More stories

  • in

    The microbiome of a bacterivorous marine choanoflagellate contains a resource-demanding obligate bacterial associate

    Worden, A. Z. et al. Rethinking the marine carbon cycle: factoring in the multifarious lifestyles of microbes. Science 347, 1257594 (2015).PubMed 
    Article 
    CAS 

    Google Scholar 
    Bjorbækmo, M. F. M., Evenstad, A., Røsæg, L. L., Krabberød, A. K. & Logares, R. The planktonic protist interactome: where do we stand after a century of research? ISME J. 14, 544–559 (2020).PubMed 
    Article 

    Google Scholar 
    Pandolfi, J. M., Staples, T. L. & Kiessling, W. Increased extinction in the emergence of novel ecological communities. Science 370, 220–222 (2020).CAS 
    PubMed 
    Article 

    Google Scholar 
    Jürgens, K. & Massana, R. in Microbial Ecology of the Oceans (ed. Kirchman, D. L.) 383–441 (John Wiley & Sons, 2008).Archibald, J. M. Endosymbiosis and eukaryotic cell evolution. Curr. Biol. 25, R911–R921 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    McCutcheon, J. P. & Moran, N. A. Extreme genome reduction in symbiotic bacteria. Nat. Rev. Microbiol. 10, 13–26 (2011).PubMed 
    Article 
    CAS 

    Google Scholar 
    Drew, G. C., Stevens, E. J. & King, K. C. Microbial evolution and transitions along the parasite–mutualist continuum. Nat. Rev. Microbiol. 19, 623–638 (2021).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Buck, K. R., Chavez, F. P. & Thomsen, H. A. Choanoflagellates of the central California waters: abundance and distribution. Ophelia 33, 179–186 (1991).Article 

    Google Scholar 
    Leadbeater, B. S. C. The Choanoflagellates: Evolution, Biology and Ecology (Cambridge Univ. Press, 2015).de Vargas, C. et al. Eukaryotic plankton diversity in the sunlit ocean. Science 348, 1261605 (2015).PubMed 
    Article 
    CAS 

    Google Scholar 
    Alegado, R. A. et al. A bacterial sulfonolipid triggers multicellular development in the closest living relatives of animals. eLife 1, e00013 (2012).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Woznica, A. et al. Bacterial lipids activate, synergize, and inhibit a developmental switch in choanoflagellates. Proc. Natl Acad. Sci. USA 113, 7894–7899 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Woznica, A., Gerdt, J. P., Hulett, R. E., Clardy, J. & King, N. Mating in the closest living relatives of animals is induced by a bacterial chondroitinase. Cell 170, 1175–1183.e11 (2017).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Needham, D. M. et al. A distinct lineage of giant viruses brings a rhodopsin photosystem to unicellular marine predators. Proc. Natl Acad. Sci. USA 116, 20574–20583 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Needham, D. M. et al. Targeted metagenomic recovery of four divergent viruses reveals shared and distinctive characteristics of giant viruses of marine eukaryotes. Phil. Trans. R. Soc. Lond. B 374, 20190086 (2019).CAS 
    Article 

    Google Scholar 
    Frank, N., Helge Abuldhauge, T. & Daniel, J. R. Bridging the gap between morphological species and molecular barcodes – exemplified by loricate choanoflagellates. Eur. J. Protistol. 57, 26–37 (2017).Article 

    Google Scholar 
    Logares, R. et al. Disentangling the mechanisms shaping the surface ocean microbiota. Microbiome 8, 55 (2020).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Eldin, C. et al. From Q fever to Coxiella burnetii infection: a paradigm change. Clin. Microbiol. Rev. 30, 115–190 (2017).PubMed 
    Article 

    Google Scholar 
    Parks, D. H. et al. A standardized bacterial taxonomy based on genome phylogeny substantially revises the tree of life. Nat. Biotechnol. 36, 996–1004 (2018).CAS 
    PubMed 
    Article 

    Google Scholar 
    Lenski, R. E. in Advances in Microbial Ecology (ed. Marshall, K. C.) 1–44 (Springer, 1988).Zimmerman, A. E. et al. Metabolic and biogeochemical consequences of viral infection in aquatic ecosystems. Nat. Rev. Microbiol. 18, 21–34 (2020).CAS 
    PubMed 
    Article 

    Google Scholar 
    Vincent, F., Sheyn, U., Porat, Z., Schatz, D. & Vardi, A. Visualizing active viral infection reveals diverse cell fates in synchronized algal bloom demise. Proc. Natl Acad. Sci. USA 118, e2021586118 (2021).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Mruwat, N. et al. A single-cell polony method reveals low levels of infected Prochlorococcus in oligotrophic waters despite high cyanophage abundances. ISME J 15, 41–54 (2021).CAS 
    PubMed 
    Article 

    Google Scholar 
    Canbäck, B., Tamas, I. & Andersson, S. G. E. A phylogenomic study of endosymbiotic bacteria. Mol. Biol. Evol. 21, 1110–1122 (2004).PubMed 
    Article 
    CAS 

    Google Scholar 
    Giovannoni, S. J. SAR11 bacteria: the most abundant plankton in the oceans. Annu. Rev. Mar. Sci. 9, 231–255 (2017).Article 

    Google Scholar 
    Parks, D. H., Imelfort, M., Skennerton, C. T., Hugenholtz, P. & Tyson, G. W. CheckM: assessing the quality of microbial genomes recovered from isolates, single cells, and metagenomes. Genome Res. 25, 1043–1055 (2015).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Pachiadaki, M. G. et al. Charting the complexity of the marine microbiome through single-cell genomics. Cell 179, 1623–1635.e11 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Cordero, O. X. & Polz, M. F. Explaining microbial genomic diversity in light of evolutionary ecology. Nat. Rev. Microbiol. 12, 263–273 (2014).CAS 
    PubMed 
    Article 

    Google Scholar 
    Delmont, T. O. et al. Single-amino acid variants reveal evolutionary processes that shape the biogeography of a global SAR11 subclade. eLife 8, e46497 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Kashtan, N. et al. Single-cell genomics reveals hundreds of coexisting subpopulations in wild. Prochlorococcus. Science 344, 416–420 (2014).CAS 
    PubMed 

    Google Scholar 
    Toft, C. & Andersson, S. G. E. Evolutionary microbial genomics: insights into bacterial host adaptation. Nat. Rev. Genet. 11, 465–475 (2010).CAS 
    PubMed 
    Article 

    Google Scholar 
    Qiu, J. & Luo, Z.-Q. Legionella and Coxiella effectors: strength in diversity and activity. Nat. Rev. Microbiol. 15, 591–605 (2017).CAS 
    PubMed 
    Article 

    Google Scholar 
    Husnik, F. et al. Bacterial and archaeal symbioses with protists. Curr. Biol. 31, R862–R877 (2021).CAS 
    PubMed 
    Article 

    Google Scholar 
    Boamah, D. K., Zhou, G., Ensminger, A. W. & O’Connor, T. J. From many hosts, one accidental pathogen: the diverse protozoan hosts of Legionella. Front. Cell. Infect. Microbiol. 7, 477 (2017).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Graf, J. S. et al. Anaerobic endosymbiont generates energy for ciliate host by denitrification. Nature 591, 445–450 (2021).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Pinhassi, J., DeLong, E. F., Béjà, O., González, J. M. & Pedrós-Alió, C. Marine bacterial and archaeal ion-pumping rhodopsins: genetic diversity, physiology, and ecology. Microbiol. Mol. Biol. Rev. 80, 929–954 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Brunet, T. et al. Light-regulated collective contractility in a multicellular choanoflagellate. Science 366, 326–334 (2019).CAS 
    PubMed 
    Article 

    Google Scholar 
    Schmitz-Esser, S. et al. ATP/ADP translocases: a common feature of obligate intracellular amoebal symbionts related to Chlamydiae and Rickettsiae. J. Bacteriol. 186, 683–691 (2004).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    George, E. E. et al. Highly reduced genomes of protist endosymbionts show evolutionary convergence. Curr. Biol. 30, 925–933.e3 (2020).CAS 
    PubMed 
    Article 

    Google Scholar 
    Deeg, C. M. et al. Chromulinavorax destructans, a pathogen of microzooplankton that provides a window into the enigmatic candidate phylum Dependentiae. PLoS Pathog. 15, e1007801–e1007801 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Major, P., Embley, T. M. & Williams, T. A. Phylogenetic diversity of NTT nucleotide transport proteins in free-living and parasitic bacteria and eukaryotes. Genome Biol. Evol. 9, 480–487 (2017).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Trentmann, O., Decker, C., Winkler, H. H. & Neuhaus, H. E. Charged amino-acid residues in transmembrane domains of the plastidic ATP/ADP transporter from Arabidopsis are important for transport efficiency, substrate specificity, and counter exchange properties. Eur. J. Biochem. 267, 4098–4105 (2000).CAS 
    PubMed 
    Article 

    Google Scholar 
    Zhang, G., Meredith, T. C. & Kahne, D. On the essentiality of lipopolysaccharide to Gram-negative bacteria. Curr. Opin. Microbiol. 16, 779–785 (2013).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Bertani, B. & Ruiz, N. Function and biogenesis of lipopolysaccharides. EcoSal Plus 8, ESP-0001–2018 (2018).Article 

    Google Scholar 
    Russell, D. G., Vanderven, B. C., Glennie, S., Mwandumba, H. & Heyderman, R. S. The macrophage marches on its phagosome: dynamic assays of phagosome function. Nat. Rev. Immunol. 9, 594–600 (2009).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Sañudo-Wilhelmy, S. A., Gómez-Consarnau, L., Suffridge, C. & Webb, E. A. The role of B vitamins in marine biogeochemistry. Annu. Rev. Mar. Sci. 6, 339–367 (2014).Article 

    Google Scholar 
    Omsland, A. & Heinzen, R. A. Life on the outside: the rescue of Coxiella burnetii from its host cell. Annu. Rev. Microbiol. 65, 111–128 (2011).CAS 
    PubMed 
    Article 

    Google Scholar 
    Weeks, A. R., Turelli, M., Harcombe, W. R., Reynolds, K. T. & Hoffmann, A. A. From parasite to mutualist: rapid evolution of Wolbachia in natural populations of Drosophila. PLoS Biol. 5, e114 (2007).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Oliver, K. M., Campos, J., Moran, N. A. & Hunter, M. S. Population dynamics of defensive symbionts in aphids. Proc. Biol. Sci. 275, 293–299 (2008).PubMed 

    Google Scholar 
    Schulz, F. & Horn, M. Intranuclear bacteria: inside the cellular control center of eukaryotes. Trends Cell Biol. 25, 339–346 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    Hamann, E. et al. Environmental Breviatea harbour mutualistic Arcobacter epibionts. Nature 534, 254–258 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Seah, B. K. B. et al. Sulfur-oxidizing symbionts without canonical genes for autotrophic CO2 fixation. mBio 10, e01112-19 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Salonen, I. S., Chronopoulou, P.-M., Bird, C., Reichart, G.-J. & Koho, K. A. Enrichment of intracellular sulphur cycle-associated bacteria in intertidal benthic foraminifera revealed by 16S and aprA gene analysis. Sci. Rep. 9, 11692 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Vallesi, A. et al. A new species of the γ-Proteobacterium Francisella, F. adeliensis sp. nov., endocytobiont in an Antarctic marine ciliate and potential evolutionary forerunner of pathogenic species. Microb. Ecol. 77, 587–596 (2019).CAS 
    PubMed 
    Article 

    Google Scholar 
    Tashyreva, D. et al. Life cycle, ultrastructure, and phylogeny of new diplonemids and their endosymbiotic bacteria. mBio 9, e02447-17 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Foster, R. A. & Zehr, J. P. Diversity, genomics, and distribution of phytoplankton-cyanobacterium single-cell symbiotic associations. Annu. Rev. Microbiol. 73, 435–456 (2019).CAS 
    PubMed 
    Article 

    Google Scholar 
    Lin, Y.-C. et al. Distribution patterns and phylogeny of marine stramenopiles in the North Pacific Ocean. Appl. Environ. Microbiol. 78, 3387–3399 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Kim, E. et al. Newly identified and diverse plastid-bearing branch on the eukaryotic tree of life. Proc. Natl Acad. Sci. USA 108, 1496–1500 (2011).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Wylezich, C., Karpov, S. A., Mylnikov, A. P., Anderson, R. & Jürgens, K. Ecologically relevant choanoflagellates collected from hypoxic water masses of the Baltic Sea have untypical mitochondrial cristae. BMC Microbiol. 12, 271 (2012).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Wilson, A. C. C. & Duncan, R. P. Signatures of host/symbiont genome coevolution in insect nutritional endosymbioses. Proc. Natl Acad. Sci. USA 112, 10255–10261 (2015).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Douglas, A. E. Multiorganismal insects: diversity and function of resident microorganisms. Annu. Rev. Entomol. 60, 17–34 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    Newton, H. J. et al. Sel1 repeat protein LpnE is a Legionella pneumophila virulence determinant that influences vacuolar trafficking. Infect. Immun. 75, 5575–5585 (2007).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Boch, J., Bonas, U. & Lahaye, T. TAL effectors–pathogen strategies and plant resistance engineering. New Phytol. 204, 823–832 (2014).CAS 
    PubMed 
    Article 

    Google Scholar 
    Schmitz-Esser, S. et al. The genome of the amoeba symbiont ‘Candidatus Amoebophilus asiaticus’ reveals common mechanisms for host cell interaction among amoeba-associated bacteria. J. Bacteriol. 192, 1045–1057 (2010).CAS 
    PubMed 
    Article 

    Google Scholar 
    Abby, S. S. et al. Identification of protein secretion systems in bacterial genomes. Sci. Rep. 6, 23080 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Bratanis, E., Andersson, T., Lood, R. & Bukowska-Faniband, E. Biotechnological potential of Bdellovibrio and like organisms and their secreted enzymes. Front. Microbiol. https://doi.org/10.3389/fmicb.2020.00662 (2020).Rose, J., Caron, D., Sieracki, M. & Poulton, N. Counting heterotrophic nanoplanktonic protists in cultures and aquatic communities by flow cytometry. Aquat. Microb. Ecol. 34, 263–277 (2004).Article 

    Google Scholar 
    Caporaso, J. G. et al. Global patterns of 16S rRNA diversity at a depth of millions of sequences per sample. Proc. Natl Acad. Sci. USA 108, 4516–4522 (2010).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Stoeck, T. et al. Multiple marker parallel tag environmental DNA sequencing reveals a highly complex eukaryotic community in marine anoxic water. Mol. Ecol. 19, 21–31 (2010).CAS 
    PubMed 
    Article 

    Google Scholar 
    Edgar, R. C. Search and clustering orders of magnitude faster than BLAST. Bioinformatics 26, 2460–2461 (2010).CAS 
    PubMed 
    Article 

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

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

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

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

    Google Scholar 
    Altschul, S. F., Gish, W., Miller, W., Myers, E. W. & Lipman, D. J. Basic local alignment search tool. J. Mol. Biol. 215, 403–410 (1990).CAS 
    PubMed 
    Article 

    Google Scholar 
    Nelson, M. C., Morrison, H. G., Benjamino, J., Grim, S. L. & Graf, J. Analysis, optimization and verification of illumina-generated 16S rRNA gene amplicon surveys. PLoS ONE 9, e94249 (2014).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Bankevich, A. et al. SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J. Comput. Biol. 19, 455–477 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Eren, A. M. et al. Anvi’o: an advanced analysis and visualization platform for ‘omics data. PeerJ 3, e1319 (2015).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Jain, C., Rodriguez-R, L. M., Phillippy, A. M., Konstantinidis, K. T. & Aluru, S. High throughput ANI analysis of 90K prokaryotic genomes reveals clear species boundaries. Nat. Commun. 9, 5114 (2018).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Darling, A. E., Mau, B. & Perna, N. T. progressiveMauve: multiple genome alignment with gene gain, loss and rearrangement. PLoS ONE 5, e11147 (2010).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Li, H. Minimap2: pairwise alignment for nucleotide sequences. Bioinformatics 34, 3094–3100 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

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

    Google Scholar 
    Gao, F. & Zhang, C.-T. Ori-Finder: a web-based system for finding oriCs in unannotated bacterial genomes. BMC Bioinformatics 9, 79 (2008).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Delcher, A. L., Phillippy, A., Carlton, J. & Salzberg, S. L. Fast algorithms for large-scale genome alignment and comparison. Nucleic Acids Res. 30, 2478–2483 (2002).PubMed 
    PubMed Central 
    Article 

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

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

    Google Scholar 
    Laslett, D. & Canback, B. ARAGORN, a program to detect tRNA genes and tmRNA genes in nucleotide sequences. Nucleic Acids Res. 32, 11–16 (2004).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Hyatt, D. et al. Prodigal: prokaryotic gene recognition and translation initiation site identification. BMC Bioinformatics 11, 119 (2010).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Huerta-Cepas, J. et al. Fast genome-wide functional annotation through orthology assignment by eggNOG-Mapper. Mol. Biol. Evol. 34, 2115–2122 (2017).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Huerta-Cepas, J. et al. eggNOG 5.0: a hierarchical, functionally and phylogenetically annotated orthology resource based on 5090 organisms and 2502 viruses. Nucleic Acids Res. 47, D309–D314 (2019).CAS 
    PubMed 
    Article 

    Google Scholar 
    Eddy, S. R. Profile hidden Markov models. Bioinformatics 14, 755–763 (1998).CAS 
    PubMed 
    Article 

    Google Scholar 
    Finn, R. D. et al. Pfam: the protein families database. Nucleic Acids Res. 42, D222–D230 (2014).CAS 
    PubMed 
    Article 

    Google Scholar 
    Arkin, A. P. et al. KBase: The United States Department of Energy Systems Biology Knowledgebase. Nat. Biotechnol. 36, 566 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Aziz, R. K. et al. The RAST Server: rapid annotations using subsystems technology. BMC Genomics 9, 75 (2008).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Karp, P. D. et al. Pathway Tools version 19.0 update: software for pathway/genome informatics and systems biology. Brief. Bioinform. 17, 877–890 (2016).CAS 
    PubMed 
    Article 

    Google Scholar 
    Käll, L., Krogh, A. & Sonnhammer, E. L. L. Advantages of combined transmembrane topology and signal peptide prediction–the Phobius web server. Nucleic Acids Res. 35, W429–W432 (2007).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Elbourne, L. D. H., Tetu, S. G., Hassan, K. A. & Paulsen, I. T. TransportDB 2.0: a database for exploring membrane transporters in sequenced genomes from all domains of life. Nucleic Acids Res. 45, D320–D324 (2017).CAS 
    PubMed 
    Article 

    Google Scholar 
    Sandoz, K. M. et al. Transcriptional profiling of Coxiella burnetii reveals extensive cell wall remodeling in the small cell variant developmental form. PLoS ONE 11, e0149957 (2016).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Rekha, S. et al. Complete genome sequence of the Q-fever pathogen Coxiella burnetii. Proc. Natl Acad. Sci. USA 100, 5455–5460 (2003).Article 
    CAS 

    Google Scholar 
    Bushnell, B. BBMap Short Read Aligner (Univ. California, Berkeley, 2016); http://sourceforge.net/projects/bbmapAnders, S., Pyl, P. T. & Huber, W. HTSeq—a Python framework to work with high-throughput sequencing data. Bioinformatics 31, 166–169 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    Chen, I.-M. A. et al. IMG/M v.5.0: an integrated data management and comparative analysis system for microbial genomes and microbiomes. Nucleic Acids Res. 47, D666–D677 (2019).CAS 
    PubMed 
    Article 

    Google Scholar 
    Delmont, T. O. et al. Nitrogen-fixing populations of Planctomycetes and Proteobacteria are abundant in surface ocean metagenomes. Nat. Microbiol. 3, 804–813 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Aylward, F. O. & Santoro, A. E. Heterotrophic Thaumarchaea with small genomes are widespread in the dark ocean. mSystems 5, e00415–20 (2020).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Rodriguez-R, L. M. & Konstantinidis, K. T. The enveomics collection: a toolbox for specialized analyses of microbial genomes and metagenomes. Preprint at PeerJ https://doi.org/10.7287/peerj.preprints.1900v1 (2016).Konstantinidis, K. T. & Tiedje, J. M. Towards a genome-based taxonomy for prokaryotes. J. Bacteriol. 187, 6258–6264 (2005).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Yoon, S.-H., Ha, S.-M., Lim, J., Kwon, S. & Chun, J. A large-scale evaluation of algorithms to calculate average nucleotide identity. Antonie Van Leeuwenhoek 110, 1281–1286 (2017).CAS 
    PubMed 
    Article 

    Google Scholar 
    Lee, I., Ouk Kim, Y., Park, S.-C. & Chun, J. OrthoANI: an improved algorithm and software for calculating average nucleotide identity. Int. J. Syst. Evol. Microbiol. 66, 1100–1103 (2016).CAS 
    PubMed 
    Article 

    Google Scholar 
    Lee, M. D. GToTree: a user-friendly workflow for phylogenomics. Bioinformatics 35, 4162–4164 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Katoh, K. & Standley, D. M. MAFFT Multiple Sequence Alignment Software Version 7: improvements in performance and usability. Mol. Biol. Evol. 30, 772–780 (2013).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Capella-Gutiérrez, S., Silla-Martínez, J. M. & Gabaldón, T. trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25, 1972–1973 (2009).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Price, M. N., Dehal, P. S. & Arkin, A. P. FastTree 2 – approximately maximum-likelihood trees for large alignments. PLoS ONE 5, e9490 (2010).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Letunic, I. & Bork, P. Interactive tree of life (iTOL) v3: an online tool for the display and annotation of phylogenetic and other trees. Nucleic Acids Res. 44, W242–W245 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

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

    Google Scholar 
    Seemann, T. barrnap 0.9: Rapid Ribosomal RNA Prediction (2018); https://github.com/tseemann/barrnapFu, L., Niu, B., Zhu, Z., Wu, S. & Li, W. CD-HIT: accelerated for clustering the next-generation sequencing data. Bioinformatics 28, 3150–3152 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Stamatakis, A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30, 1312–1313 (2014).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Ronquist, F. et al. 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 Scholar 
    Warren, D. L., Geneva, A. J. & Lanfear, R. RWTY (R We There Yet): an R package for examining convergence of Bayesian phylogenetic analyses. Mol. Biol. Evol. 34, 1016–1020 (2017).CAS 
    PubMed 

    Google Scholar 
    Bi, D. et al. SecReT4: a web-based bacterial type IV secretion system resource. Nucleic Acids Res. 41, D660–D665 (2013).CAS 
    PubMed 
    Article 

    Google Scholar 
    Darriba, D., Taboada, G. L., Doallo, R. & Posada, D. ProtTest 3: fast selection of best-fit models of protein evolution. Bioinformatics 27, 1164–1165 (2011).CAS 
    PubMed 
    Article 

    Google Scholar 
    Sunagawa, S. et al. Structure and function of the global ocean microbiome. Science 348, 1261359 (2015).PubMed 
    Article 
    CAS 

    Google Scholar 
    Tully, B. J., Graham, E. D. & Heidelberg, J. F. The reconstruction of 2,631 draft metagenome-assembled genomes from the global oceans. Sci. Data 5, 170203 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Olson, D. K., Yoshizawa, S., Boeuf, D., Iwasaki, W. & DeLong, E. F. Proteorhodopsin variability and distribution in the North Pacific Subtropical Gyre. ISME J. 12, 1047–1060 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Philosof, A. & Béjà, O. Bacterial, archaeal and viral-like rhodopsins from the Red Sea. Environ. Microbiol. Rep. 5, 475–482 (2013).CAS 
    PubMed 
    Article 

    Google Scholar 
    Boeuf, D., Audic, S., Brillet-Guéguen, L., Caron, C. & Jeanthon, C. MicRhoDE: a curated database for the analysis of microbial rhodopsin diversity and evolution. Database 2015, bav080 (2015).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Robert, X. & Gouet, P. Deciphering key features in protein structures with the new ENDscript server. Nucleic Acids Res. 42, W320–W324 (2014).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Demir-Hilton, E. et al. Global distribution patterns of distinct clades of the photosynthetic picoeukaryote Ostreococcus. ISME J. 5, 1095–1107 (2011).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Parada, A. E., Needham, D. M. & Fuhrman, J. A. Every base matters: assessing small subunit rRNA primers for marine microbiomes with mock communities, time-series and global field samples. Environ. Microbiol. 18, 1403–1414 (2016).CAS 
    PubMed 
    Article 

    Google Scholar 
    Barbera, P. et al. EPA-ng: massively parallel evolutionary placement of genetic sequences. Syst. Biol. 68, 365–369 (2019).PubMed 
    Article 

    Google Scholar 
    Amaral-Zettler, L. A., McCliment, E. A., Ducklow, H. W. & Huse, S. M. A method for studying protistan diversity using massively parallel sequencing of V9 hypervariable regions of small-subunit ribosomal RNA genes. PLoS ONE 4, e6372 (2009).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Needham, D. M. & Fuhrman, J. A. Pronounced daily succession of phytoplankton, archaea and bacteria following a spring bloom. Nat. Microbiol. 1, 16005 (2016).CAS 
    PubMed 
    Article 

    Google Scholar 
    Xia, L. C. et al. Extended local similarity analysis (eLSA) of microbial community and other time series data with replicates. BMC. Syst. Biol. 5, S15 (2011).
    Google Scholar 
    Choi, H. M. T. et al. Third-generation in situ hybridization chain reaction: multiplexed, quantitative, sensitive, versatile, robust. Development 145, dev165753 (2018).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Schramm, A., Fuchs, B. M., Nielsen, J. L., Tonolla, M. & Stahl, D. A. Fluorescence in situ hybridization of 16S rRNA gene clones (Clone-FISH) for probe validation and screening of clone libraries. Environ. Microbiol. 4, 713–720 (2002).CAS 
    PubMed 
    Article 

    Google Scholar 
    Weiss, S. et al. Correlation detection strategies in microbial data sets vary widely in sensitivity and precision. ISME J. 10, 1669–1681 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Gloor, G. B., Macklaim, J. M., Pawlowsky-Glahn, V. & Egozcue, J. J. Microbiome datasets are compositional: and this is not optional. Front. Microbiol. 8, 2224 (2017).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Darling, A. C. E., Mau, B., Blattner, F. R. & Perna, N. T. Mauve: multiple alignment of conserved genomic sequence with rearrangements. Genome Res. 14, 1394–1403 (2004).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar  More

  • in

    Even a small nuclear war threatens food security

    Nuclear weapons obliterate targets. The soot ejected into the stratosphere spreads, changing global weather patterns. When weapons are especially high yielding, the resultant soot could trigger global famine.About 66 million years ago, roughly three-quarters of all species on Earth died when a 10–15-km-diameter asteroid travelling at 72,000 km h−1 struck at Chicxulub, Mexico1. Sulfates and soot lofted high in the atmosphere, cutting off sunlight. The Earth cooled, weather changed and primary productivity crashed. While the best-known victims of the asteroid impact were dinosaurs, the resultant food scarcity impacted the entire Earth; those not affected immediately by the impact eventually died from starvation. Any mechanism that can loft massive quantities of aerosols high into the atmosphere, such as massive volcanic explosions2 or nuclear wars3, can interfere with the weather globally and change world food security. More

  • in

    Pollinator and floral odor specificity among four synchronopatric species of Ceropegia (Apocynaceae) suggests ethological isolation that prevents reproductive interference

    Fly pollination in African and in Asian Ceropegia
    The four Ceropegia we investigated were pollinated by Diptera of two families, Chloropidae and Milichiidae, the latter represented by two genera, Milichiella and Neophyllomyza. Pollinator assemblages were significantly different between all pairs of plant species. With regard to the pollinator specialization, there are two aspects: functional specialization, i.e. interacting with a functional group of pollinators, and ecological specialization, i.e. interacting with relatively few species of pollinators. We consider the four Ceropegia species in our study as equally functionally specialized since they are pollinated by Chloropidae and Milichiidae flies, both of which are long-tongued flies and often kleptoparasites. Therefore, the way these flies interact with Ceropegia flowers may exert similar selective pressures on floral traits. Functional specialization on pollination by Diptera of only one or two families has been generally documented in Ceropegia21,22. However, ecological specialization18,19 is so far infrequent in the genus. It has been shown by Heiduk et al.24 for other species of Ceropegia from East Africa and South Africa that they are not ecologically highly specialized, because the functional group of pollinators often included several fly species. A different trend is revealed here: three of the four are ecologically specialized on pollination by only two, species-specific, pollinating fly morphospecies. Ceropegia citrina, which only relied on two fly morphospecies of the same genus, i.e., Neophyllomyza, is the most specialized. In contrast, C. boonjarasii is most generalist, being pollinated by six Chloropidae morphospecies. Each of the other two plant species, C. acicularis and C. tenuicaulis, was pollinated by two fly morphospecies of different families. In terms of the diversity of pollinator species, these two were thus more ecologically specialized than C. boojarasii, but less than C. citrina. Both Chloropidae and Milichiidae were also reported as pollinators of C. thaithongiae, another Thai endemic species33. Neophyllomyza flies were found to be pollinators of C. dolichophylla in China24,26 and visitors of C. nilotica in southern Africa21, while Milichiella flies were visitors of C. arabica var. powysii in East Africa21 and C. sandersonii in South Africa23.Large-scale studies of Ceropegia (including the stapeliads s. str.) covering the five major centers of diversity21,22, i.e., the Indian subcontinent (barring Southeast Asia), the Arabian Peninsula, East Africa, southern Africa, and West Africa, revealed flies of 16 families as pollinators. These studies also showed that the proportional use of different fly families as pollinators did not differ across all centers of diversity, thus no difference was found in this respect between Ceropegia from Africa and Asia. Moreover, in all regions, Chloropidae and Milichiidae were important pollinators. Note that the pollinator families used by Ceropegia s. str. were very different from those of the stapeliads s. str.Pollinator attraction and pollination strategyA main driving force behind floral visitation by adult Diptera, which can subsequently lead to successful pollination, is their requirement for sugar as a primary energy source, which can be obtained from floral nectar, and/or nutrients necessary for reproduction, which can be obtained from pollen35. Ceropegia flowers, however, do not offer pollen that can be consumed by flies, and are generally known to be nectarless. It has been suggested that they attract fly pollinators by deception through food source or brood site mimicry20,21,24. Nectar has thus far only been shown to be present in an African species (C. ampliata25) and a species endemic to Thailand, C. thaithongiae33. Also, a preliminary study on C. acicularis and C. tenuicaulis revealed the presence of viscous nectar in very small volume. In both species, nectar contained about 60% of fructose and 40% of glucose (unpublished data). Despite the presence of small quantities of nectar, these species can still be regarded as deceptive since they advertise another reward via scent mimicry21,22,24. As Ceropegia flowers trap flies for several hours or days, the function of nectar has been thought to be to maintain pollinators alive until they are released22,29,36. In the four Ceropegia species studied here, oviposition has not been observed; therefore, brood-site deception is unlikely.Recently, prey mimicry has been described in C. sandersonii23 and C. dolichophylla26. Both plant species were experimentally shown to be kleptomyiophilous, targeting their kleptoparasitic Desmometopa (Milichiidae) pollinators through the emission of VOCs that mimicked a particular food source for female flies. Many species of Choropidae and Milichiidae are kleptoparasites, attracted by various types of prey, frequently true bugs, killed by other arthropods. There is increasing evidence that flowers pollinated by such flies are kleptomyiophilous22,23,24,28. Typically, only females are kleptoparasitic26,37, because they require protein for egg maturation. The overwhelming majority (99%) of Diptera found in flowers of the four studied Ceropegia species were female, which is in accordance with previous findings for floral visitors of this plant group21,22. Forty percent of all major compounds identified in the four Ceropegia species are known to be released from insects, in particular by certain Hemiptera (Coreidae, Miridae, Pentatomidae). For example, (Z)-3-octen-1-yl acetate, (E)-2-octen-1-yl acetate and butyl butyrate were reported from Miridae38,39, and (E)-2-decen-1-yl acetate from Pentatomidae40. Hexyl butyrate, which was detected in the floral scents of C. citrina in a large proportion, is known to be released in prey defense secretions and responsible for the attraction of kleptoparasitic Neophyllomyza flies39,41,42,43 Hexyl acetate and 1-hexanol, the main compounds in the floral scents respectively emitted by C. tenuicaulis and C. acicularis, are known as components of the secretion of Coreidae bugs44,45. Interestingly, a species of Milichiella, the genus of flies that pollinated these two Ceropegia species, was reported to be attracted to various crushed Coreidae and Pentatomidae46. Certain VOCs have also been reported in other Ceropegia species, e.g., (E)-2-hexenyl acetate in C. sandersonii23, which is pollinated by kleptoparasitic Desmometopa flies and also visited by Milichiella flies. Aristolochia rotunda (Aristolochiaceae)28, a species with trap flowers similar to those of Ceropegia, emits volatiles similar to those of freshly killed Miridae bugs that were shown to attract its pollinating kleptoparasitic Chloropidae flies. These components included, inter alia, hexyl butyrate and octyl butyrate, which were present in C. acicularis, whose main pollinators we have also shown to be Chloropidae flies. Pollination by kleptoparasitic flies and emission of compounds also released in insect secretions support our assumption that, like other Ceropegia spp., C. acicularis, C. boonjarasii, C. citrina and C. tenuicaulis are kleptomyiophilous, and it is likely that VOCs play a major role in attracting their dipteran pollinators from a distance. In our study, a clear distinction in both the chemical profiles of VOC emissions of the four Ceropegia species and the identities of pollinating flies attracted to them suggests that each of them, if truly kleptomyiophilous, mimicks a different model. Further investigation should determine which compounds actually attract the particular pollinating flies, and which insects (models) are potentially mimicked by each species.In addition to VOCs, which appear to play a key functional role, morphological attributes, such as shape and size of floral parts, colors and patterns or other functional traits of flowers, which were very distinctive between the investigated Ceropegia species, may also affect the attractiveness of flowers. The experiments performed here on C. boonjarasii support the hypothesis that the vibratile trichomes on the corolla lobe tips play a role in attracting the fly pollinators of this species. In their natural habitats, Ceropegia generally grow close to the ground, hidden among the tall grasses or in small bushes47. It has therefore been suggested that olfactory cues are more important than visual cues in pollinator attraction20. However, in the case of the twiner C. boonjarasii, which climbs higher up in the vegetation and can present its flowers at a level up to 200 cm above the ground, the visually attractive vibratile trichomes of its flowers appear well adapted to its growth habit and the place where it grows, playing a key role in short-distance attraction, as proposed by Vogel20. Similar motile floral appendages48 are known in other Asclepiadoideae, including Stapelia spp. (Ceropegieae)49, and in plants of unrelated angiosperm families, e.g., Trichosalpinx spp. (Orchidaceae, Pleurothallidinae)50, pollinated by Diptera28,31,50.Our video recordings of flies trapped inside the flowers of C. tenuicaulis (Video 2) and C. acicularis (Video 3) are the first such recordings for Ceropegia trap flowers and showed that flies actively moved downwards to the base of the corolla inflation. This has been described by Vogel20 as positive phototaxis due to lighter-colored epidermal cells (“light windows”) around the gynostegium that direct the flies to move into contact with the reproductive organs, where they then remove pollinaria (the corpusculum being attached to the insect’s proboscis) or deposit pollinia (the guide rails take up the insertion crest of a pollinium). This uniform placement of pollinaria on different regions of insect mouthparts (e.g. rostrum, labellum or trichomes on lip pads) has been shown to be a phylogenetically conserved trait in the entire tribe Ceropegieae, which use flies, wasps, or beetles as pollinators20,21,22,29,49. In this regard, nectar (known to be present in certain species, see above) or probably minimal amounts of sugar-containing secretion contained in the cups formed by the corona appendages around the central gynostegium, may play a role in positioning the pollinators correctly.Entrapment of flies by the tubular corolla and guidance of trapped flies to the floral reproductive organs by the light windows are the important mechanisms ensuring successful pollination in Ceropegia. The circumstance observed on C. acicularis—most of whose Chloropidae pollinators moved purposefully downward into the corolla tube, but often left the flower after a short while—implies that the flower of this species is less efficient in trapping flies. Interestingly, the main pollinators of this species preferentially visited its flowers at the end of the day, when the light intensity had decreased drastically. Therefore, once flies entered the flower, whose opening at the top of the corolla tube (i.e., the fly’s exit) was already obscured by the darkness outside, their directional movement in response to light was impeded, forcing flies to stay overnight inside the flower. Whether these circumstances reflect coincidence or alternatively specialization for pollination by flies that seek overnight shelter, they seem to mitigate the effect of inefficient entrapment of flies by flowers of C. acicularis.Reproductive isolationThe four species of Ceropegia under consideration were found in the same habitat with individuals of different species occurring close to each other, and had largely overlapping flowering phenology. All species seem to share the same pollination strategy and attracted flies of only one or two families (Chloropidae and Milichiidae) with similar ecological traits. Their flowers were receptive all day long, and visits of fly pollinators, although varying according to the Ceropegia species they visited, largely overlapped. Overall, the combination of these factors might lead to pollinator sharing and, in turn, to interspecific pollinia transfer. Hereof, Meve and Liede-Schumann51 suggested that hybridization mediated by small flies was most likely responsible for morphological transitions between the taxa of Ceropegia s. str. and those of Brachystelma s. str. (recently merged into the former). Moreover, one of the authors (M. Kidyoo, unpublished data) found putative hybrids in various Ceropegia populations across Thailand. Thus, natural hybridization in the genus is known from co-occurring species. Nevertheless, in continued observations from 2014 to 2019, no putative hybrid plant individuals (i.e., with vegetative and floral morphology intermediate between the four studied species) have ever been found in Pha Taem National Park, indicating that species integrity is well maintained through efficient reproductive isolation among the four species. Whether or not post-zygotic reproductive isolation occurs, we expect pre-zygotic isolation mechanisms to have evolved because of the relatively high value of each pollinarium (only five per flower and few flowers per plant).The four species have diverged over time in particular floral traits. There are sharp floral morphological differences in shape, color patterns and other functional traits such as vibratile trichomes, which were shown to contribute to pollinator attraction in C. boonjarasii. Some similarities between the VOCs emitted by flowers of the four species suggest that the identities and proportions of compounds have changed over time due to selection exerted by different flies at this locality, which are all likely to be kleptoparasites but with different preferences for certain volatiles. Overall, specific pollinator attraction mediated by olfactory and visual floral cues—‘ethological isolation’—is a crucial pre-zygotic isolation mechanism to avoid interspecific visitation and pollinator limitation due to sharing. A contrasting case has been found in two other Ceropegia species, each native to different geographical regions, C. sandersonii from South Africa and C. dolichophylla from China. In that case, there is thus no selection pressure for divergence. In their natural habitats, each species uses different floral scent profiles to attract fly pollinators of the same genus, i.e., the kleptoparasitic Desmometopa spp. It has thus been proposed that these two Ceropegia exploit different olfactory preferences of flies of the same genus23,24,26.Pre-zygotic isolation can also be achieved through ‘mechanical isolation’17, requiring morphological fit either between flower and pollinator and/or between male (pollinium, in particular insertion crest) and female (guide rails) reproductive organs, known as a lock-and-key mechanism52, which can be achieved through various morphological adaptations. The role of floral proportions in regulating the fly sizes that can function as pollinators has been suggested in Ceropegia taxa21. It has also been highlighted in a stem-succulent, open-flowered stapeliad, Orbea lutea subsp. lutea, the head width of whose effective pollinating Atherigona (Muscidae) flies is small enough to fit and probe the cavity formed by the inner corona lobes beneath the guide rails for nectar53. In the present study, the size of the floral entrance, i.e. the smallest diameter of the corolla tube, was thought to be the first filter screening out visitors of the four Ceropegia flowers. However, no effect of corolla tube diameter on the fly sizes was detected. Each fly morphospecies that pollinates one of these four Ceropegia species was small enough to enter the tubular corolla of the other co-occurring species. Moreover, the pollinator exchange experiment performed in this study demonstrated that, in spite of distinctive corona structure (shape, size and spatial arrangement of corona lobes) between the investigated Ceropegia species, which concealed to varying extents the central floral reproductive organs, all fly morphospecies tested, including flies that were not natural pollinators (even a laboratory strain of D. melanogaster, a widespread Diptera of a family completely different from the identified pollinators), were able to remove the pollinaria, even those of C. citrina that appeared to be well hidden (Fig. 2G–I). Therefore, corona structure seems not to restrict visitors from removing the pollinaria. The fact that some removed pollinaria fell off and were found free inside the flowers shows that the pollinaria were not firmly attached to the experimental non-natural visitors. It is thus likely that their micromorphology did not fit properly that of the plant’s reproductive organs. Therefore, non-specific flies can lose pollinaria in flowers and are unlikely to be able to actually insert pollinia. Whatever the fate of the removed pollinia, they are lost from the flower if they do not reach another flower of the same species. The fact that pollinaria can be removed by flies that are not naturally attracted to and visiting flowers suggests that there is selective pressure for efficient ethological isolation. The distinct floral scents among Ceropegia species evidenced in this study might be the result of such selective pressure.
    Another aspect of mechanical isolation is the lock-and-key relation between the guide rails (lock) and pollinium (key) that has been reported from several Asclepiadoideae, e.g., Cynanchum s. l. (former Sarcostemma)54 and the stapeliads32. Other studies, however, confirmed the lack of such a precise mechanism, e.g., in Asclepias55 and Vincetoxicum s. l. (former Tylophora)52. Here, the preliminary study of morphology of pollinaria showed strong differences between the studied Ceropegia species in the shape and size of corpuscula, translator arms and pollinia, especially the insertion crest, which is the portion that is inserted as a “key” between the guide rails (the “lock”). This result is suggestive of a possible lock-and-key mechanism.Although this first study has not yet reached a full conclusion about the mechanisms enforcing the reproductive boundaries between the four Ceropegia species, it advances our understanding of how these investigated species maintain their integrity, and highlights the crucial role of ethological isolation. Other core elements of work need to be done to exactly examine the risk of interspecific pollinia transfers and hybridization. On the one hand, the morphological fit between the guide rails and pollinia has to be inspected by a detailed micromorphological study. On the other, further field experiments must be conducted to test if a pollinator of a given plant species, e.g., C. acicularis, can insert pollinia of this plant into the guide rails of any of the other three ‘wrong’ Ceropegia species (and all possible combinations thereof). A simpler experiment would test if those fly species that are attracted to more than one species (e.g., Milichiella sp. 1) are actually capable of exporting pollen to flowers of the wrong species. More

  • in

    The overlooked role of a biotin precursor for marine bacteria – desthiobiotin as an escape route for biotin auxotrophy

    Wakil SJ, Titchener EB, Gibson DM. Evidence for the participation of biotin in the enzymic synthesis of fatty acids. Biochim Biophys Acta. 1958;29:225–6.CAS 
    PubMed 
    Article 

    Google Scholar 
    Lardy HA, Peanasky R. Metabolic functions of biotin. Physiol Rev. 1953;33:560–5.CAS 
    PubMed 
    Article 

    Google Scholar 
    Zeczycki TN, Menefee AL, Adina-Zada A, Jitrapakdee S, Surinya KH, Wallace JC, et al. Novel Insights into the biotin carboxylase domain reactions of pyruvate carboxylase from Rhizobium etli. Biochemistry. 2011;50:9724–37.CAS 
    PubMed 
    Article 

    Google Scholar 
    Waldrop GL, Holden HM, Maurice MS. The enzymes of biotin dependent CO2 metabolism: What structures reveal about their reaction mechanisms. Protein Sci Publ Protein Soc. 2012;21:1597–619.CAS 
    Article 

    Google Scholar 
    Entcheva P, Phillips DA, Streit WR. Functional analysis of Sinorhizobium meliloti genes involved in biotin synthesis and transport. Appl Environ Microbiol. 2002;68:2843–8.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    E. Webb M, Marquet A, R. Mendel R, Rébeillé F, G. Smith A. Elucidating biosynthetic pathways for vitamins and cofactors. Nat Prod Rep. 2007;24:988–1008.Article 
    CAS 

    Google Scholar 
    Streit WR, Entcheva P. Biotin in microbes, the genes involved in its biosynthesis, its biochemical role and perspectives for biotechnological production. Appl Microbiol Biotechnol. 2003;61:21–31.CAS 
    PubMed 
    Article 

    Google Scholar 
    Tang YZ, Koch F, Gobler CJ. Most harmful algal bloom species are vitamin B1 and B12 auxotrophs. Proc Natl Acad Sci. 2010;107:20756–61.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Croft MT, Warren MJ, Smith AG. Algae need their vitamins. Eukaryot Cell. 2006;5: 1175–83.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Sañudo-Wilhelmy SA, Gómez-Consarnau L, Suffridge C, Webb EA. The role of B vitamins in marine biogeochemistry. Annu Rev Mar Sci. 2014;6:339–67.Article 

    Google Scholar 
    Rodionov DA, Arzamasov AA, Khoroshkin MS, Iablokov SN, Leyn SA, Peterson SN, et al. Micronutrient requirements and sharing capabilities of the human gut microbiome. Front Microbiol. 2019;10:1316.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Croft MT, Lawrence AD, Raux-Deery E, Warren MJ, Smith AG. Algae acquire vitamin B12 through a symbiotic relationship with bacteria. Nature. 2005;438:90–3.CAS 
    PubMed 
    Article 

    Google Scholar 
    Carini P, Campbell EO, Morré J, Sañudo-Wilhelmy SA, Cameron Thrash J, Bennett SE, et al. Discovery of a SAR11 growth requirement for thiamin’s pyrimidine precursor and its distribution in the Sargasso Sea. ISME J. 2014;8:1727–38.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Paerl RW, Sundh J, Tan D, Svenningsen SL, Hylander S, Pinhassi J, et al. Prevalent reliance of bacterioplankton on exogenous vitamin B1 and precursor availability. Proc Natl Acad Sci USA. 2018;115:E10447–56.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Helliwell KE, Lawrence AD, Holzer A, Kudahl UJ, Sasso S, Kräutler B, et al. Cyanobacteria and eukaryotic algae use different chemical variants of vitamin B12. Curr Biol. 2016;26:999–1008.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Heal KR, Carlson LT, Devol AH, Armbrust EV, Moffett JW, Stahl DA, et al. Determination of four forms of vitamin B12 and other B vitamins in seawater by liquid chromatography/tandem mass spectrometry. Rapid Commun Mass Spectrom RCM. 2014;28:2398–404.CAS 
    PubMed 
    Article 

    Google Scholar 
    Wienhausen G, Dlugosch L, Jarling R, Wilkes H, Giebel H-A, Simon M. Availability of vitamin B12 and its lower ligand intermediate α-ribazole impact prokaryotic and protist communities in oceanic systems. ISME J. 2022;16:2002–14.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Cooper MB, Kazamia E, Helliwell KE, Kudahl UJ, Sayer A, Wheeler GL, et al. Cross-exchange of B-vitamins underpins a mutualistic interaction between Ostreococcus tauri and Dinoroseobacter shibae. ISME J. 2019;13:334–45.CAS 
    PubMed 
    Article 

    Google Scholar 
    Cruz-López R, Maske H, Yarimizu K, Holland NA. The B-Vitamin mutualism between the Dinoflagellate Lingulodinium polyedrum and the bacterium Dinoroseobacter shibae. Front Mar Sci. 2018;5:274.Article 

    Google Scholar 
    Blifernez-Klassen O, Klassen V, Wibberg D, Cebeci E, Henke C, Rückert C, et al. Phytoplankton consortia as a blueprint for mutually beneficial eukaryote-bacteria ecosystems based on the biocoenosis of Botryococcus consortia. Sci Rep. 2021;11:1726.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Burkholder PR, Lewis S. Some patterns of B vitamin requirements among neritic marine bacteria. Can J Microbiol. 1968;14:537–43.CAS 
    PubMed 
    Article 

    Google Scholar 
    Leonian LH, Lilly VG. Conversion of desthiobiotin into biotin or biotinlike substances by some microorganisms. J Bacteriol. 1945;3:291–7.Article 

    Google Scholar 
    Lilly VG, Leonian LH. The anti-biotin effect of desthiobiotin. Science. 1944;99:205–6.CAS 
    PubMed 
    Article 

    Google Scholar 
    Dittmer K, Melville DB, du Vigneaud V. The possible synthesis of biotin from desthiobiotin by yeast and the anti-biotin effect of desthiobiotin for L. casei. Science. 1944;99:203–5.CAS 
    PubMed 
    Article 

    Google Scholar 
    Prakash O, Eisenberg MA. Active transport of biotin in Escherichia coli K-12. J Bacteriol. 1974;120:785–91.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Longnecker K, Sievert SM, Sylva SP, Seewald JS, Kujawinski EB. Dissolved organic carbon compounds in deep-sea hydrothermal vent fluids from the East Pacific Rise at 9°50′N. Org Geochem. 2018;125:41–49.CAS 
    Article 

    Google Scholar 
    Johnson WM, Soule MCK, Longnecker K, Bhatia MP, Hallam SJ, Lomas MW, et al. Insights into the controls on metabolite distributions along a latitudinal transect of the western Atlantic Ocean. 2021. bioRxiv. 2021.03.09.434501; https://doi.org/10.1101/2021.03.09.434501.Suffridge CP, Gómez‐Consarnau L, Monteverde DR, Cutter L, Arístegui J, Alvarez‐Salgado XA, et al. B vitamins and their congeners as potential drivers of microbial community composition in an oligotrophic marine ecosystem. J Geophys Res Biogeosciences. 2018;123:2890–907.CAS 
    Article 

    Google Scholar 
    Suffridge C, Cutter L, Sañudo-Wilhelmy SA. A new analytical method for direct measurement of particulate and dissolved B-vitamins and their congeners in seawater. Front Mar Sci. 2017;4:2296.Article 

    Google Scholar 
    Natarajan KV. Distribution of thiamine, biotin, and niacin in the sea. Appl Microbiol. 1968;16:366–9.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Ohwada K. Bioassay of biotin and its distribution in the sea. Mar Biol. 1972;14:10–17.CAS 
    Article 

    Google Scholar 
    Luo H, Moran MA. Evolutionary ecology of the marine roseobacter clade. Microbiol Mol Biol Rev. 2014;78:573–87.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Guillard RRL, Ryther JH. Studies of marine planktonic diatoms: i. Cyclotella nana hustedt, and Detonula confervacea (cleve) gran. Can J Microbiol. 1962;8:229–39.CAS 
    PubMed 
    Article 

    Google Scholar 
    Lunau M, Lemke A, Walther K, Martens-Habbena W, Simon M. An improved method for counting bacteria from sediments and turbid environments by epifluorescence microscopy. Environ Microbiol. 2005;7:961–8.PubMed 
    Article 

    Google Scholar 
    Osterholz H, Niggemann J, Giebel H-A, Simon M, Dittmar T. Inefficient microbial production of refractory dissolved organic matter in the ocean. Nat Commun. 2015;6:7422.CAS 
    PubMed 
    Article 

    Google Scholar 
    Taga ME, Xavier KB. Methods for analysis of bacterial autoinducer-2 production. Curr Protoc Microbiol. 2011;Chapter 1:Unit1C.1.PubMed 

    Google Scholar 
    Cakić N, Kopke B, Rabus R, Wilkes H. Suspect screening and targeted analysis of acyl coenzyme A thioesters in bacterial cultures using a high-resolution tribrid mass spectrometer. Anal Bioanal Chem. 2021;413:3599–610.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Shelton AN, Seth EC, Mok KC, Han AW, Jackson SN, Haft DR, et al. Uneven distribution of cobamide biosynthesis and dependence in bacteria predicted by comparative genomics. ISME J. 2019;13:789–804.CAS 
    PubMed 
    Article 

    Google Scholar 
    Raes J, Korbel JO, Lercher MJ, von Mering C, Bork P. Prediction of effective genome size in metagenomic samples. Genome Biol. 2007;8:R10.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Brown CT, Hug LA, Thomas BC, Sharon I, Castelle CJ, Singh A, et al. Unusual biology across a group comprising more than 15% of domain Bacteria. Nature. 2015;523:208–11.CAS 
    PubMed 
    Article 

    Google Scholar 
    Dlugosch L, Poehlein A, Wemheuer B, Pfeiffer B, Badewien TH, Daniel R, et al. Significance of gene variants for the functional biogeography of the near-surface Atlantic Ocean microbiome. Nat Commun. 2022;13:456.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Nurk S, Meleshko D, Korobeynikov A, Pevzner PA. metaSPAdes: a new versatile metagenomic assembler. Genome Res. 2017;27:824–34.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Hyatt D, Chen G-L, LoCascio PF, Land ML, Larimer FW, Hauser LJ. Prodigal: prokaryotic gene recognition and translation initiation site identification. BMC Bioinforma. 2010;11:119.Article 
    CAS 

    Google Scholar 
    Edgar RC. Search and clustering orders of magnitude faster than BLAST. Bioinformatics. 2010;26:2460–1.CAS 
    PubMed 
    Article 

    Google Scholar 
    Menzel P, Ng KL, Krogh A. Fast and sensitive taxonomic classification for metagenomics with Kaiju. Nat Commun. 2016;7:11257.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Mende DR, Letunic I, Huerta-Cepas J, Li SS, Forslund K, Sunagawa S, et al. proGenomes: a resource for consistent functional and taxonomic annotations of prokaryotic genomes. Nucleic Acids Res. 2017;45:D529–D534.CAS 
    PubMed 
    Article 

    Google Scholar 
    Wildiers E. Nouvelle substance indispensable au developpement de la levure. La Cellule. 1901;18:311–33.
    Google Scholar 
    du Vigneaud V, Hofmann K, Melville DB, Rachele JR. The preparation of free crystaline biotin. J Biol Chem. 1941;140:763–6.Article 

    Google Scholar 
    Firestone BY, Koser SA. Growth promoting effect of some biotin analogues for Candida albicans. J Bacteriol. 1960;79:674–6.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Gómez-Consarnau L, Sachdeva R, Gifford SM, Cutter LS, Fuhrman JA, Sañudo-Wilhelmy SA, et al. Mosaic patterns of B-vitamin synthesis and utilization in a natural marine microbial community. Environ Microbiol. 2018;20:2809–23.PubMed 
    Article 
    CAS 

    Google Scholar 
    Carini PJ. Genome-enabled investigation of the minimal growth requirements andphosphate metabolism for Pelagibacter marine bacteria. Oregon State University 2014; dissertation: 9593v081hBertrand EM, Saito MA, Rose JM, Riesselman CR, Lohan MC, Noble AE, et al. Vitamin B12 and iron colimitation of phytoplankton growth in the Ross Sea. Limnol Oceanogr. 2007;52:1079–93.CAS 
    Article 

    Google Scholar 
    Azhar A, Booker G, Polyak S. Mechanisms of biotin transport. Biochem Anal Biochem. 2015,4:210.
    Google Scholar 
    Agarwal S, Dey S, Ghosh B, Biswas M, Dasgupta J. Mechanistic basis of vitamin B12 and cobinamide salvaging by the Vibrio species. Biochim Biophys Acta BBA—Proteins Proteom. 2019;1867:140–51.CAS 
    Article 

    Google Scholar 
    Giovannoni SJ, Cameron Thrash J, Temperton B. Implications of streamlining theory for microbial ecology. ISME J. 2014;8:1553–65.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Morris JJ, Lenski RE, Zinser ER. The black queen hypothesis: evolution of dependencies through adaptive gene loss. mBio. 2012;3:e00036–12.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Morris JJ. Black queen evolution: the role of leakiness in structuring microbial communities. Trends Genet. 2015;31:475–82.CAS 
    PubMed 
    Article 

    Google Scholar 
    Wienhausen G, Noriega-Ortega BE, Niggemann J, Dittmar T, Simon M. The exometabolome of two model strains of the roseobacter group: a marketplace of microbial metabolites. Front Microbiol. 2017;8:1985.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Wienhausen G, Paerl RW, Bittner M. Key knowledge gaps to fill at the cell-to-ecosystem level in marine B-vitamin cycling. Front Perspect. 2022;9:835.
    Google Scholar 
    Cohen NR A, Ellis K, Burns WG, Lampe RH, Schuback N, Johnson Z, et al. Iron and vitamin interactions in marine diatom isolates and natural assemblages of the Northeast Pacific Ocean. Limnol Oceanogr. 2017;62:2076–96.Article 
    CAS 

    Google Scholar 
    Marinov I, Doney SC, Lima ID. Response of ocean phytoplankton community structure to climate change over the 21st century: partitioning the effects of nutrients, temperature and light. Biogeosciences. 2010;7:3941–59.Article 

    Google Scholar 
    Thole S, Kalhoefer D, Voget S, Berger M, Engelhardt T, Liesegang H, et al. Phaeobacter gallaeciensis genomes from globally opposite locations reveal high similarity of adaptation to surface life. ISME J. 2012;6:2229–44.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Wagner-Döbler I, Ballhausen B, Berger M, Brinkhoff T, Buchholz I, Bunk B, et al. The complete genome sequence of the algal symbiont Dinoroseobacter shibae: a hitchhiker’s guide to life in the sea. ISME J. 2010;4:61–77.PubMed 
    Article 
    CAS 

    Google Scholar  More

  • in

    Accelerated marsh erosion following the Deepwater Horizon oil spill confirmed, ameliorated by planting

    We start by looking at our field erosion data, collected through 6 years post-spill (7 years total), to determine the duration of oiling effects, and to look for longer-term cleanup treatment influences. We previously observed increased erosion in our heavily oiled marsh sites for a 2-year period after the spill, the duration of our study at that time7. We also observed no major differences in erosion among oiled sites that were untreated versus those with manual cleanup treatments. In a separate but related experiment, using slightly different methods, there were indirect indications that mechanical cleanup treatments may have further worsened erosion and direct evidence that planting limited erosion in mechanically treated sites over a 1 year study period7.
    In the present study, our field-based comparisons of oiling/treatment categories included reference, oiled and untreated, oiled and manually treated, and oiled and mechanically treated sites. Untreated sites had no active cleanup (i.e., natural recovery), an approach which is commonly prescribed for oiled marshes (see Zengel et al.7 for background on the trade-offs of typical oiled marsh treatment tactics). Manual cleanup treatment involved raking, cutting, and removal of oiled wrack, oiled vegetation mats (laid over oiled and dead vegetation that was still rooted), and underlying thick oil on the substrate by small crews using hand tools, to remove surface oiling to the extent possible and to better expose remaining oil to natural weathering and degradation processes7. Hand crews used walking boards to minimize foot traffic on the marsh surface. Mechanical cleanup treatment involved mechanized grappling to remove oiled wrack and mechanized raking, cutting, and scraping to remove or reduce oiled vegetation mats and oil on the substrate. The mechanical treatments were applied using long-reach hydraulic arms mounted on shallow-draft barges and large airboats stationed just seaward of the marsh shoreline7,23. Mechanical treatment was aimed toward the same goals as manual cleanup but with anticipated increases in speed and scale; however, mechanical treatment can also be less precise, resulting in removal of soils, mixing of oil into the substrate, etc. Oiling conditions were highly consistent across all the oiled sites, characterized as heavy oiling using Shoreline Cleanup Assessment Technique (SCAT) methods7, although we agree that oiling in our sites could be considered “very heavy” as proposed by others19. Oiling conditions consisted of a continuous 6–13 meters (m) wide oiling band along the marsh shoreline, with heavily oiled wrack and vegetation mats overlying a 2–3 centimeter (cm) layer of emulsified oil on the marsh surface with ~ 90–100% oil cover7. The heavily oiled sites all experienced complete or near complete vegetation die-off, with vegetation recovery spanning multiple years7,23. Nearby reference sites on the same shoreline had lighter to no oiling, intact vegetation structure, and no cleanup treatments7. Further details on oiling conditions, cleanup treatments, and vegetation response are included in our prior papers (including photographs)7,23.Annual shoreline erosion rates were determined each year by ground surveys using tape measures and differentially corrected GPS (± 10 cm horizontal accuracy) to measure shoreline position along established transects. In the present study, we observed 147–198% greater marsh shoreline erosion for the oiled versus reference sites over 2 years (2010–2011 and 2011–2012), with no clear differences among oiled sites with or without cleanup treatments (Fig. 2, Supplementary Table S1). There was some indication that mechanical treatment may possibly have worsened erosion in some sites in 2011–2012, which matched our field observations and prior experiments7, though our sample size was small and highly variable in this case.Figure 2Field measured marsh shoreline erosion rates 2010–2016 (m yr−1). Data are means with 90% confidence intervals, n = 5 for Reference, 9 for Oiled-Untreated, 5 for Oiled-Manual, and 2–6 for Oiled-Mechanical treatments (n = 14–20 for Oiled sites combined) depending on year. Due to missing values, the desire to use as much data as possible, and the lack of clear differences among cleanup treatments, we pooled the oiled site data for statistical analysis. Marsh erosion rates differed among Reference and Oiled sites (F1,17 = 9.751, p = 0.006); among years (F2.32,39.40 = 2.703, p = 0.072); and for the interaction of oiling and year (F2.32,39.40 = 2.648, p = 0.076). Pairwise differences (Tukey’s test) among Reference and Oiled sites were observed for 2010–2011 (p = 0.003) and 2011–2012 (p = 0.001), but not for other years. See Supplementary Table S1 for detailed two-way mixed ANOVA results.Full size imagePotential causes for increased shoreline erosion in oiled marshes may lie most directly with the die-off of marsh vegetation, which baffles wave energy and binds marsh soils. Die-off of vegetation at the marsh edge was likely caused by several related factors, including thick persistent oiling covering all or most of the aboveground vegetation and soil surface, repetitive oiling, and penetration and mixing of oil into the soils, resulting in fouling and smothering effects on the plants such as interference with photosynthesis, gas exchange, thermal regulation, etc., leading to plant death. Our prior work showed substantial reductions (77–100%) in aboveground total plant cover (all species) and Spartina alterniflora cover (the dominant marsh species) in our oiled sites versus reference over 2010–2011 (and 45–99% reductions over 2011–2012)7. Belowground plant biomass was likely also reduced, although this was not measured in our study. However, belowground biomass was reduced in similarly oiled sites in studies by others5,9,14,24. Significant relationships have been established between marsh belowground biomass and soil shear strength (a measure of erosion potential)25, including in oiled marshes9, and belowground biomass is the main vegetation trait that resists erosion in coastal marshes26. Working in collaboration with us, Lin et al.9 conducted ancillary sampling in a subset of our study sites over the 2011–2012 period and provided their unpublished soil shear strength data (0–6 cm) for our use (their sampling did not include our planted sites). There was a 42% reduction in soil shear strength in the oiled sites relative to the reference sites (Fig. 3, Supplementary Table S2). Our reference site mean values are very similar to marsh soil shear strength values reported by others in our study region9,27, whereas our oiled site values are much lower. Thus, there is evidence that oiling affected both the marsh vegetation and the erodibility of the marsh soils, which likely led to the observed differences in marsh erosion between the reference and oiled sites. Oyster beds were not present near the marsh edge in our study area; thus, oyster cover impacts did not contribute to observed erosion differences between our oiled and reference sites (see Powers et al.17).Figure 3Field measured marsh soil shear strength 2011–2012 (KPa). Data are means with 90% confidence intervals, n = 4 for Reference and 13 for Oiled sites. Soil shear strength differences were observed among the Reference and Oiled sites (t6.29 = − 3.877, p = 0.007, Welch’s t-test). See Supplementary Table S2 for further details.Full size image More

  • in

    The gill transcriptome of threatened European freshwater mussels

    Allendorf, F. W., Hohenlohe, P. A. & Luikart, G. Genomics and the future of conservation genetics. Nature Reviews Genetics 2010 11:10 11, 697–709 (2010).CAS 

    Google Scholar 
    Formenti, G. et al. The era of reference genomes in conservation genomics. Trends in Ecology and Evolution 37, 197–202 (2022).PubMed 
    Article 

    Google Scholar 
    Hohenlohe, P. A., Funk, W. C. & Rajora, O. P. Population genomics for wildlife conservation and management. Molecular Ecology 30, 62–82 (2021).PubMed 
    Article 

    Google Scholar 
    Meek, M. H. & Larson, W. A. The future is now: Amplicon sequencing and sequence capture usher in the conservation genomics era. Molecular Ecology Resources 19, 795–803 (2019).PubMed 
    Article 

    Google Scholar 
    McCartney, M. A. et al. The genome of the zebra mussel, Dreissena polymorpha: a resource for comparative genomics, invasion genetics, and biocontrol. G3 Genes|Genomes|Genetics 12 (2022).Vaughn, C. C., Nichols, S. J. & Spooner, D. E. Community and foodweb ecology of freshwater mussels. 27, 409–423, https://doi.org/10.1899/07-058.1 (2015).Vaughn, C. C. Ecosystem services provided by freshwater mussels. Hydrobiologia 2017 810:1 810, 15–27 (2017).
    Google Scholar 
    Lopes-Lima, M. et al. Biology and conservation of freshwater bivalves: Past, present and future perspectives. Hydrobiologia 735, 1–13 (2014).Article 

    Google Scholar 
    Haag, W. R. North American Freshwater Mussels: Natural History, Ecology, and Conservation. (Cambridge University Press, 2012).Lopes-Lima, M. et al. Conservation status of freshwater mussels in Europe: state of the art and future challenges. Biological Reviews 92, 572–607 (2017).PubMed 
    Article 

    Google Scholar 
    Cuttelod, A., Seddon, M. & Neubert, E. European red list of non-marine molluscs. (Publications Office of the European Union Luxembourg, 2011).Lopes-Lima, M. et al. Conservation of freshwater bivalves at the global scale: diversity, threats and research needs. Hydrobiologia 810, 1–14 (2018).Article 

    Google Scholar 
    Gomes-dos-Santos, A. et al. The Crown Pearl: a draft genome assembly of the European freshwater pearl mussel Margaritifera margaritifera (Linnaeus, 1758). DNA Research, https://doi.org/10.1093/dnares/dsab002 (2021).Smith, C. H. A High-Quality Reference Genome for a Parasitic Bivalve with Doubly Uniparental Inheritance (Bivalvia: Unionida). Genome Biology and Evolution 13 (2021).Rogers, R. L. et al. Gene family amplification facilitates adaptation in freshwater unionid bivalve Megalonaias nervosa. Molecular Ecology 30, 1155–1173 (2021).CAS 
    PubMed 
    Article 

    Google Scholar 
    Renaut, S. et al. Genome Survey of the Freshwater Mussel Venustaconcha ellipsiformis (Bivalvia: Unionida) Using a Hybrid De Novo Assembly Approach. Genome Biology and Evolution 10, 1637–1646 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Roznere, I., Sinn, B. T. & Watters, G. T. The Amblema plicata Transcriptome as a Resource to Assess Environmental Impacts on Freshwater Mussels. Freshwater Mollusk Biology and Conservation 21, 57–64 (2018).
    Google Scholar 
    Wang, R. et al. Rapid development of molecular resources for a freshwater mussel, Villosa lienosa (Bivalvia:Unionidae), using an RNA-seq-based approach. 31, 695–708, https://doi.org/10.1899/11-149.1 (2015).Luo, Y. et al. Transcriptomic Profiling of Differential Responses to Drought in Two Freshwater Mussel Species, the Giant Floater Pyganodon grandis and the Pondhorn Uniomerus tetralasmus. PLOS ONE 9, e89481 (2014).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Patnaik, B. B. et al. Sequencing, De Novo Assembly, and Annotation of the Transcriptome of the Endangered Freshwater Pearl Bivalve, Cristaria plicata, Provides Novel Insights into Functional Genes and Marker Discovery. PLOS ONE 11, e0148622 (2016).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Wang, X., Liu, Z. & Wu, W. Transcriptome analysis of the freshwater pearl mussel (Cristaria plicata) mantle unravels genes involved in the formation of shell and pearl. Molecular Genetics and Genomics 292, 343–352 (2017).CAS 
    PubMed 
    Article 

    Google Scholar 
    Yang, Q. et al. Histopathology, antioxidant responses, transcriptome and gene expression analysis in triangle sail mussel Hyriopsis cumingii after bacterial infection. Developmental & Comparative Immunology 124, 104175 (2021).CAS 
    Article 

    Google Scholar 
    Bertucci, A. et al. Transcriptomic responses of the endangered freshwater mussel Margaritifera margaritifera to trace metal contamination in the Dronne River, France. Environmental Science and Pollution Research 24, 27145–27159 (2017).CAS 
    PubMed 
    Article 

    Google Scholar 
    Robertson, L. S., Galbraith, H. S., Iwanowicz, D., Blakeslee, C. J. & Cornman, R. S. RNA sequencing analysis of transcriptional change in the freshwater mussel Elliptio complanata after environmentally relevant sodium chloride exposure. Environmental Toxicology and Chemistry 36, 2352–2366 (2017).CAS 
    PubMed 
    Article 

    Google Scholar 
    Capt, C. et al. Deciphering the Link between Doubly Uniparental Inheritance of mtDNA and Sex Determination in Bivalves: Clues from Comparative Transcriptomics. Genome Biology and Evolution 10, 577–590 (2018).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Huang, D., Shen, J., Li, J. & Bai, Z. Integrated transcriptome analysis of immunological responses in the pearl sac of the triangle sail mussel (Hyriopsis cumingii) after mantle implantation. Fish & Shellfish Immunology 90, 385–394 (2019).CAS 
    Article 

    Google Scholar 
    Capt, C., Renaut, S., Stewart, D. T., Johnson, N. A. & Breton, S. Putative Mitochondrial Sex Determination in the Bivalvia: Insights From a Hybrid Transcriptome Assembly in Freshwater Mussels. Frontiers in Genetics 10, 840 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Chen, X., Bai, Z. & Li, J. The Mantle Exosome and MicroRNAs of Hyriopsis cumingii Involved in Nacre Color Formation. Marine Biotechnology 21, 634–642 (2019).CAS 
    PubMed 
    Article 

    Google Scholar 
    Cornman, R. S., Robertson, L. S., Galbraith, H. & Blakeslee, C. Transcriptomic Analysis of the Mussel Elliptio complanata Identifies Candidate Stress-Response Genes and an Abundance of Novel or Noncoding Transcripts. PLOS ONE 9, e112420 (2014).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Ganser, A. M., Newton, T. J. & Haro, R. J. Effects of elevated water temperature on physiological responses in adult freshwater mussels. Freshwater Biology 60, 1705–1716 (2015).Article 

    Google Scholar 
    Haney, A., Abdelrahman, H. & Stoeckel, J. A. Effects of thermal and hypoxic stress on respiratory patterns of three unionid species: implications for management and conservation. Hydrobiologia 847, 787–802 (2020).Article 

    Google Scholar 
    Geist, J. Strategies for the conservation of endangered freshwater pearl mussels (Margaritifera margaritifera L.): a synthesis of Conservation Genetics and Ecology. Hydrobiologia 644, 69–88 (2010).Article 

    Google Scholar 
    Moorkens, E., Cordeiro, J., Seddon, M. B. & von Proschwitz, T. Woolnough, D. Margaritifera margaritifera (Freshwater Pearl Mussel). The IUCN Red List of Threatened Species https://www.iucnredlist.org/species/12799/128686456 (2017).Lopes-Lima, M., Kebapçı, U. & van Damme, D. Unio crassus (Thick Shelled River Mussel). The IUCN Red List of Threatened Species https://www.iucnredlist.org/species/22736/42465628 (2014).Lopes-Lima, M. & Seddon, M. B. Unio mancus. The IUCN Red List of Threatened Species https://www.iucnredlist.org/species/22737/42466471 (2014).Araujo, R. Unio delphinus. The IUCN Red List of Threatened Species https://www.iucnredlist.org/species/195510/8975648 (2011).Díaz, S. et al. IPBES, 2019: Summary for policymakers of the global assessment report on biodiversity and ecosystem services of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services. Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services (2019).Bolger, A. M., Lohse, M. & Usadel, B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics 30, 2114–2120 (2014).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Song, L. & Florea, L. Rcorrector: Efficient and accurate error correction for Illumina RNA-seq reads. Gigascience 4, 48 (2015).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Kim, D., Song, L., Breitwieser, F. P. & Salzberg, S. L. Centrifuge: rapid and sensitive classification of metagenomic sequences. Genome Research 26, 1721–1729 (2016).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Grabherr, M. G. et al. Full-length transcriptome assembly from RNA-Seq data without a reference genome. Nature Biotechnology 2011 29:7 29, 644–652 (2011).CAS 

    Google Scholar 
    Haas, B. J. et al. De novo transcript sequence reconstruction from RNA-seq using the Trinity platform for reference generation and analysis. Nature Protocols 8, 1494–1512 (2013).CAS 
    PubMed 
    Article 

    Google Scholar 
    Agarwala, R. et al. Database resources of the National Center for Biotechnology Information. Nucleic Acids Research 44, D7–D19 (2016).CAS 
    Article 

    Google Scholar 
    Camacho, C. et al. BLAST+: Architecture and applications. BMC Bioinformatics 10, 1–9 (2009).Article 
    CAS 

    Google Scholar 
    Davidson, N. M. & Oshlack, A. Corset: Enabling differential gene expression analysis for de novo assembled transcriptomes. Genome Biology 15, 1–14 (2014).Article 
    CAS 

    Google Scholar 
    Simão, F. A., Waterhouse, R. M., Ioannidis, P., Kriventseva, E. V. & Zdobnov, E. M. BUSCO: Assessing genome assembly and annotation completeness with single-copy orthologs. Bioinformatics 31, 3210–3212 (2015).PubMed 
    Article 
    CAS 

    Google Scholar 
    Lang-Unnasch, N. Purification and properties of Plasmodium falciparum malate dehydrogenase. Molecular and Biochemical Parasitology 50, 17–25 (1992).CAS 
    PubMed 
    Article 

    Google Scholar 
    Bateman, A. et al. UniProt: the universal protein knowledgebase. Nucleic Acids Research 45, D158–D169 (2017).CAS 
    Article 

    Google Scholar 
    Punta, M. et al. The Pfam protein families database. Nucleic Acids Research 40, D290–D301 (2012).CAS 
    PubMed 
    Article 

    Google Scholar 
    Finn, R. D., Clements, J. & Eddy, S. R. HMMER web server: interactive sequence similarity searching. Nucleic Acids Research 39, W29–W37 (2011).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Dainat, J., Hereñú, D. & Pucholt, P. AGAT: Another Gff Analysis Toolkit to handle annotations in any GTF/GFF format, https://doi.org/10.5281/zenodo.4205393 (2020).Pruitt, K. D., Tatusova, T. & Maglott, D. R. NCBI reference sequences (RefSeq): a curated non-redundant sequence database of genomes, transcripts and proteins. Nucleic Acids Research 35, D61–D65 (2007).CAS 
    PubMed 
    Article 

    Google Scholar 
    Buchfink, B., Xie, C. & Huson, D. H. Fast and sensitive protein alignment using DIAMOND. Nature Methods 12, 59–60 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    NCBI Sequence Read Archive https://identifiers.org/ncbi/insdc.sra:SRP375793 (2022). NCBI BioProject, https://identifiers.org/ncbi/bioproject:PRJNA839062 (2022).Gomes-dos-Santos, A. The gill transcriptome of threatened European freshwater mussels, figshare, https://doi.org/10.6084/m9.figshare.19787566.v2 (2022).Machado, A. M. et al. The male and female gonad transcriptome of the edible sea urchin, Paracentrotus lividus: Identification of sex-related and lipid biosynthesis genes. Aquaculture Reports 22, 100936 (2022).Article 

    Google Scholar 
    Machado, A. M. et al. Liver transcriptome resources of four commercially exploited teleost species. Scientific Data 7, 1–9 (2020).Article 
    CAS 

    Google Scholar 
    Lehner, B. & Grill, G. Global river hydrography and network routing: Baseline data and new approaches to study the world’s large river systems. Hydrological Processes 27, 2171–2186 (2013).ADS 
    Article 

    Google Scholar  More

  • in

    Effects of decadal climate variability on spatiotemporal distribution of Indo-Pacific yellowfin tuna population

    Burrows, M. T. et al. The pace of shifting climate in marine and terrestrial ecosystems. Science 334, 652–655 (2011).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Cheung, W. W. L., Dunne, J., Sarmiento, J. L. & Pauly, D. Integrating ecophysiology and plankton dynamics into projected maximum fisheries catch potential under climate change in the Northeast Atlantic. ICES J. Mar. Sci. 68, 1008–1018 (2011).Article 

    Google Scholar 
    Muhling, B. A. et al. Potential impact of climate change on the Intra-Americas Sea: Part 2. Implications for Atlantic bluefin tuna and skipjack tuna adult and larval habitats. J. Mar. Syst. 148, 1–13 (2015).Article 

    Google Scholar 
    Erauskin-Extramiana, M. et al. Large-scale distribution of tuna species in a warming ocean. Glob. Change Biol. 25, 2043–2060 (2019).ADS 
    Article 

    Google Scholar 
    Cheung, W. W. et al. Large-scale redistribution of maximum fisheries catch potential in the global ocean under climate change. Glob. Change Biol. 16, 24–35 (2010).ADS 
    Article 

    Google Scholar 
    Townhill, B. L., Couce, E., Bell, J., Reeves, S. & Yates, O. Climate change impacts on Atlantic oceanic island tuna fisheries. Front. Mar. Sci. 8, 140 (2021).Article 

    Google Scholar 
    Wu, Y. L., Lan, K. W. & Tian, Y. J. Determining the effect of multiscale climate indices on the global yellowfin tuna (Thunnus albacares) population using a time series analysis. Deep Sea Res. Part II Top. Stud. Oceanogr. 175, 104808 (2020).Article 

    Google Scholar 
    Faillettaz, R., Beaugrand, G., Goberville, E. & Kirby, R. R. Atlantic Multidecadal Oscillations drive the basin-scale distribution of Atlantic bluefin tuna. Sci. Adv. 5(1), eaar6993 (2019).ADS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Lan, K. W., Evans, K. & Lee, M. A. Effects of climate variability on the distribution and fishing conditions of yellowfin tuna (Thunnus albacares) in the western Indian Ocean. Clim. Change 119, 63–77 (2013).ADS 
    Article 

    Google Scholar 
    Lan, K. W., Chang, Y. J. & Wu, Y. L. Influence of oceanographic and climatic variability on the catch rate of yellowfin tuna (Thunnus albacares) cohorts in the Indian Ocean. Deep Sea Res. Part II Top. Stud. Oceanogr. 175, 104681 (2019).Article 

    Google Scholar 
    Drinkwater, K. et al. Climate forcing on marine ecosystems. In Marine Ecosystems and Global Change 11–39 (2010).Lan, K. W., Wu, Y. L., Chen, L. C., Naimullah, M. & Lin, T. H. Effects of climate change in marine ecosystems based on the spatiotemporal age structure of top predators: A case study of bigeye tuna in the Pacific Ocean. Front. Mar. Sci. 8, 352 (2021).Article 

    Google Scholar 
    Li, S. et al. The Pacific Decadal Oscillation less predictable under greenhouse warming. Nat. Clim. Chang. 10, 30–34 (2020).ADS 
    CAS 
    Article 

    Google Scholar 
    Debertin, A. J., Irvine, J. R., Holt, C. A., Oka, G. & Trudel, M. Marine growth patterns of southern British Columbia chum salmon explained by interactions between density-dependent competition and changing climate. Can. J. Fish. Aquat. Sci. 74(7), 1077–1087 (2017).Article 

    Google Scholar 
    Di Lorenzo, E. et al. North Pacific Gyre Oscillation links ocean climate and ecosystem change. Geophys. Res. Lett. https://doi.org/10.1029/2007GL032838 (2008).Article 

    Google Scholar 
    Oceanic Fisheries Programme Pacific Community. Western and central Pacific fisheries commission tuna fishery yearbook (2020).IOTC. Report of the Twelfth Session of the Scientific Committee of the Indian Ocean Tuna Commsion. Victoria, Seychelles, 190 (2009).Pecoraro, C. et al. Putting all the pieces together: Integrating current knowledge of the biology, ecology, fisheries status, stock structure and management of yellowfin tuna (Thunnus albacares). Rev. Fish. Biol. Fish. 27(4), 811–841 (2017).Article 

    Google Scholar 
    Lee, Y. C., Nishida, T. & Mohri, M. Separation of the Taiwanese regular and deep tuna longliners in the Indian Ocean using bigeye tuna catch ratios. Fish. Sci. 71(6), 1256–1263 (2005).CAS 
    Article 

    Google Scholar 
    Marsac, F. Outlook of ocean climate variability in the west tropical Indian Ocean, 1997–2008. Working document for IOTC Indian Ocean Tuna Commission (2008).Lehodey, P., Chai, F. & Hampton, J. Modelling climate-related variability of tuna populations from a coupled ocean–biogeochemical-populations dynamics model. Fish Oceanogr. 12(4–5), 483–494 (2003).Article 

    Google Scholar 
    Torres-Faurrieta, L. K., Dreyfus-León, M. J. & Rivas, D. Recruitment forecasting of yellowfin tuna in the eastern Pacific Ocean with artificial neuronal networks. Ecol. Inform. 36, 106–113 (2016).Article 

    Google Scholar 
    Planque, B. et al. How does fishing alter marine populations and ecosystems sensitivity to climate?. J. Mar. Syst. 79(3–4), 403–417 (2010).Article 

    Google Scholar 
    Perry, R. I. et al. Sensitivity of marine systems to climate and fishing: Concepts, issues and management responses. J. Mar. Syst. 79(3–4), 427–435 (2010).Article 

    Google Scholar 
    Sen Gupta, A. & McNeil, B. Variability and change in the ocean. In The Future of the World’s Climate 141–165 (2012).Welch, H., Pressey, R. L. & Reside, A. E. Using temporally explicit habitat suitability models to assess threats to mobile species and evaluate the effectiveness of marine protected areas. J. Nat. Conserv. 41, 106–115 (2018).Article 

    Google Scholar 
    Shin, A., Yoon, S. C., Lee, S. I., Park, H. W. & Kim, S. The relationship between fishing characteristics of Pacific bluefin tuna (Thunnus orientalis) and ocean conditions around Jeju Island. Fish. Quat. Sci. 21, 1–12 (2018).
    Google Scholar 
    Monllor-Hurtado, A., Pennino, M. G. & Sanchez-Lizaso, J. L. Shift in tuna catches due to ocean warming. PLoS ONE 12, e0178196 (2017).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Arrizabalaga, H. et al. Global habitat preferences of commercially valuable tuna. Deep Sea Res. Part II Top. Stud. Oceanogr. 113, 102–112 (2015).ADS 
    Article 

    Google Scholar 
    Yen, K. W. et al. Using remote-sensing data to detect habitat suitability for yellowfin tuna in the Western and Central Pacific Ocean. Int. J. Remote Sens. 33(23), 7507–7522 (2012).Article 

    Google Scholar 
    Liu, Q. et al. Seasonal and intraseasonal thermocline variability in the central South China Sea. Geophys. Res. Lett. 28(23), 4467–4470 (2001).ADS 
    Article 

    Google Scholar 
    Schaefer, K. M., Fuller, D. W. & Block, B. A. Movements, behavior, and habitat utilization of yellowfin tuna (Thunnus albacares) in the northeastern Pacific Ocean, ascertained through archival tag data. Mar. Biol. 152, 503–525 (2007).Article 

    Google Scholar 
    Song, L. M. et al. Environmental preferences of longlining for yellowfin tuna (Thunnus albacares) in the tropical high seas of the Indian Ocean. Fish Oceanogr. 17, 239–253 (2008).Article 

    Google Scholar 
    Bismuto, E. et al. Molecular dynamics simulation of the acidic compact state of apomyoglobin from yellowfin tuna. Proteins 74, 273–290 (2009).CAS 
    PubMed 
    Article 

    Google Scholar 
    Galli, G. L. J., Shiels, H. A. & Brill, R. W. Temperature sensitivity of cardiac function in pelagic fishes with different vertical mobilities: yellowfin tuna (Thunnus albacares), bigeye tuna (Thunnus obesus), mahimahi (Coryphaena hippurus), and swordfish (Xiphias gladius). Physiol. Biochem. Zool. 82, 280–290 (2009).PubMed 
    Article 

    Google Scholar 
    Weng, K. C. et al. Habitat and behaviour of yellowfin tuna Thunnus albacares in the Gulf of Mexico determined using pop-up satellite archival tags. J. Fish Biol. 74, 1434–1449 (2009).CAS 
    PubMed 
    Article 

    Google Scholar 
    Tseng, C. T. et al. Spatio-temporal distributions of tuna species and potential habitats in the Western and Central Pacific Ocean derived from multi-satellite data. Int. J. Remote Sens. 31, 4543–4558 (2010).Article 

    Google Scholar 
    Báez, J. C., Czerwinski, I. A. & Ramos, M. L. Climatic oscillations effect on the yellowfin tuna (Thunnus albacares) Spanish captures in the Indian Ocean. Fish Oceanogr. 29(6), 572–583 (2020).Article 

    Google Scholar 
    Mantua, N. J., Hare, S. R., Zhang, Y., Wallace, J. M. & Francis, R. C. A Pacific interdecadal climate oscillation with impacts on salmon production. Bull. Am. Meteorol. Soc. 78, 1069–1080 (1997).ADS 
    Article 

    Google Scholar 
    Messié, M. & Chavez, F. Global modes of sea surface temperature variability in relation to regional climate indices. J. Clim. 24, 4314–4331 (2011).ADS 
    Article 

    Google Scholar 
    Michael, P. E., Tuck, G. N., Strutton, P. & Hobday, A. Environmental associations with broad-scale Japanese and Taiwanese pelagic longline effort in the southern Indian and Atlantic Oceans. Fish. Oceanogr. 24(5), 478–493 (2015).Article 

    Google Scholar 
    Chavez, F. P., Ryan, J., Lluch-Cota, S. E. & Ñiquen, M. From anchovies to sardines and back: Multidecadal change in the Pacific Ocean. Science 299, 217–221 (2003).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Chiba, S. et al. Temperature and zooplankton size structure: climate control and basin-scale comparison in the North Pacific. Ecol. Evol. 5(4), 968–978 (2015).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Olson, R. J. et al. Decadal diet shift in yellowfin tuna (Thunnus albacares) suggests broad-scale food web changes in the eastern tropical Pacific Ocean. Mar. Ecol.-Prog. Ser. 497, 157–178 (2014).ADS 
    Article 

    Google Scholar 
    Deepa, J. S. et al. The tropical Indian Ocean decadal sea level response to the Pacific decadal oscillation forcing. Clim. Dyn. 52, 5045–5058 (2019).Article 

    Google Scholar 
    Vibhute, A. et al. Decadal variability of tropical Indian Ocean Sea surface temperature and its impact on the Indian summer monsoon. Theor. Appl. Climatol. 141, 551–566 (2020).ADS 
    Article 

    Google Scholar 
    Ummenhofer, C. C., Biastoch, A. & Böning, C. W. Multidecadal Indian Ocean variability linked to the Pacific and implications for preconditioning Indian Ocean dipole events. J. Clim. 30, 1739–1751 (2017).ADS 
    Article 

    Google Scholar 
    Latif, M. The ocean’s role in modeling and predicting decadal climate variations. In International Geophysics 645–665 (Academic Press, 2013).Sun, C. et al. Western tropical Pacific multidecadal variability forced by the Atlantic multidecadal oscillation. Nat. Commun. 8, 1–10 (2017).Article 
    CAS 

    Google Scholar 
    Xie, T., Li, J., Chen, K., Zhang, Y. & Sun, C. Origin of Indian Ocean multidecadal climate variability: Role of the North Atlantic Oscillation. Clim. Dyn. 56, 3277–3294 (2021).Article 

    Google Scholar 
    Myers, R. A. & Worm, B. Rapid worldwide depletion of predatory fish communities. Nature 423, 280–283 (2003).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Ciannelli, L. et al. Climate forcing, food web structure and community dynamics in pelagic marine ecosystems. In Aquatic Food Webs: An Ecosystem Approach 143–169 (Oxford University Press, Oxford, 2005).Enfield, D. B., Mestas-Nuñez, A. M. & Trimble, P. J. The Atlantic multidecadal oscillation and its relation to rainfall and river flows in the continental US. Geophys. Res. Lett. 28, 2077–2080 (2001).ADS 
    Article 

    Google Scholar 
    Zuo, H., Balmaseda, M., Tietsche, S., Mogensen, K. & Mayer, M. The ECMWF operational ensemble reanalysis-analysis system for ocean and sea-ice: A description of the system and assessment. Ocean Sci. 15(3), 779–808 (2019).ADS 
    Article 

    Google Scholar 
    Harley, S. J., Myers, R. A. & Dunn, A. Is catch-per-unit-effort proportional to abundance?. Can J. Fish. Aquat. Sci. 58, 1760–1772 (2001).Article 

    Google Scholar 
    Guyomard, D., Desruisseaux, M., Poisson, F., Taquet, M., Petit, M. GAM analysis of operational and environmental factors affecting swordfish (Xiphias gladius) catch and CPUE of the Reunion Island longline fishery, in the South Western Indian Ocean. IOTC-2004-WPB-08, 38 (2004).Su, N. J., Sun, C. L., Punt, A. E., Yeh, S. Z. & DiNardo, G. Modelling the impacts of environmental variation on the distribution of blue marlin, Makaira nigricans, in the Pacific Ocean. ICES J. Mar. Sci. 68, 1072–1080 (2011).Article 

    Google Scholar 
    Bonett, D. G. & Wright, T. A. Sample size requirements for estimating Pearson, Kendall and Spearman correlations. Psychometrika 65(1), 23–28 (2000).MATH 
    Article 

    Google Scholar 
    Weaver, B. & Koopman, R. An SPSS macro to compute confidence intervals for Pearson’s correlation. Quant. Methods Psychol. 10(1), 29–39 (2014).Article 

    Google Scholar 
    Naimullah, M. et al. Effect of the El Niño-Southern Oscillation (ENSO) cycle on the catches and habitat patterns of three swimming crabs in the Taiwan Strait. Front. Mar. Sci. https://doi.org/10.3389/fmars.2021.763543 (2021).Article 

    Google Scholar 
    Chen, X. J., Li, G., Feng, B. & Tian, S. Q. Habitat suitability index of Chub mackerel (Scomber japonicus) from July to September in the East China Sea. J. Oceanogr. 65, 93–102 (2009).Article 

    Google Scholar 
    Urich, D. L. & Graham, J. P. Applying habitat evaluation procedures (HEP) to wildlife area planning in Missouri. Wildl. Soc. Bull. 11(3), 215–222 (1983).
    Google Scholar 
    Chen, X. J., Tian, S. Q., Chen, Y. & Liu, B. L. A modeling approach to identify optimal habitat and suitable fishing grounds for neon flying squid (Ommastrephes bartramii) in the Northwest Pacific Ocean. Fish. Bull. 108, 1–14 (2010).
    Google Scholar 
    Tian, S. Q., Chen, X. J., Chen, Y., Xu, L. X. & Dai, X. J. Evaluating habitat suitability indices derived from CPUE and fishing effort data for Ommatrephes bratramii in the northwestern Pacific Ocean. Fish Res. 95, 181–188 (2009).Article 

    Google Scholar 
    Rouyer, T., Sadykov, A., Ohlberger, J. & Stenseth, N. C. Does increasing mortality change the response of fish populations to environmental fluctuations?. Ecol. Lett. 15, 658–665 (2012).PubMed 
    Article 

    Google Scholar 
    Grinsted, A., Moore, J. C. & Jevrejeva, S. Application of the cross wavelet transform and wavelet coherence to geophysical time series. Nonlinear Process Geophys. 11, 561–566 (2004).ADS 
    Article 

    Google Scholar 
    Torrence, C. & Compo, G. P. A practical guide to wavelet analysis. Bull. Amer. Meteorol. Soc. 79, 61–78 (1998).ADS 
    Article 

    Google Scholar  More

  • in

    A life history model of the ecological and evolutionary dynamics of polyaneuploid cancer cells

    Housman, G. et al. Drug resistance in cancer: An overview. Cancers 6(3), 1769 (2014).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Vasan, N. Baselga, J. & Hyman, D. M. A View on Drug Resistance in Cancer, 11 (2019).Casás-Selves, M. & Degregori, J. How cancer shapes evolution and how evolution shapes cancer (2011).Dujon, A. M. et al. Identifying key questions in the ecology and evolution of cancer. Evol. Appl. 14, 4 (2021).
    Google Scholar 
    Korolev, K. S., Xavier, J. B. & Gore, J. Turning ecology and evolution against cancer (2014).Merlo, L. M. F., Pepper, J. W., Reid, B. J. & Maley, C. C. Cancer as an evolutionary and ecological process. Nat. Rev. Cancer 6(12), 924–935 (2006).CAS 
    PubMed 
    Article 

    Google Scholar 
    Ujvari, B., Roche, B. & Thomas, F. Ecology and Evolution of Cancer 1st edn. (Academic Press, 2017).
    Google Scholar 
    Brown, R. L. What evolvability really is. Brit. J. Philos. Sci. 65, 3 (2014).MathSciNet 
    Article 

    Google Scholar 
    Crother, B. I. & Murray, C. M. Early usage and meaning of evolvability. Ecol. Evol. 9, 7 (2019).Article 

    Google Scholar 
    Pigliucci, M. Is evolvability evolvable? (2008).Sniegowski, P. D. & Murphy, H. A. Evolvability (2006).Bukkuri, A. & Brown, J. S. Evolutionary game theory: Darwinian dynamics and the G function approach. MDPI Games 12(4), 1–19 (2021).MathSciNet 
    MATH 

    Google Scholar 
    Fisher, R. A. The Genetical Theory of Natural Selection (The Clarendon Press, 1930).MATH 
    Book 

    Google Scholar 
    Li, C. C. Fundamental theorem of natural selection. Nature 214(5087), 4 (1967).Article 

    Google Scholar 
    Vincent, T. L. & Brown, J. S. Evolutionary Game Theory, Natural Selection, and Darwinian Dynamics (Cambridge University Press, 2005).MATH 
    Book 

    Google Scholar 
    Hanahan, D. & Weinberg, R. A. The next generation. Leading edge review hallmarks of cancer. Cell 144, 646–674 (2011).CAS 
    PubMed 
    Article 

    Google Scholar 
    Pienta, K. J. et al. Cancer cells employ an evolutionarily conserved polyploidization program to resist therapy. Semin. Cancer Biol. 20, 1–15 (2020).
    Google Scholar 
    Virchow, R. As based upon physiological and pathological histology: Cellular pathology. Nutr. Rev. 47(1), 23–25 (1989).CAS 
    PubMed 
    Article 

    Google Scholar 
    Razmik, M., Bonnie, A. & David, M. Roles of polyploid/multinucleated giant cancer cells in metastasis and disease relapse following anticancer treatment. Cancers 10(4), 4 (2018).
    Google Scholar 
    Amend, S. R. et al. Polyploid giant cancer cells: Unrecognized actuators of tumorigenesis, metastasis, and resistance. Prostate 79(13), 1489–1497 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Kuczler, M. D., Olseen, A. M., Pienta, K. J. & Amend, S. R. ROS-induced cell cycle arrest as a mechanism of resistance in polyaneuploid cancer cells (PACCs). Prog. Biophys. Mol. Biol. 20, 3–7 (2021).Article 
    CAS 

    Google Scholar 
    Kostecka, L. G., Pienta, K. J. & Amend, S. R. Polyaneuploid cancer cell dormancy: Lessons from evolutionary phyla. Front. Ecol. Evol. 9, 439 (2021).Article 

    Google Scholar 
    Rajaraman, R., Rajaraman, M. M., Rajaraman, S. R. & Guernsey, D. L. Neosis—-a paradigm of self-renewal in cancer. Cell Biol. Int. 29(12), 1084–1097 (2005).CAS 
    PubMed 
    Article 

    Google Scholar 
    Rajaraman, R., Guernsey, D. L., Rajaraman, M. M. & Rajaraman, S. R. Neosis—a parasexual somatic reduction division in cancer. Int. J. Hum. Genet. 7(1), 29–48 (2007).CAS 
    Article 

    Google Scholar 
    Sundaram, M., Guernsey, D. L., Rajaraman, M. M. & Rajaraman, R. Neosis: A novel type of cell division in cancer. Cancer Biol. Ther. 3(2), 207–218 (2004).CAS 
    PubMed 
    Article 

    Google Scholar 
    Illidge, T. M., Cragg, M. S., Fringes, B., Olive, P. & Erenpreisa, J. A. Polyploid giant cells provide a survival mechanism for p53 mutant cells after DNA damage. Cell Biol. Int. 24(9), 621–633 (2000).CAS 
    PubMed 
    Article 

    Google Scholar 
    Puig, P. E. et al. Tumor cells can escape DNA-damaging cisplatin through DNA endoreduplication and reversible polyploidy. Cell Biol. Int. 32(9), 1031–1043 (2008).CAS 
    PubMed 
    Article 

    Google Scholar 
    Zhang, S. et al. Generation of cancer stem-like cells through the formation of polyploid giant cancer cells. Oncogene 33(1), 116–128 (2014).CAS 
    PubMed 
    Article 

    Google Scholar 
    Comai, L. The advantages and disadvantages of being polyploid. Nat. Rev. Genet. 6(11), 836–846 (2005).CAS 
    PubMed 
    Article 

    Google Scholar 
    Hassel, C., Zhang, B., Dixon, M. & Calvi, B. R. Induction of endocycles represses apoptosis independently of differentiation and predisposes cells to genome instability. Development (Cambridge) 141(1), 112–123 (2014).CAS 
    Article 

    Google Scholar 
    Lee, H. O., Davidson, J. M. & Duronio, R. J. Endoreplication: Polyploidy with purpose. Genes Dev. 23(21), 2461–2477 (2009).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Basener, W. F. & Sanford, J. C. The fundamental theorem of natural selection with mutations. J. Math. Biol. 76(7), 1589–1622 (2018).MathSciNet 
    PubMed 
    MATH 
    Article 

    Google Scholar 
    Frank, S. A. & Slatkin, M. Fisher’s fundamental theorem of natural selection. Trends Ecol. Evol. 7(3), 92–95 (1992).CAS 
    PubMed 
    Article 

    Google Scholar 
    Lessard, S. Fisher’s fundamental theorem of natural selection revisited. Theor. Popul. Biol. 52(2), 119–136 (1997).MathSciNet 
    CAS 
    PubMed 
    MATH 
    Article 

    Google Scholar 
    Das, P., Mukherjee, S. & Das, P. An investigation on Michaelis–Menten kinetics based complex dynamics of tumor-immune interaction. Chaos Solitons Fractals 1, 28 (2019).MathSciNet 
    CAS 
    MATH 

    Google Scholar 
    Renee Fister, K. & Panetta, J. C. Optimal control applied to competing chemotherapeutic cell-kill strategies. SIAM J. Appl. Math. 63, 6 (2003).MathSciNet 
    MATH 

    Google Scholar 
    López, Á. G., Seoane, J. M. & Sanjuán, M. A. F. Decay dynamics of tumors. PLoS One 11, 6 (2016).
    Google Scholar 
    Pienta, K. J., Hammarlund, E. U., Brown, J. S., Amend, S. R. & Axelrod, R. M. Cancer recurrence and lethality are enabled by enhanced survival and reversible cell cycle arrest of polyaneuploid cells. Proc. Natl. Acad. Sci. U.S.A. 118(7), 2 (2021).Article 
    CAS 

    Google Scholar 
    Pienta, K. J., Hammarlund, E. U., Axelrod, R., Brown, J. S. & Amend, S. R. Poly-aneuploid cancer cells promote evolvability, generating lethal cancer. Evol. Appl. 13(7), 1626–1634 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Mittal, K. et al. Multinucleated polyploidy drives resistance to Docetaxel chemotherapy in prostate cancer. Br. J. Cancer 116, 9 (2017).Article 
    CAS 

    Google Scholar 
    Cunningham, J. J., Bukkuri, A., Gatenby, R., Brown, J. S. & Gillies, R. J. Coupled source-sink habitats produce spatial and temporal variation of cancer cell molecular properties as an alternative to branched clonal evolution and stem cell paradigms. Front. Ecol. Evol. 9, 472 (2021).Article 

    Google Scholar 
    Fujiwara, M. & Diaz-Lopez, J. Constructing stage-structured matrix population models from life tables: Comparison of methods. PeerJ 5(10), 1–27 (2017).
    Google Scholar 
    Kendall, B. E. et al. Persistent problems in the construction of matrix population models. Ecol. Model. 406, 33–43 (2019).Article 

    Google Scholar 
    Law, R. & Edley, M. T. Transient dynamics of populations with age- and size-dependent vital rates. Ecology 71(5), 1863–1870 (1990).Article 

    Google Scholar 
    Velde, R. V. et al. Resistance to targeted therapies as a multifactorial, gradual adaptation to inhibitor specific selective pressures. Nat. Commun. 11(1), 1–13 (2020).Article 
    CAS 

    Google Scholar 
    Salmina, K. et al. The cancer aneuploidy paradox: In the light of evolution. Genes 10(2), 83 (2019).CAS 
    PubMed Central 
    Article 

    Google Scholar 
    Turajlic, S., Sottoriva, A., Graham, T. & Swanton, C. Resolving genetic heterogeneity in cancer. Nat. Rev. Genet. 20(7), 404–416 (2019).CAS 
    PubMed 
    Article 

    Google Scholar 
    Miller, A. K., Brown, J. S., Enderling, H., Basanta, D. & Whelan, C. J. The evolutionary ecology of dormancy in nature and in cancer. Front. Ecol. Evol. 9, 5 (2021).Article 

    Google Scholar 
    Geiser, F. Hibernation. Curr. Biol. 23(5), R188–R193 (2013).CAS 
    PubMed 
    Article 

    Google Scholar 
    Lyman, C. P. & Chatfield, P. O. Physiology of hibernation in mammals. Physiol. Rev. 35(2), 403–425 (1955).CAS 
    PubMed 
    Article 

    Google Scholar 
    Lin, K. C. et al. The role of heterogeneous environment and docetaxel gradient in the emergence of polyploid, mesenchymal and resistant prostate cancer cells. Clin. Exp. Metas. 36(2), 97–108 (2019).Article 

    Google Scholar 
    Lin, K. C. et al. An: In vitro tumor swamp model of heterogeneous cellular and chemotherapeutic landscapes. Lab Chip 20(14), 2453–2464 (2020).CAS 
    PubMed 
    Article 

    Google Scholar 
    Kawamura, E. et al. Identification of novel small molecule inhibitors of centrosome clustering in cancer cells. Oncotarget 4(10), 1763–1776 (2013).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Kostecka, L. G. et al. High KIFC1 expression is associated with poor prognosis in prostate cancer. Med. Oncol. 38, 1–9 (2021).Article 
    CAS 

    Google Scholar 
    Sekino, Y. et al. KIFC1 induces resistance to docetaxel and is associated with survival of patients with prostate cancer. Urol. Oncol. Semin. Original Investig. 35(1), 1–8 (2017).Article 

    Google Scholar 
    Xiao, Y. X. & Yang, W. X. KIFC1: A promising chemotherapy target for cancer treatment?. Oncotarget 7(30), 1–9 (2016).
    Google Scholar 
    Law, M. E., Corsino, P. E., Narayan, S. & Law, B. K. Cyclin-dependent kinase inhibitors as anticancer therapeutics. Mol. Pharmacol. 88, 5 (2015).Article 
    CAS 

    Google Scholar 
    Tadesse, S., Caldon, E. C., Tilley, W. & Wang, S. Cyclin-Dependent Kinase 2 Inhibitors in Cancer Therapy: An Update (2019).Zhang, M. et al. CDK inhibitors in cancer therapy, an overview of recent development. Am. J. Cancer Res. 11, 5 (2021).CAS 

    Google Scholar 
    Kostecka, L. G., Pienta, K. J. & Amend, S. R. Lipid droplet evolution gives insight into polyaneuploid cancer cell lipid droplet functions. Med. Oncol. 38(11), 1–10 (2021).Article 
    CAS 

    Google Scholar 
    Strobl, M. A. R. et al. Turnover modulates the need for a cost of resistance in adaptive therapy. Can. Res. 81, 4 (2021).Article 

    Google Scholar 
    West, J., Ma, Y. & Newton, P. K. Capitalizing on competition: An evolutionary model of competitive release in metastatic castration resistant prostate cancer treatment. J. Theor. Biol. 4, 55 (2018).MathSciNet 
    MATH 

    Google Scholar  More