of the Marinoan glaciation. Nature 489, 546–549 (2012).
ADS CAS PubMed Article Google Scholar69.
Guo, Z., Peng, X., Czaja, A. D., Chen, S. & Ta, K. Cellular taphonomy of well-preserved Gaoyuzhuang microfossils: a window into the preservation of ancient cyanobacteria. Precambrian Res. 304, 88–98 (2018).
ADS CAS Article Google Scholar70.
Czaja, A. D., Beukes, N. J. & Osterhout, J. T. Sulfur-oxidizing bacteria prior to the Great Oxidation Event from the 2.52 Ga Gamohaan Formation of South Africa. Geology 44, 983–986 (2016).
ADS CAS Article Google Scholar71.
Pang, K. et al. The nature and origin of nucleus-like intracellular inclusions in Paleoproterozoic eukaryote microfossils. Geobiology 11, 499–510 (2013).
CAS PubMed Google Scholar72.
Zhang, J. et al. Improved precision and spatial resolution of sulfur isotope analysis using NanoSIMS. J. Anal. Spectrom. 20, 1934–1943 (2014).
CAS Article Google Scholar73.
Chen, L. et al. Extreme variation of sulfur isotopic compositions in pyrite from the Qiuling sediment-hosted gold deposit, West Qinling orogen, central China: an in situ SIMS study with implications for the source of sulfur. Mineral. Depos. 50, 643–656 (2015).
ADS CAS Article Google Scholar74.
Roberts, N. M. W. & Walker, R. J. U-Pb geochronology of calcite-mineralized faults: absolute timing of rift-related fault events on the northeast Atlantic margin. Geology 44, 531–534 (2016).
ADS CAS Article Google Scholar75.
Roberts, N. M. W. et al. A calcite reference material for LA-ICP-MS U-Pb geochronology. Geochem. Geophys. Geosyst. 18, 2807–2814 (2017).
ADS CAS Article Google Scholar76.
Hu, Z. et al. Signal enhancement in laser ablation ICP-MS by addition of nitrogen in the central channel gas. J. Anal. Spectrom. 23, 1093–1101 (2008).
CAS Article Google Scholar77.
Liu, Y. et al. Reappraisement and refinement of zircon U-Pb isotope and trace element analyses by LA-ICP-MS. Chin. Sci. Bull. 55, 1535–1546 (2010).
CAS Article Google Scholar78.
Zhang, Y. & Yuan, X. New data on multicellular thallophytes and fragments of cellular tissues from late Proterozoic phosphate rocks, South China. Lethaia 25, 1–18 (1992).
Article Google Scholar79.
Nie, W., Ma, D., Pan, J., Zhou, J. & Wu, K. δ13C excursions of phosphorite-bearing rocks in Neoproterozoic-Early Cambrian interval in Guizhou, South China: implications for palaeoceanic evolutions. J. Nanjing Univ. Nat. Sci. 42, 257–268 (2006).
CAS Google Scholar80.
Barfod, G. H. et al. New Lu-Hf and Pb-Pb age constraints on the earliest animal fossils. Earth Planet. Sci. Lett. 201, 203–212 (2002).
ADS CAS Article Google Scholar81.
Igisu, M. et al. Micro-FTIR spectroscopic signatures of bacterial lipids in Proterozoic microfossils. Precambrian Res. 173, 19–26 (2009).
ADS CAS Article Google Scholar82.
Wang, X.-H. Interfacial electrochemistry of pyrite oxidation and flotation: II. FTIR studies of xanthate adsorption on pyrite surfaces in neutral pH solutions. J. Colloid Interface Sci. 171, 413–428 (1995).
ADS CAS Article Google Scholar83.
Igisu, M. et al. FTIR microspectroscopy of Ediacaran phosphatized microfossils from the Doushantuo Formation, Weng’an, South China. Gondwana Res. 25, 1120–1138 (2014).
ADS CAS Article Google Scholar84.
Beyssac, O., Goffé, B., Chopin, C. & Rouzaud, J. N. Raman spectra of carbonaceous material in metasediments: a new geothermometer. J. Metamorphic. Geol. 20, 859–871 (2002).
ADS CAS Article Google Scholar85.
Turcotte, S. B. et al. Application of Raman spectroscopy to metal-sulfide surface analysis. Appl. Opt. 32, 935–938 (1993).
ADS CAS PubMed Article Google Scholar More200 Shares109 Views
in EcologyNovel gene rearrangement in the mitochondrial genome of Muraenesox cinereus and the phylogenetic relationship of Anguilliformes
1.
Mehta, R. S. Ecomorphology of the moray bite: Relationship between dietary extremes and morphological diversity. Physiol. Biochem. Zool. 82, 90–103 (2009).
PubMed Article ADS PubMed Central Google Scholar
2.
Mehta, R. S. & Wainwright, P. C. Raptorial jaws in the throat help moray eels swallow large prey. Nature 449, 79–82 (2007).
CAS PubMed Article ADS PubMed Central Google Scholar3.
Robins, C. R. The phylogenetic relationships of the anguilliform fishes. Fishes Western N. Atl. 1, 9–23 (1989).
Google Scholar4.
Greenwood, P. H. Notes on the anatomy and classification of elopomorph fishes. Bull. Mus. Comp. Zool. 32, 65–102 (1977).5.
Nelson, G. J. Relationships of clupeomorphs, with remarks on the structure of the lower jaw in fishes. Interrelat. Fishes 333–349 (1973).6.
Inoue, J. G. et al. Deep-ocean origin of the freshwater eels. Biol. Let. 6, 363–366. https://doi.org/10.1098/rsbl.2009.0989 (2010).
Article Google Scholar7.
Santini, F. et al. A multi-locus molecular timescale for the origin and diversification of eels (Order: Anguilliformes). Mol. Phylogenet. Evol. 69, 884–894. https://doi.org/10.1016/j.ympev.2013.06.016 (2013).
CAS Article PubMed PubMed Central Google Scholar8.
Inoue, J. G., Miya, M., Tsukamoto, K. & Nishida, M. Complete Mitochondrial DNA Sequence of Conger myriaster (Teleostei: Anguilliformes): Novel gene order for vertebrate mitochondrial genomes and the phylogenetic implications for Anguilliform Families. J. Mol. Evol. 52, 311–320 (2001).
CAS PubMed Article ADS PubMed Central Google Scholar9.
Tang, K. L. & Fielitz, C. Phylogeny of moray eels (Anguilliformes: Muraenidae), with a revised classification of true eels (Teleostei: Elopomorpha: Anguilliformes). Mitochondrial DNA. 24, 55–66 (2013).
CAS PubMed Article PubMed Central Google Scholar10.
Russell, B. & Houston, W. Offshore fishes of the Arafura Sea. Beagle Rec. Mus. Art Galleries Northern Territory 6, 69–84 (1989).
Google Scholar11.
Chen, D., Ye, Y., Chen, J., Zhan, P. & Lou, Y. Molecular nutritional characteristics of vinasse pike eel (Muraenesox cinereus) during pickling. Food Chem. 224, 359–364. https://doi.org/10.1016/j.foodchem.2016.12.089 (2017).
CAS Article PubMed PubMed Central Google Scholar12.
Boore, J. L. Animal mitochondrial genomes. Nucleic Acids Res. 27, 1767–1780 (1999).
CAS PubMed PubMed Central Article Google Scholar13.
Lu, Z. Z. et al. Complete mitochondrial genome of Ophichthus brevicaudatus reveals novel gene order and phylogenetic relationships of Anguilliformes. Int. J. Biol. Macromol. 135, 609–618. https://doi.org/10.1016/j.ijbiomac.2019.05.139 (2019).
CAS Article PubMed PubMed Central Google Scholar14.
Bibb, M., Etten, R., Wright, C., Walberg, M. & Clayton, D. Sequence and gene organization of mouse mitochondrial DNA. Cell 26, 167–180. https://doi.org/10.1016/0092-8674(81)90300-7 (1981).
CAS Article PubMed PubMed Central Google Scholar15.
Anderson, S. B. et al. Sequence and organization of the human mitochondrial genome. Nature 290, 457–465. https://doi.org/10.1038/290457a0 (1981).
CAS Article PubMed ADS PubMed Central Google Scholar16.
Roe, B. A., Ma, D. P., Wilson, R. K. & Wong, F. H. The complete nucleotide sequence of the Xenopus laevis mitochondrial genome. J. Biol. Chem. 260, 9759–9774 (1985).
CAS PubMed Article PubMed Central Google Scholar17.
Brown, W., George, M. J. & Wilson, A. C. Rapid evolution of animal mitochondrial DNA. Proc. Natl. Acad. Sci. USA 76, 1967–1971. https://doi.org/10.1073/pnas.76.4.1967 (1979).
CAS Article PubMed ADS PubMed Central Google Scholar18.
Macey, J. R., Larson, A., Ananjeva, N. B., Fang, Z. & Papenfuss, T. J. Two novel gene orders and the role of light-strand replication in rearrangement of the vertebrate mitochondrial genome. Mol. Biol. Evol. 14, 91–104. https://doi.org/10.1093/oxfordjournals.molbev.a025706 (1997).
CAS Article PubMed PubMed Central Google Scholar19.
Zhang, J. Y., Zhang, L. P., Yu, D. N., Storey, K. B. & Zheng, R. Q. Complete mitochondrial genomes of Nanorana taihangnica and N. yunnanensis (Anura: Dicroglossidae) with novel gene arrangements and phylogenetic relationship of Dicroglossidae. BMC Evol. Biol. 18, 1–13 (2018).
Article CAS Google Scholar20.
Yan, J., Li, H. & Zhou, K. Evolution of the mitochondrial genome in snakes: Gene rearrangements and phylogenetic relationships. BMC Genom. 9, 569. https://doi.org/10.1186/1471-2164-9-569 (2008).
CAS Article Google Scholar21.
Liu, J., Yu, J., Zhou, M. & Yang, J. Complete mitochondrial genome of Japalura flaviceps: Deep insights into the phylogeny and gene rearrangements of Agamidae species. Int. J. Biol. Macromol. 125, 423–431. https://doi.org/10.1016/j.ijbiomac.2018.12.068 (2019).
CAS Article PubMed PubMed Central Google Scholar22.
Verkuil, Y. I., Piersma, T. & Baker, A. J. A novel mitochondrial gene order in shorebirds (Scolopacidae, Charadriiformes). Mol. Phylogenet. Evol. 57, 411–416. https://doi.org/10.1016/j.ympev.2010.06.010 (2010).
CAS Article PubMed PubMed Central Google Scholar23.
Eberhard, J. R. & Wright, T. F. Rearrangement and evolution of mitochondrial genomes in parrots. Mol. Phylogenet. Evol. 94, 34–46 (2015).
PubMed PubMed Central Article CAS Google Scholar24.
Pääbo, S., Thomas, W. K., Whitfield, K. M., Kumazawa, Y. & Wilson, A. C. Rearrangements of mitochondrial transfer RNA genes in marsupials. J. Mol. Evol. 33, 426–430. https://doi.org/10.1007/bf02103134 (1991).
Article PubMed ADS PubMed Central Google Scholar25.
Gong, L., Shi, W., Yang, M., Li, D. & Kong, X. Novel gene arrangement in the mitochondrial genome of Bothus myriaster(Pleuronectiformes: Bothidae): Evidence for the dimer-mitogenome and non-random loss model. Mitochondrial DNA Part A. 27, 3089–3092 (2015).
Article CAS Google Scholar26.
Miya, M. N. Organization of the Mitochondrial Genome of a Deep-Sea Fish, Gonostoma gracile (Teleostei: Stomiiformes): First example of transfer RNA gene rearrangements in Bony Fishes. Mar. Biotechnol. 1, 416–0426 (1999).
CAS Article Google Scholar27.
Shi, W., Miao, X. G. & Kong, X. Y. A novel model of double replications and random loss accounts for rearrangements in the Mitogenome of Sssamariscus latus (Teleostei: Pleuronectiformes). BMC Genom. 15, 352 (2014).
Article CAS Google Scholar28.
Kong, X. et al. A novel rearrangement in the mitochondrial genome of tongue sole, Cynoglossus semilaevis: Control region translocation and a tRNA gene inversion. Genome. 52, 975–984. https://doi.org/10.1139/g09-069 (2009).
CAS Article PubMed PubMed Central Google Scholar29.
Gong, L., Shi, W., Si, L. Z. & Kong, X. Y. Rearrangement of mitochondrial genome in fishes. Zool. Res. 34, 666–673 (2013).
CAS Google Scholar30.
Inoue, J. G., Masaki, M., Katsumi, T. & Mutsumi, N. evolution of the deep-sea gulper eel mitochondrial genomes: Large-scale gene rearrangements originated within the eels. Mol. Biol. Evol. 20, 1917–1924 (2003).
CAS PubMed Article PubMed Central Google Scholar31.
Ishikawa, S., Kimura, Y., Tokai, T., Tsukamoto, K. & Nishida, M. Gene rearrangement around the control region in the mitochondrial genome of conger myriaster. Fish. Sci. 66, 1186–1188 (2002).
Article Google Scholar32.
Miya, M., Kawaguchi, A. & Nishida, M. Mitogenomic exploration of higher teleostean phylogenies: A case study for moderate-scale evolutionary genomics with 38 newly determined complete mitochondrial DNA sequences. Mol. Biol. Evol. 18, 1993–2009 (2001).
CAS PubMed Article PubMed Central Google Scholar33.
Miya, M. T. et al. Major patterns of higher teleostean phylogenies: A new perspective based on 100 complete mitochondrial DNA sequences. Mol. Phylogenet. Evol. 26, 121–138 (2003).
CAS PubMed Article PubMed Central Google Scholar34.
Poulton, J. et al. Families of mtDNA re-arrangements can be detected in patients with mtDNA deletions: Duplications may be a transient intermediate form. Hum. Mol. Genet. 2, 23–30 (1993).
CAS PubMed Article PubMed Central Google Scholar35.
Lunt, D. H. & Hyman, B. C. Animal mitochondrial DNA recombination. Nature 387, 247 (1997).
CAS PubMed Article ADS PubMed Central Google Scholar36.
Ladoukakis, E. D. & Zouros, E. Recombination in animal mitochondrial DNA: Evidence from published sequences. Mol. Biol. Evol. 18, 2127–2131 (2001).
CAS PubMed Article Google Scholar37.
Sammler, S., Bleidorn, C. & Tiedemann, R. Full mitochondrial genome sequences of two endemic Philippine hornbill species (Aves: Bucerotidae) provide evidence for pervasive mitochondrial DNA recombination. BMC Genom. 12, 35. https://doi.org/10.1186/1471-2164-12-35 (2011).
CAS Article Google Scholar38.
Atsushi, K. et al. Phylogeny, recombination, and mechanisms of stepwise mitochondrial genome reorganization in mantellid frogs from madagascar. Mol. Biol. Evol. 5, 874–891 (2008).
Google Scholar39.
Arndt, A. & Smith, M. J. Mitochondrial gene rearrangement in the sea cucumber genus Cucumaria. Mol. Biol. Evol. 8, 1009–1016 (1998).
Article Google Scholar40.
Moritz, C., Dowling, T. E. & Brown, W. M. Evolution of animal mitochondrial DNA: Relevance for population biology and systematics. Annu. Rev. Ecol. Syst. 18, 269–292 (1987).
Article Google Scholar41.
Erin, E. S. et al. Multiple independent origins of mitochondrial control region duplications in the order Psittaciformes. Mol. Phylogenet. Evol. 64, 342–356. https://doi.org/10.1016/j.ympev.2012.04.009 (2012).
Article Google Scholar42.
Mauro, D. S., Gower, D. J., Rafael, Z. & Mark, W. A hotspot of gene order rearrangement by tandem duplication and random loss in the vertebrate mitochondrial genome. Mol. Biol. Evol. 23, 227–234 (2006).
Article CAS Google Scholar43.
Lavrov, D. V., Boore, J. L. & Brown, W. M. Complete mtDNA sequences of two millipedes suggest a new model for mitochondrial gene rearrangements: Duplication and nonrandom loss. Mol. Biol. Evol. 19, 163–169 (2002).
CAS PubMed Article PubMed Central Google Scholar44.
Smith, M. J., Arndt, A., Gorski, S. & Fajber, E. The phylogeny of echinoderm classes based on mitochondrial gene arrangements. J. Mol. Evol. 36, 545–554 (1993).
CAS PubMed Article ADS PubMed Central Google Scholar45.
Schierup, M. H. & Hein, J. Consequences of recombination on traditional phylogenetic analysis. Genetics 156, 879–891 (2000).
CAS PubMed PubMed Central Google Scholar46.
Zhi, J. J. et al. Comparative mitochondrial genomics of snakes: Extraordinary substitution rate dynamics and functionality of the duplicate control region. BMC Evol. Biol. 7, 123. https://doi.org/10.1186/1471-2148-7-123 (2007).
CAS Article Google Scholar47.
Shi, W., Dong, X. L., Wang, Z. M., Miao, X. G. & Kong, X. Y. Complete mitogenome sequences of four flatfishes (Pleuronectiformes) reveal a novel gene arrangement of L-strand coding genes. BMC Evol. Biol. 13, 173 (2013).
PubMed PubMed Central Article CAS Google Scholar48.
Kumazawa, Y., Ota, H., Nishida, M. & Ozawa, T. The complete nucleotide sequence of a snake (Dinodon semicarinatus) mitochondrial genome with two identical control regions. Genetics 150, 313–329 (1998).
CAS PubMed PubMed Central Google Scholar49.
Liu, Y. et al. Mitochondrial genome of the yellow catfish Pelteobagrus fulvidraco and insights into Bagridae phylogenetics. Genomics 111, 1258–1265. https://doi.org/10.1016/j.ygeno.2018.08.005 (2019).
CAS Article PubMed PubMed Central Google Scholar50.
Gong, L., Lü, Z. M., Guo, B. Y., Ye, Y. Y. & Liu, L. Q. Characterization of the complete mitochondrial genome of the tidewater goby, Eucyclogobius newberryi (Gobiiformes; Gobiidae; Gobionellinae) and its phylogenetic implications. Conserv. Genet. Resour. 10, 93–97 (2017).
Article Google Scholar51.
Lin, J. P. et al. The first complete mitochondrial genome of the sand dollar Sinaechinocyamus mai (Echinoidea: Clypeasteroida). Genomics 112, 1686–1693. https://doi.org/10.1016/j.ygeno.2019.10.007 (2020).
CAS Article PubMed PubMed Central Google Scholar52.
Prabhu, V. R. et al. Characterization of the complete mitochondrial genome of Barilius malabaricus and its phylogenetic implications. Genomics 112, 2154–2163. https://doi.org/10.1016/j.ygeno.2019.12.009 (2020).
CAS Article PubMed PubMed Central Google Scholar53.
Xu, T. J., Cheng, Y. Z., Sun, Y. N., Shi, G. & Wang, R. X. The complete mitochondrial genome of bighead croaker, Collichthys niveatus (Perciformes, Sciaenidae): Structure of control region and phylogenetic considerations. Mol. Biol. Rep. 38, 4673–4685. https://doi.org/10.1007/s11033-010-0602-4 (2011).
CAS Article PubMed PubMed Central Google Scholar54.
Ojala, D., Montoya, J. & Attardi, G. tRNA punctuation model of RNA processing in human mitochondria. Nature 290, 470–474. https://doi.org/10.1038/290470a0 (1981).
CAS Article PubMed ADS PubMed Central Google Scholar55.
Vandana, R. P. et al. Characterization of the complete mitochondrial genome of Barilius malabaricus and its phylogenetic implications. Genomics 112, 2154–2163 (2019).
Google Scholar56.
Wang, X., Wang, J., He, S. & Mayden, R. L. The complete mitochondrial genome of the Chinese hook snout carp Opsariichthys bidens (Actinopterygii: Cypriniformes) and an alternative pattern of mitogenomic evolution in vertebrate. Gene 399, 0–19 (2007).
CAS Article Google Scholar57.
Gong, L., Liu, B., Lü, Z. M. & Liu, L. Q. Characterization of the complete mitochondrial genome of Wuhaniligobius polylepis (Gobiiformes: Gobiidae) and phylogenetic studies of Gobiiformes. Mitochondrial DNA Part B 3, 1117–1119. https://doi.org/10.1080/23802359.2018.1519380 (2018).
Article Google Scholar58.
Dowton, M. & Campbell, N. J. H. Intramitochondrial recombination—is it why some mitochondrial genes sleep around?. Trends Ecol. Evol. 16, 269–271 (2001).
CAS PubMed Article PubMed Central Google Scholar59.
Kong, X. D. et al. A novel rearrangement in the mitochondrial genome of tongue sole, Cynoglossus semilaevis: Control region translocation and a tRNA gene inversion. Genome 52, 975–984. https://doi.org/10.1139/g09-069 (2009).
CAS Article PubMed PubMed Central Google Scholar60.
Shi, W., Gong, L., Wang, S. Y., Miao, X. G. & Kong, X. Y. Tandem duplication and random loss for mitogenome rearrangement in symphurus (Teleost: Pleuronectiformes). BMC Genom. 16, 355 (2015).
Article CAS Google Scholar61.
Gong, L. et al. Large-scale mitochondrial gene rearrangements in the hermit crab Pagurus nigrofascia and phylogenetic analysis of the Anomura. Gene 695, 75–83 (2019).
CAS PubMed Article PubMed Central Google Scholar62.
Wang, Z. W. et al. Complete mitochondrial genome of Parasesarma affine (Brachyura: Sesarmidae): Gene rearrangements in Sesarmidae and phylogenetic analysis of the Brachyura. Int. J. Biol. Macromol. 118, 31–40 (2018).
CAS PubMed Article PubMed Central Google Scholar63.
Inoue, J. G., Miya, M., Tsukamoto, K. & Nishida, M. Evolution of the deep-sea gulper eel mitochondrial genomes: Large-scale gene rearrangements originated within the eels. Mol. Biol. Evol. 20, 1917–1924. https://doi.org/10.1093/molbev/msg206 (2003).
CAS Article PubMed PubMed Central Google Scholar64.
Inoue, J. G. et al. Deep-ocean origin of the freshwater eels. Biol. Lett. 6, 363–366. https://doi.org/10.1098/rsbl.2009.0989 (2010).
Article PubMed PubMed Central Google Scholar65.
Chen, J. N., López, J. A., Lavoué, S., Miya, M. & Chen, W. J. Phylogeny of the Elopomorpha (Teleostei): Evidence from six nuclear and mitochondrial markers. Mol. Phylogenet. Evol. 70, 152–161 (2014).
PubMed Article PubMed Central Google Scholar66.
Reece, J. S., Bowen, B. W., Smith, D. G. & Larson, A. Molecular phylogenetics of moray eels (Muraenidae) demonstrates multiple origins of a shell-crushing jaw (Gymnomuraena, Echidna) and multiple colonizations of the Atlantic Ocean. Mol. Phylogenet. Evol. 57, 829–835 (2010).
CAS PubMed Article PubMed Central Google Scholar67.
Johnson, G. D., Ida, H., Sakaue, J., Sado, T. & Asahida, T. A “living fossil” eel (Anguilliformes: Protanguillidae, fam. nov.) from an undersea cave in Palau. Proc. R. Soc. B Biol. Sci. 279, 934–943. https://doi.org/10.1098/rspb.2011.1289 (2012).
Article Google Scholar68.
Kumazawa, Y. & Nishida, M. Variations in mitochondrial tRNA gene organization of reptiles as phylogenetic markers. Mol. Biol. Evol. 12, 759–772. https://doi.org/10.1093/oxfordjournals.molbev.a040254 (1995).
CAS Article PubMed PubMed Central Google Scholar69.
Loh, K. H. et al. Next-generation sequencing yields the complete mitochondrial genome of the longfang moray, Enchelynassa canina (Anguilliformes: Muraenidae). Mitochondrial DNA Part A 27, 2431–2432 (2015).
Article CAS Google Scholar70.
Loh, K. H. et al. Next generation sequencing yields the complete mitochondrial genome of the Zebra moray, Gymnomuraena zebra (Anguilliformes: Muraenidae). Mitochondrial DNA Part A 27, 1–2 (2015).
Google Scholar71.
Perna, N. T. & Kocher, T. D. Patterns of nucleotide composition at fourfold degenerate sites of animal mitochondrial genomes. J. Mol. Evol. 41, 353–358 (1995).
CAS PubMed Article ADS PubMed Central Google Scholar72.
Sudhir, K. et al. MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol. Biol. Evol. 35, 1547–1549 (2018).
Article CAS Google Scholar73.
Paul, S. & Wishart, D. S. Circular genome visualization and exploration using CGView. Bioinformatics 21, 537–539 (2004).
Google Scholar74.
Nelson, J. S. Fishes of the Word 4th edn. (Wiley, New York, 2006).
Google Scholar75.
Xia, X. DAMBE7: New and improved tools for data analysis in molecular biology and evolution. Mol. Biol. Evol. 35, 1550–1552. https://doi.org/10.1093/molbev/msy073 (2018).
CAS Article PubMed PubMed Central Google Scholar76.
Shi, W., Kong, X., Wang, Z. M. & Jiang, J. X. Utility of tRNA Genes from the Complete Mitochondrial Genome of Psetta maxima for Implying a Possible Sister-group Relationship to the Pleuronectiformes. Zool. Stud. 50, 665–681 (2011).
CAS Google Scholar77.
Larkin, M. A. et al. Clustal W and Clustal X version 2.0. Bioinformatics 23, 2947–2948. https://doi.org/10.1093/bioinformatics/btm404 (2007).
CAS Article Google Scholar78.
Hall, T. BioEdit: A user-friendly biological sequence alignment program for Windows 95/98/NT. Nucleic Acids Sympos. Ser. 41, 95–98 (1999).
CAS Google Scholar79.
Gerard, T. & Jose, C. Improvement of phylogenies after removing divergent and ambiguously aligned blocks from protein sequence alignments. Syst. Biol. 56, 564–577 (2007).
Article CAS Google Scholar80.
Gascuel, O. New algorithms and methods to estimate maximum-likelihood phylogenies: Assessing the performance of PhyML 3.0. Syst. Biol. 59, 307–321 (2010).
PubMed Article CAS PubMed Central Google Scholar81.
Huelsenbeck, J. P. MrBayes 3.2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol. 61, 539–542 (2012).
PubMed PubMed Central Article Google Scholar82.
Posada, D. & Crandall, K. A. MODELTEST: Testing the model of DNA substitution. Bioinformatics 14, 817–818 (1998).
CAS PubMed PubMed Central Article Google Scholar83.
Nylander, J. A. A., Fredrik, R., Huelsenbeck, J. P. & Nieves-Aldrey, J. Bayesian phylogenetic analysis of combined data. Syst. Biol. 53, 47–67 (2004).
PubMed Article PubMed Central Google Scholar84.
Sitnikova, T., Rzhetsky, A. & Nei, M. Interior-branch and bootstrap tests of phylogenetic trees. Mol. Biol. Evol. 12, 319–333 (1995).
CAS PubMed PubMed Central Google Scholar85.
Antezana, M. When being “most likely” is not enough: Examining the performance of three uses of the parametric bootstrap in phylogenetics. J. Mol. Evol. 56, 198–222 (2003).
CAS PubMed Article ADS PubMed Central Google Scholar More88 Shares179 Views
in EcologyForaging niche partitioning in sympatric seabird populations
1.
Gause, G. F. Experimental populations of microscopic organisms. Ecology 18, 173–179 (1937).
Google Scholar
2.
DeBach, P. & Sundby, R. A. Competitive displacement between ecological homologues. Hilgardia 34, 105–166 (1963).
Article Google Scholar3.
Johnson, C. A. & Bronstein, J. L. Coexistence and competitive exclusion in mutualism. Ecology 100, e02708 (2019).
PubMed Article PubMed Central Google Scholar4.
Gause, G. F. Experimental analysis of Vito Volterra’s mathematical theory of the struggle for existence. Science 79, 16–17 (1934).
ADS CAS Article Google Scholar5.
Hutchinson, G. E. Concluding remarks Cold Spring Harbor Symp. Quant. Biol. 22, 415–427 (1958).
Article Google Scholar6.
MacArthur, R. H. Population ecology of some warblers of northeastern coniferous forests. Ecology 39, 599–619 (1958).
Article Google Scholar7.
Kooyers, N. J., James, B. & Blackman, B. K. Competition drives trait evolution and character displacement between Mimulus species along an environmental gradient. Evolution 71, 1205–1221 (2017).
CAS PubMed Article PubMed Central Google Scholar8.
Whittaker, R. H., Levin, S. A. & Root, R. B. Niche, Habitat, and Ecotope. Am. Nat. 107, 321–338 (1973).
Article Google Scholar9.
Wilson, R. P. Resource partitioning and niche hyper-volume overlap in free-living Pygoscelid penguins. Funct. Ecol. 24, 646–657 (2010).
Article Google Scholar10.
Berendse, F. Interspecific competition and niche differentiation between Plantago Ianceolata and Anthoxanthum odoratum in a natural hayfield. J. Ecol. 71, 379–390 (1983).
Article Google Scholar11.
Peterson, A. T. & Holt, R. D. Niche differentiation in Mexican birds: using point occurrences to detect ecological innovation. Ecol. Lett. 6, 774–782 (2003).
Article Google Scholar12.
Preez, B., Purdon, J., Trethowan, P., Macdonald, D. W. & Loveridge, A. J. Dietary niche differentiation facilitates coexistence of two large carnivores. J. Zool. 302, 149–156 (2017).
Article Google Scholar13.
Pratte, I., Robertson, G. J. & Mallory, M. L. Four sympatrically nesting auks show clear resource segregation in their foraging environment. Mar. Ecol. Prog. Ser. 572, 243–254 (2017).
ADS Article Google Scholar14.
Cody, M. L. Coexistence, coevolution and convergent evolution in seabird communities. Ecology 54, 31–44 (1973).
Article Google Scholar15.
Hjernquist, M. B., Hjernquist, M., Hjernquist, B. & Thuman Hjernquist, K. A. Common Guillemot Uria aalge differentiate their niche to coexist with colonizing Great Cormorants Phalacrocorax carbo. Atlantic Seabird 7, 83–89 (2005).
Google Scholar16.
Gaston, A. J. Seabirds: A Natural History (Yale University Press, New Haven, 2004).
Google Scholar17.
Frere, E., Quintana, F., Gandini, P. & Wilson, R. P. Foraging behaviour and habitat partitioning of two sympatric cormorants in Patagonia, Argentina. Ibis 150, 558–564 (2008).
Article Google Scholar18.
Linnebjerg, J. F., Reuleaux, A., Mouritsen, K. N. & Frederiksen, M. Foraging ecology of three sympatric breeding alcids in a declining colony in southwest Greenland. Waterbirds 38, 143–152 (2015).
Article Google Scholar19.
Symons, S. C. Ecological Segregation Between Two Closely Related Species: Exploring Atlantic Puffin and Razorbill Foraging Hotspots (M. Sc. thesis, University of New Brunswick) (2018).20.
Jenkins, E. J. & Davoren, G. K. Seabird species-and assemblage-level isotopic niche shifts associated with changing prey availability during breeding in coastal Newfoundland. Ibis https://doi.org/10.1111/ibi.12873 (2020).
Article Google Scholar21.
Gaston, A. J., Ydenberg, R. C. & Smith, G. E. J. Ashmole’s halo and population regulation in seabirds. Mar. Ornithol. 35, 119–126 (2007).
Google Scholar22.
Sánchez, S. et al. Within-colony spatial segregation leads to foraging behaviour variation in a seabird. Mar. Ecol. Prog. Ser. 606, 215–230 (2018).
ADS Article Google Scholar23.
Elliott, K. H. et al. Central-place foraging in an Arctic seabird provides evidence for Storer-Ashmole’s halo. Auk 126, 613–625 (2009).
Article Google Scholar24.
Wakefield, E. D. et al. Space partitioning without territoriality in gannets. Science 341, 68–70 (2013).
ADS CAS PubMed Article PubMed Central Google Scholar25.
Linnebjerg, J. F. et al. Sympatric breeding auks shift between dietary and spatial resource partitioning across the annual cycle. PLoS ONE 8, e72987 (2013).
ADS CAS PubMed PubMed Central Article Google Scholar26.
Masello, J. F. et al. Diving seabirds share foraging space and time within and among species. Ecosphere 1, 1–28 (2010).
Article Google Scholar27.
Hinke, J. T. et al. Spatial and isotopic niche partitioning during winter in chindstrap and Adélie penguins from the South Shetland Islands. Ecosphere 6, 1–32 (2015).
Article Google Scholar28.
Gulka, J., Ronconi, R. A. & Davoren, G. K. Spatial segregation contrasting dietary overlap: niche partitioning of two sympatric alcids during shifting resource availability. Mar. Biol. 166, 155 (2019).
Article Google Scholar29.
Ricklefs, R. E. & White, S. C. Growth and energetics of chicks of the Sooty Tern (Sterna fuscata) and Common Tern (S. hirundo). Auk 98, 361–378 (1981).
Google Scholar30.
Lance, M. M. & Thompson, C. W. Overlap in diets and foraging of Common Murres and Rhinoceros Auklets after the breeding season. Auk 122, 887–901 (2005).
Article Google Scholar31.
Maynard, L. D. & Davoren, G. K. Inter-colony and interspecific differences in the isotopic niche of two sympatric gull species in Newfoundland. Mar. Ornithol. 48, 103–109 (2020).
Google Scholar32.
Thaxter, C. B. et al. Influence of wing loading on the trade-off between pursuit-diving and flight in common guillemots and razorbills. J. Exp. Biol. 213, 1018–1025 (2010).
CAS PubMed Article PubMed Central Google Scholar33.
Thaxter, C. B. et al. Modelling the effects of prey size and distribution on prey capture rates of two sympatric marine predators. PLoS ONE 8, e79915 (2013).
ADS PubMed PubMed Central Article CAS Google Scholar34.
Shoji, A., Aris-Brosou, S. & Elliott, K. H. Physiological constraints and dive behavior scale in tandem with body mass in auks: A comparative analysis. Comp. Biochem. Physiol. A 196, 54–60 (2016).
CAS Article Google Scholar35.
Guigueno, M. F., Shoji, A., Elliott, K. H. & Aris-Brosou, S. Flight costs in volant vertebrates: A phylogenetically-controlled meta-analysis of birds and bats. Comp. Biochem. Physiol. A 235, 193–201 (2019).
CAS Article Google Scholar36.
Robertson, G. S., Bolton, M. & Monaghan, P. Influence of diet and foraging strategy on reproductive success in two morphologically similar sympatric seabirds. Bird Study 63, 316–329 (2016).
Article Google Scholar37.
Gaston, A. J. & Jones, I. L. The Auks: Alcidae (Oxford University Press, USA, 1998).
Google Scholar38.
Orians, G. H. & Pearson, N. E. On the theory of central place foraging. Analysis of ecological systems. Ohio State Univ. Press Columbus 2, 155–177 (1979).
Google Scholar39.
Rail, J. F. & Cotter, R. C. Seventeenth census of seabird populations in the sanctuaries of the North Shore of the Gulf of St. Lawrence, 2010. Can. Field-Nat. 121, 287–294 (2015).
Article Google Scholar40.
Lavoie, R. A., Rail, J. F. & Lean, D. R. Diet composition of seabirds from Corossol Island, Canada, using direct dietary and stable isotope analyses. Waterbirds 35, 402–419 (2012).
Article Google Scholar41.
Elliott, K. H. et al. High flight costs, but low dive costs, in auks support the biomechanical hypothesis for flightlessness in penguins. Proc. Natl. Acad. Sci. USA 110, 9380–9384 (2013).
ADS CAS PubMed Article PubMed Central Google Scholar42.
Croll, D. A., Gaston, A. J., Burger, A. E. & Konnoff, D. Foraging behavior and physiological adaptation for diving in Thick-billed Murres. Ecology 73, 344–356 (1992).
Article Google Scholar43.
Thaxter, C. B. et al. Sex-specific food provisioning in a monomorphic seabird, the common guillemot Uria aalge: nest defence, foraging efficiency or parental effort?. J. Avian Biol. 40, 75–84 (2009).
Article Google Scholar44.
Gremillet, D. & Boulinier, T. Spatial ecology and conservation of seabirds facing global climate change a review. Mar. Ecol. Prog. Ser. 391, 121–137 (2009).
ADS Article Google Scholar45.
Domalik, A. D., Hipfner, J. M., Studholme, K. R., Crossin, G. T. & Green, D. J. At-sea distribution and fine-scale habitat use patterns of zooplanktivorous Cassin’s auklets during the chick-rearing period. Mar. Biol. 165, 177 (2018).
Article Google Scholar46.
Lack, D. Ecological Isolation in Birds (Blackwell, Oxford, 1971).
Google Scholar47.
Wanless, S., Harris, M. P. & Morris, J. A. A comparison of feeding areas used by individual Common Murres (Uria aalge), Razorbills (Alca torda) and an Atlantic Puffin (Fratercula arctica) during the breeding season. Colonial Waterbirds 13, 16–24 (1990).
Article Google Scholar48.
Paredes, R., Jones, I. L., Boness, D. J., Tremblay, Y. & Renner, M. Sex-specific differences in diving behaviour of two sympatric Alcini species: thick-billed murres and razorbills. Can. J. Zool. 86, 610–622 (2008).
Article Google Scholar49.
Kotzerka, J., Garthe, S. & Hatch, S. A. GPS tracking devices reveal foraging strategies of Black-legged Kittiwakes. J. Ornithol. 151, 459–467 (2009).
Article Google Scholar50.
John, M. A. S., MacDonald, J. S., Harrison, P. J., Beamish, R. J. & Choromanski, E. The Fraser River plume: some preliminary observations on the distribution of juvenile salmon, herring, and their prey. Fish. Oceanogr. 1, 153–162 (1992).
Article Google Scholar51.
Phillips, E. M., Horne, J. K., Adams, J. & Zamon, J. E. Selective occupancy of a persistent yet variable coastal river plume by two seabird species. Mar. Ecol. Prog. Ser. 594, 245–261 (2018).
ADS Article Google Scholar52.
Shoji, A. et al. Foraging behaviour of sympatric razorbills and puffins. Mar. Ecol. Prog. Ser. 520, 257–267 (2015).
ADS Article Google Scholar53.
Chimienti, M. et al. Taking movement data to new depths: inferring prey availability and patch profitability from seabird foraging behavior. Ecol. Evol. 7, 10252–10256 (2017).
PubMed PubMed Central Article Google Scholar54.
Gaglio, D., Cook, T. R., Connan, M., Ryan, P. G. & Sherley, R. B. Dietary studies in birds: testing a non-invasive method using digital photography in seabirds. Methods Ecol. Evol. 8, 214–222 (2017).
Article Google Scholar55.
R Core Team. R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna, Austria. http://www.R-project.org/. (2019)56.
Calenge, C. The package adehabitat for the R software: a tool for the analysis of space and habitat use by animals. Ecol. Model. 197, 516–519 (2006).
Article Google Scholar57.
Hijmans, R. J. Geosphere: Spherical Trigonometry. R Package Version 1.5–10 (2019).58.
Burger, A. E. Arrival and departure behavior of common murres at colonies: evidence for an information halo?. Colonial Waterbirds 20, 55–65 (1997).
Article Google Scholar59.
Bradstreet, M. S. W. & Brown, R. G. B. Feeding studies. In Population Estimation, Productivity, and Food Habits of Nesting Seabirds at Cape Pierce and the Pribilof Islands, Bering Sea (ed. Johnson, S. R.) 257–306 (Anchorage, Alaska, 1985).
Google Scholar60.
Vaughn, H. R. H. Flight speed of guillemots, razorbills and puffins. Br. Birds 31, 123 (1937).
Google Scholar61.
Fifield, D. A., Lewis, K. P., Gjerdrum, C., Robertsoan, G. J. & Wells, R. Offshore seabird monitoring program. Environ. Stud. Res. Funds Rep. 183, 68 (2009).
Google Scholar62.
Pennycuick, C. J. Actual and “optimum” flight speeds: field data reassessed. J. Exp. Biol. 200, 2355–2361 (1997).
CAS PubMed PubMed Central Google Scholar63.
Elliott, K. H. et al. Windscapes shape seabird instantaneous energy costs but adult behavior buffers impact on offspring. Mov. Ecol. 2, 17 (2014).
PubMed PubMed Central Article Google Scholar64.
Gaston, A. J. et al. Modeling foraging range for breeding colonies of thick-billed murres Uria lomvia in the Eastern Canadian Arctic and potential overlap with industrial development. Biol. Conserv. 168, 134–143 (2013).
Article Google Scholar65.
Kruskal, W. H. & Wallis, W. A. Use of ranks in one-criterion variance analysis. J. Am. Stat. Assoc. 47, 583–621 (1952).
MATH Article Google Scholar66.
Neuhäuser, M. Wilcoxon–Mann–Whitney Test. In International Encyclopedia of Statistical Science (ed. Lovric, M.) (Springer, Berlin, Heidelberg, 2011).
Google Scholar67.
Haynes, W. Bonferroni Correction in Encyclopedia of Systems Biology (Dubitzky, W., Wolkenhauer, O., Cho, K.H., Yokota, H.) (Springer, New York, 2013).68.
Bates, D., Maechler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1–48 (2015).
Article Google Scholar69.
Wickham, H. ggplot2: Elegant Graphics for Data Analysis (Springer-Verlag, New York, 2016).
Google Scholar70.
Dunnington, D. ggspatial: Spatial Data Framework for ggplot2. R package version 1.1.4. https://CRAN.R-project.org/package=ggspatial (2020).71.
South, A. rnaturalearth: World Map Data from Natural Earth. R package version 0.1.0. https://CRAN.R-project.org/package=rnaturalearth (2017).72.
South, A. rnaturalearthdata: World Vector Map Data from Natural Earth used in ‘rnaturalearth’. https://CRAN.R-project.org/package=rnaturalearthdata (2017).73.
Pabesma, E. Simple features for R: standardized support for spatial vector data. R J 10, 439–446 (2018).
Article Google Scholar74.
Fieberg, J. & Kochanny, C. O. Quantifying home-range overlap: The importance of the utilization distribution. J. Wildl. Manag. 69, 1346–1359 (2005).
Article Google Scholar75.
Delord, K. et al. Movements of three alcid species breeding sympatrically in Saint Pierre and Miquelon, northwestern Atlantic Ocean. J. Ornithol. 161, 359–371 (2019).
Article Google Scholar76.
Wei, T. & Simko, V. R package “corrplot”: Visualization of a Correlation Matrix (Version 0.84) (2017). https://github.com/taiyun/corrplot.77.
Lê, S., Josse, J. & Husson, F. FactoMineR: an R package for multivariate analysis. J. Stat. Softw. 25, 1–18 (2008).
Article Google Scholar78.
Kassambara, A. & Mundt, F. factoextra: Extract and Visualize the Results of Multivariate Data Analyses. https://CRAN.R-project.org/package=factoextra, (Version 1.0.3) (2016).79.
Rasband, W. S. ImageJ, U. S. National Institutes of Health, Bethesda, Maryland, USA, https://imagej.nih.gov/ij/ (1997–2018).80.
Lambert, J. D. & Bernier, B. Observations on 4RST capelin in the Gulf of St. Lawrence (A retrospective, 1984–1987). CAFSAC Res. Document 89 (1989).81.
Elliott, K. H. & Gaston, A. J. Mass-length relationships and energy content of fishes and invertebrates delivered to nesting Thick-billed Murres Uria lomvia in the Canadian Arctic, 1981–2007. Mar. Ornithol. 36, 25–34 (2008).
Google Scholar82.
Noble, V. R. & Clark, D. S. Seasonal length: weight relationships of Grenadiers, Chimaeras, and Atlantic Herring caught by Fisheries and Oceans Canada’s Maritimes Region Ecosystem Surveys, using different measurement techniques at sea. Can. Data Rep. Fish. Aquat. Sci. 1291, 4–14 (2019).
Google Scholar83.
Silva, J. F., Ellis, J. R. & Ayers, R. A. Length-weight relationships of marine fish collected from around the British Isles. Science 150, 109 (2013).
Google Scholar84.
Morin, R., Ricard, D., Benoît, H. & Surette, T. A review of the biology of Atlantic hagfish (Myxine glutinosa), its ecology, and its exploratory fishery in the southern Gulf of St. Lawrence (NAFO Div. 4T). DFO Canadian Science Advisory Secretariat, v + 39 (2017).85.
Erguden, D., Turan, F. & Turan, C. Length–weight and length–length relationships for four shad species along the western Black Sea coast of Turkey. J. Appl. Ichthyol. 27, 942–944 (2011).
Article Google Scholar86.
Sievert, C. plotly for R. https://plotly-book.cpsievert.me. [p506] (2018).87.
Wickham, H., François, R., Henry, L. & Müller, K. dplyr: A Grammar of Data Manipulation. R package version 0.8.3. https://CRAN.R-project.org/package=dplyr (2019).88.
Zeileisk, A., Kleiber, C. & Jackman, S. Regression models for count data in R. J. Stat. Softw. 27, 1–25 (2008).
Google Scholar89.
Chambers, J. M. Linear models. In Statistical Models (eds Chambers, J. M. & Hastie, T. J.) (Wadsworth and Brooks/Cole, Belmont, 1992).
Google Scholar90.
Ahlmann-Eltze, C. ggsignif: Significance Brackets for ggplot2. https://CRAN.R-project.org/package=ggsignif (2019). More213 Shares159 Views
in EcologyModerately decreasing fertilizer in fields does not reduce populations of cereal aphids but maximizes fitness of parasitoids
Through a three-year investigation, we found that a moderate decrease of nitrogen from 280 to 140–210 kg N ha−1 did not markedly influence the populations of cereal aphids or the parasitism rate. However, a moderate decrease of nitrogen input from 280 to 140–210 kg N ha−1 maximized the fitness of two predominant Aphidiinae parasitoid species, suggesting parasitoid control of cereal aphid would get benefit from the moderate decrease of nitrogen fertilizer. Those results showed that moderately decreasing nitrogen fertilizer could boost the parasitoid control of cereal aphids. Our research suggests that moderately decreasing nitrogen input is qualitatively beneficial to parasitoids but would not control cereal aphids quantitatively.
Effect of decreasing nitrogen fertilizer on the cereal aphid population
This study demonstrated that nitrogen fertilizer has the potential to positively influence densities of S. avenae and R. padi among all manipulated nitrogen fertilizer levels (70–280 kg N ha−1) (Fig. 1). Similar conclusions have been documented in research linked with aphids, including cereal aphids5,17,24. First, the plant usually responds monotonously and positively to nitrogen fertilizer. The percentage of nitrogen in the dry weight of tobacco leaves was positively associated with fertilizer levels25. Nitrogen fertilizer in the range of 0–225 kg N ha−1 improved nitrogen concentration of canola throughout the growing season26. It has been reported that fertilization has a positive influence on plants, indicating a cascading effect on herbivorous pests24,26,27. Nitrogen input could enhance the nutritional quality of the host, as nitrogen input increases sugars and amino acids availability for aphids, thereby accelerating the population growth of the herbivores28,29. Second, fertilization negatively affects plant defensive responses to herbivores and lessens the amounts of toxins in host plants27. For example, nitrogen fertilizer employed for walnut seedlings decreased the allocation to defensive toxins such as juglone, thereby lowering resistance to walnut aphids30. Third, fertilization alters the microclimate of crops and thereby contributes to the population growth of aphids17,31.
However, only the lowest nitrogen level manipulated in our experiment (70 kg N ha−1) significantly reduced the population of cereal aphids compared with the conventional nitrogen level (280 kg N ha−1) in 2016 and 2017 (Fig. 1). Those results showed that the magnitude of decreasing fertilizer input from the conventional level (280 kg N ha−1) to a moderate level (140–210 kg N ha−1) was insufficient to contain the population of cereal aphids. The performance of cereal aphids could remain unaffected when fertilizer input was decreased to a low level, as aphids could adapt to the pressure of deficient nutrition by sucking more strongly10. Therefore, to reduce the population of cereal aphids, the nitrogen level should be decreased to 70 kg·N·ha−1 or lower. Similarly, as fertilizer was applied to tobacco in the range of 0–200 ppm N, the nymph weights of whiteflies on tobacco plants did not diminish markedly until the nitrogen concentration level was reduced from 200 to 0 ppm N25.
Nevertheless, cereal yield responds to nitrogen levels as a negatively accelerating curve based on previous studies7,9. Far lower nitrogen input sharply reduces grain yield, and moderate nitrogen fertilizer is always imperative in agricultural production2,7. Therefore, the tradeoff between maintaining the essential grain yield and reduction of the pest population would not have been optimized solely by decreasing nitrogen input.
The wheat variety adopted in our experiment was susceptible to cereal aphids. The landscape around our field employed in this experiment was predominated by winter wheat, and thus the landscape was extremely simplified. By comparison, use of a resistant variety and intercropping wheat with another crop mediated the impact of nitrogen input on densities of cereal aphids10,12. If these factors are taken into consideration, it then seems more unlikely that the pest population can be controlled solely by decreasing nitrogen input in complex realistic agricultural environments.
Effect of decreasing nitrogen fertilizer on the densities of parasitoids and parasitism rate
The results showed that the parasitism rate remained unchanged with nitrogen input (Fig. 2), similar to the results of Garratt, who pointed out that fertilizer levels did not affect the parasitism rate in a cereal-aphid-parasitoid system, as the densities of aphids and their parasitoids increased synchronously with the amount of fertilizer18. Similar findings were observed in a walnut aphid-Aphidiinae parasitoid system24. Mixed results were reported in previous studies5,11. The densities of cereal aphids and parasitoids increased when input of nitrogen fertilizer increased from 115 to 170 kg N ha−1, while the parasitism rate increased steadily5.
Parasitoids are subject to pressures derived from higher trophic level. Coincidental intraguild predation is ubiquitous in the form of parasitized aphids suffering from predation. The effect of coincidental intraguild predation on biocontrol and the abundance of parasitoids remains controversial32,33. Importantly, the Aphidiinae parasitoids have the potential to identify the odors of ladybird beetles and reduce searching efficiency by themselves and their offspring, a trait-mediated indirect effect unrelated with the densities of ladybird beetles34. It is possible that the behavior of Aphidiinae parasitoids and the parasitism rate could have been mediated indirectly by ladybird beetles and other predators. Furthermore, the hyperparasitoids also could have relieved biocontrol by Aphidiinae parasitoids35. Hence, the higher trophic level could relieve the effects of nitrogen levels on densities of parasitoids and the parasitism rate.
Effect of decreasing nitrogen fertilizer on the body size of Aphidiinae parasitoids
This research has shown that nitrogen fertilizer application impacted the body sizes of the two Aphidiinae parasitoids (Figs. 3, 4). It has been reported that the body sizes of parasitoids increased monotonically with nitrogen fertilizer under low densities of aphids in the laboratory18,22, meanwhile the dispersion capacity of parasitoid adults, the fecundity of adult females, the emergence rate, the adult longevity of parasitoids, and the parasitism rate increased with the body sizes of parasitoids19,20,22. In contrast to previous reports, this field study found that a moderate decrease in nitrogen application from 280 to 140–210 kg N ha−1 maximized the body sizes of parasitoids. The body sizes of parasitoids depend negatively on the abundance of parasitoids and positively on the hosts diversity19,36,37. Hence, combining the positive effect of the abundance of aphids and of the nitrogen input with the negative effect of parasitoid abundance, it is assumed that an equilibrium should emerge balancing the positive effect of abundance of aphids and the negative effect of abundance of parasitoids. Analogously, It has been reported that an optimized nitrogen level maximized the ratio of predators to prey in a canola-mustard aphid-predatory gall midge system26.
Manipulating nitrogen fertilizer to maximize the fitness of parasitoids plays a crucial role in natural pest control. Increasing the body sizes of parasitoids means greater fertility and dispersal ability of adults20,21, higher fitness of offspring38, and the resulting greater capacity to control the aphid. Thus, decreasing nitrogen fertilizer from the conventional level to more environmentally-friendly magnitudes (140–210 kg N ha−1) could increase the fitness of Aphidiinae parasitoids and boost the biocontrol by parasitoids. Regrettably, this research study did not validate such a viewpoint since the parasitism rate was not maximized under the moderate nitrogen levels. First, there may be hysteresis effects. The parasitoids that were measured for body sizes came from mummies that were sampled in the flowering phases. These parasitoids came into play and mummified cereal aphids after more than ten days. The mummies remained scarce before the flowering phase. Thus, a notable lag occurred and the effect of parasitoid fitness on the parasitism rate could have been unobservable in this study. Second, apart from affecting parasitoid fitness, nitrogen application affected pest fitness. A moderate amount nitrogen maximized the performance of the green peach aphid and the Bertha armyworm23,39. A positive relationship between aphid weight and hind tibia length of parasitoids has been reported18. Combined with the finding in this study that the body sizes of parasitoids were maximized by moderate nitrogen levels, these results imply that the fitness of cereal aphids also benefited from moderate nitrogen levels. However, the densities of cereal aphids in moderate nitrogen levels were similar to those under higher nitrogen levels, suggesting that there could be a compensation between the effect of nitrogen input on fitness of cereal aphids and the effect of nitrogen input on fitness of parasitoids. Currently, long-term agricultural intensification limited biocontrol of parasitoids5. Previous study has reported that the parasitoids were more strongly influenced by agricultural intensification compared to cereal aphids5,13,14. If serious agricultural intensification had mediated, for example decreasing nitrogen fertilizer to an optimized extent, the equilibrium between the impact of moderate decreasing nitrogen fertilizer on parasitoids and the counterpart on cereal aphids would be reshaped. Thus, the positive influence of decreasing nitrogen fertilizer on parasitoids would prevail. Coincidentally, such a magnitude of decreasing nitrogen application would maintain the current wheat yield and lessen the potential environmental risks9.
Relationship between the parasitism rate and the population growth of cereal aphids
From flowering to milking phase, the population of the cereal aphid R. padi that escaped from Aphidiinae parasitoids increased substantially in both 2017 and 2018, while the population of the cereal aphid S. avenae decreased markedly in both 2016 and 2017 (Table 1). Combining the differences between dynamics of the two cereal aphid species with the fact that the Aphidiinae parasitoids rarely parasitize R. padi in China40, it is apparent that the Aphidiinae parasitoids play a pivotal role in suppressing the cereal aphid S. avenae. Furthermore, a higher parasitism rate had a greater suppression effect on the population of the cereal aphid S. avenae, in line with previous research6,14,41.
Year-to-year fluctuation of the cereal aphids-Aphidiinae parasitoids interaction
Obvious fluctuations in the cereal aphids-Aphidiinae parasitoids interaction across years have been documented in this study. Such population fluctuations of aphids and their natural enemies are ubiquitous14,17,42. It has been assumed that a disadvantageous climate accounted for the fluctuations17. The climate changes could not have been manipulated in our study, but they play essential roles in population fluctuations43. Climate warming induced an outbreak of the cereal aphids, but the parasitism rate remained unchanged43,44. Lack of Aphidiinae parasitoids caused higher populations of the cereal aphid S. avenae in a simulated warmed wheat field. However, abundant Aphidiinae parasitoids retained effective suppression of the cereal aphids even when the wheat field was warmed45. The synchronization of parasitoids with pests is vitally important for maintaining biocontrol46, while climate change has the potential to mismatch the pests with parasitoids and cause strong population fluctuations of pests and natural enemies47.
In this study, the parasitism rate was evaluated according to the densities of discernible mummies, a conventional method widely adopted5,6,24. We keep in mind that this method neglects the fact that the symptomless aphids that have been parasitized. Consequently, the parasitism rare was underestimated and the annual fluctuations of abundance of the parasitoids and the parasitism rate were magnified, especially early in the season. Molecular detection, which has the capacity to evaluate whether symptomless aphids have been parasitized and if so by which parasitoid species, presents an exceedingly promising alternative for exploring the aphid-parasitoid interaction11,33. This burgeoning method should be employed to more accurately evaluate the aphids-parasitoids interaction. More
