More stories

  • in

    Trait biases in microbial reference genomes

    Overmann, J., Abt, B. & Sikorski, J. Present and future of culturing bacteria. Annual Review of Microbiology 71, 711–730 (2017).CAS 

    Google Scholar 
    O’Leary, N. A. et al. Reference sequence (RefSeq) database at NCBI: current status, taxonomic expansion, and functional annotation. Nucleic Acids Research 44, D733 (2016).
    Google Scholar 
    Bobay, L. M. & Ochman, H. Biological species are universal across life’s domains. Genome Biology and Evolution 9, 491–501 (2017).
    Google Scholar 
    Magnabosco, C., Moore, K., Wolfe, J. & Fournier, G. Dating phototrophic microbial lineages with reticulate gene histories. Geobiology 16, 179–189 (2018).CAS 

    Google Scholar 
    Louca, S. et al. Function and functional redundancy in microbial systems. Nature Ecology & Evolution 2, 936–943 (2018).ADS 

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

    Google Scholar 
    Zhu, Q. et al. Phylogenomics of 10,575 genomes reveals evolutionary proximity between domains bacteria and archaea. Nature Communications 10, 5477 (2019).ADS 
    CAS 

    Google Scholar 
    Royalty, T.M. & Steen, A.D. Quantitatively partitioning microbial genomic traits among taxonomic ranks across the microbial tree of life. mSphere 4 (2019).Murray, C. S., Gao, Y. & Wu, M. Re-evaluating the evidence for a universal genetic boundary among microbial species. Nature Communications 12, 4059 (2021).ADS 
    CAS 

    Google Scholar 
    Powell, S. et al. eggNOG v4.0: nested orthology inference across 3686 organisms. Nucleic Acids Research 42, D231–D239 (2014).CAS 

    Google Scholar 
    Stoddard, S. F., Smith, B. J., Hein, R., Roller, B. R. & Schmidt, T. M. rrnDB: improved tools for interpreting rRNA gene abundance in bacteria and archaea and a new foundation for future development. Nucleic Acids Research 43, D593–D598 (2014).
    Google Scholar 
    Douglas, G. M. et al. Picrust2 for prediction of metagenome functions. Nature Biotechnology 38, 685–688 (2020).CAS 

    Google Scholar 
    Wemheuer, F. et al. Tax4Fun2: prediction of habitat-specific functional profiles and functional redundancy based on 16S rRNA gene sequences. Environmental Microbiome 15, 1–12 (2020).
    Google Scholar 
    Louca, S., Parfrey, L. W. & Doebeli, M. Decoupling function and taxonomy in the global ocean microbiome. Science 353, 1272–1277 (2016).ADS 
    CAS 

    Google Scholar 
    Wu, D. et al. A phylogeny-driven genomic encyclopaedia of bacteria and archaea. Nature 462, 1056–1060 (2009).ADS 
    CAS 

    Google Scholar 
    Louca, S. & Pennell, M. W. A general and efficient algorithm for the likelihood of diversification and discrete-trait evolutionary models. Systematic Biology 69, 545–556 (2020).
    Google Scholar 
    Tyson, G. W. et al. Community structure and metabolism through reconstruction of microbial genomes from the environment. Nature 428, 37–43 (2004).ADS 
    CAS 

    Google Scholar 
    Sharon, I. & Banfield, J. F. Genomes from metagenomics. Science 342, 1057–1058 (2013).ADS 
    CAS 

    Google Scholar 
    Parks, D. H. et al. Recovery of nearly 8,000 metagenome-assembled genomes substantially expands the tree of life. Nature Microbiology 2, 1533–1542 (2017).CAS 

    Google Scholar 
    Chen, L. X., Anantharaman, K., Shaiber, A., Eren, A. M. & Banfield, J. F. Accurate and complete genomes from metagenomes. Genome Research 30, 315–333 (2020).CAS 

    Google Scholar 
    Konstantinidis, K. T. & Tiedje, J. M. Genomic insights that advance the species definition for prokaryotes. Proceedings of the National Academy of Sciences 102, 2567–2572 (2005).ADS 
    CAS 

    Google Scholar 
    Kim, M., Oh, H. S., Park, S. C. & Chun, J. Towards a taxonomic coherence between average nucleotide identity and 16S rRNA gene sequence similarity for species demarcation of prokaryotes. Journal of Systematic and Evolutionary Microbiology 64, 346–351 (2014).CAS 

    Google Scholar 
    Shapiro, B.J. What microbial population genomics has taught us about speciation. In Polz, M.F. & Rajora, O.P. (eds.) Population Genomics: Microorganisms, 31–47 (Springer International Publishing, Cham, Switzerland, 2019).Olm, M. R. et al. Consistent metagenome-derived metrics verify and delineate bacterial species boundaries. mSystems 5, e00731–19 (2020).CAS 

    Google Scholar 
    Lagkouvardos, I., Overmann, J. & Clavel, T. Cultured microbes represent a substantial fraction of the human and mouse gut microbiota. Gut Microbes 8, 493–503 (2017).
    Google Scholar 
    Zhang, Z., Wang, J., Wang, J., Wang, J. & Li, Y. Estimate of the sequenced proportion of the global prokaryotic genome. Microbiome 8, 1–9 (2020).
    Google Scholar 
    Aramaki, T. et al. KofamKOALA: KEGG Ortholog assignment based on profile HMM and adaptive score threshold. Bioinformatics 36, 2251–2252 (2019).
    Google Scholar 
    Mira, A., Ochman, H. & Moran, N. A. Deletional bias and the evolution of bacterial genomes. Trends in Genetics 17, 589–596 (2001).CAS 

    Google Scholar 
    Morris, J. J., Lenski, R. E. & Zinser, E. R. The Black Queen Hypothesis: evolution of dependencies through adaptive gene loss. MBio 3, e00036–12 (2012).
    Google Scholar 
    Giovannoni, S. J., Cameron Thrash, J. & Temperton, B. Implications of streamlining theory for microbial ecology. ISME Journal 8, 1553–1565 (2014).
    Google Scholar 
    Nayfach, S., Shi, Z. J., Seshadri, R., Pollard, K. S. & Kyrpides, N. C. New insights from uncultivated genomes of the global human gut microbiome. Nature 568, 505–510 (2019).ADS 
    CAS 

    Google Scholar 
    Gary, P.R. Adjusting for nonresponse in surveys. In Smart, J.C. (ed.) Higher Education: Handbook of Theory and Research, chap. 8, 411–449 (Springer, Dordrecht, Netherlands, 2007).Maguire, F. et al. Metagenome-assembled genome binning methods with short reads disproportionately fail for plasmids and genomic islands. Microbial Genomics 6, mgen000436 (2020).
    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 Research 47, D309–D314 (2019).CAS 

    Google Scholar 
    Abdel-Hamid, A.M., Solbiati, J.O., Cann, I.K.O., Sariaslani, S. & Gadd, G.M. Insights into lignin degradation and its potential industrial applications, vol. 82, chap. 1, 1–28 (Academic Press, 2013).El-Bondkly, A.M. Sequence analysis of industrially important genes from trichoderma. In Biotechnology and biology of Trichoderma, chap. 28, 377–392 (Elsevier, 2014).Dawood, A. & Ma, K. Applications of microbial β-mannanases. Frontiers in Bioengineering and Biotechnology 8 (2020).Khelaifia, S., Raoult, D. & Drancourt, M. A versatile medium for cultivating methanogenic archaea. PLOS ONE 8, e61563 (2013).ADS 
    CAS 

    Google Scholar 
    Khelaifia, S. et al. Aerobic culture of methanogenic archaea without an external source of hydrogen. European Journal of Clinical Microbiology & Infectious Diseases 35, 985–991 (2016).CAS 

    Google Scholar 
    Michał, B. et al. Phymet2: a database and toolkit for phylogenetic and metabolic analyses of methanogens. Environmental Microbiology Reports 10, 378–382 (2018).
    Google Scholar 
    Albright, S. & Louca, S. Trait biases in microbial reference genomes, figshare., https://doi.org/10.6084/m9.figshare.c.6055004.v1 (2022).Castelle, C. J. & Banfield, J. F. Major new microbial groups expand diversity and alter our understanding of the tree of life. Cell 172, 1181–1197 (2018).CAS 

    Google Scholar 
    Murray, A. E. et al. Roadmap for naming uncultivated archaea and bacteria. Nature Microbiology 5, 987–994 (2020).CAS 

    Google Scholar 
    Palleroni, N. J. Prokaryotic diversity and the importance of culturing. Antonie van Leeuwenhoek 72, 3–19 (1997).CAS 

    Google Scholar 
    Langille, M. G. et al. Predictive functional profiling of microbial communities using 16S rRNA marker gene sequences. Nature Biotechnology 31, 814–821 (2013).CAS 

    Google Scholar 
    Tran, P. Q. et al. Depth-discrete metagenomics reveals the roles of microbes in biogeochemical cycling in the tropical freshwater Lake Tanganyika. The ISME Journal 15, 1971–1986 (2021).CAS 

    Google Scholar 
    Kroeger, M. E. et al. New biological insights into how deforestation in amazonia affects soil microbial communities using metagenomics and metagenome-assembled genomes. Frontiers in Microbiology 9, 1635 (2018).
    Google Scholar 
    Nathani, N. M. et al. 309 metagenome assembled microbial genomes from deep sediment samples in the Gulfs of Kathiawar Peninsula. Scientific Data 8, 194 (2021).
    Google Scholar 
    Irazoqui, J. M., Eberhardt, M. F., Adjad, M. M., Amadio, A. F. & Collado, M. C. Identification of key microorganisms in facultative stabilization ponds from dairy industries, using metagenomics. PeerJ 10, e12772 (2022).
    Google Scholar 
    Hwang, Y. et al. Leave no stone unturned: individually adapted xerotolerant Thaumarchaeota sheltered below the boulders of the Atacama Desert hyperarid core. Microbiome 9, 234 (2021).CAS 

    Google Scholar 
    Tully, B., Wheat, C. G., Glazer, B. T. & Huber, J. A dynamic microbial community with high functional redundancy inhabits the cold, oxic subseafloor aquifer. ISME Journal 12, 1–16 (2018).CAS 

    Google Scholar 
    Vanwonterghem, I., Jensen, P. D., Rabaey, K. & Tyson, G. W. Genome-centric resolution of microbial diversity, metabolism and interactions in anaerobic digestion. Environmental Microbiology 18, 3144–3158 (2016).CAS 

    Google Scholar 
    Glasl, B. et al. Comparative genome-centric analysis reveals seasonal variation in the function of coral reef microbiomes. The ISME Journal 14, 1435–1450 (2020).
    Google Scholar 
    Robbins, S. J. et al. A genomic view of the reef-building coral Porites lutea and its microbial symbionts. Nature Microbiology 4, 2090–2100 (2019).
    Google Scholar 
    Engelberts, J. P. et al. Characterization of a sponge microbiome using an integrative genome-centric approach. The ISME Journal 14, 1100–1110 (2020).CAS 

    Google Scholar 
    Bowerman, K. L. et al. Disease-associated gut microbiome and metabolome changes in patients with chronic obstructive pulmonary disease. Nature Communications 11, 5886 (2020).ADS 
    CAS 

    Google Scholar 
    Chen, Y. J. et al. Hydrodynamic disturbance controls microbial community assembly and biogeochemical processes in coastal sediments. The ISME Journal 16, 750–763 (2022).CAS 

    Google Scholar 
    Hugerth, L. W. et al. Metagenome-assembled genomes uncover a global brackish microbiome. Genome Biology 16, 279 (2015).
    Google Scholar 
    Alneberg, J. et al. Ecosystem-wide metagenomic binning enables prediction of ecological niches from genomes. Communications Biology 3, 119 (2020).
    Google Scholar 
    Di Cesare, A. et al. Genomic comparison and spatial distribution of different Synechococcus phylotypes in the Black Sea. Frontiers in Microbiology 11, 1979 (2020).
    Google Scholar 
    van Vliet, D. M. et al. The bacterial sulfur cycle in expanding dysoxic and euxinic marine waters. Environmental Microbiology 23, 2834–2857 (2021).
    Google Scholar 
    Dalcin Martins, P. et al. Enrichment of novel Verrucomicrobia, Bacteroidetes, and Krumholzibacteria in an oxygen-limited methane- and iron-fed bioreactor inoculated with Bothnian Sea sediments. MicrobiologyOpen 10, e1175 (2021).CAS 

    Google Scholar 
    Stewart, R. D. et al. Compendium of 4,941 rumen metagenome-assembled genomes for rumen microbiome biology and enzyme discovery. Nature Biotechnology 37, 953–961 (2019).CAS 

    Google Scholar 
    Segura-Wang, M., Grabner, N., Koestelbauer, A., Klose, V. & Ghanbari, M. Genome-resolved metagenomics of the chicken gut microbiome. Frontiers in Microbiology 12, 726923 (2021).
    Google Scholar 
    Ruuskanen, M. O. et al. Microbial genomes retrieved from High Arctic lake sediments encode for adaptation to cold and oligotrophic environments. Limnology and Oceanography 65, S233–S247 (2020).CAS 

    Google Scholar 
    Haas, S., Desai, D. K., LaRoche, J., Pawlowicz, R. & Wallace, D. W. R. Geomicrobiology of the carbon, nitrogen and sulphur cycles in Powell Lake: a permanently stratified water column containing ancient seawater. Environmental Microbiology 21, 3927–3952 (2019).CAS 

    Google Scholar 
    Spasov, E. et al. High functional diversity among Nitrospira populations that dominate rotating biological contactor microbial communities in a municipal wastewater treatment plant. The ISME Journal 14, 1857–1872 (2020).CAS 

    Google Scholar 
    Vigneron, A. et al. Genomic evidence for sulfur intermediates as new biogeochemical hubs in a model aquatic microbial ecosystem. Microbiome 9, 46 (2021).CAS 

    Google Scholar 
    Galambos, D., Anderson, R. E., Reveillaud, J. & Huber, J. A. Genome-resolved metagenomics and metatranscriptomics reveal niche differentiation in functionally redundant microbial communities at deep-sea hydrothermal vents. Environmental Microbiology 21, 4395–4410 (2019).CAS 

    Google Scholar 
    Stewart, R. D. et al. Assembly of 913 microbial genomes from metagenomic sequencing of the cow rumen. Nature Communications 9, 870 (2018).ADS 

    Google Scholar 
    Xing, P. et al. Stratification of microbiomes during the holomictic period of Lake Fuxian, an alpine monomictic lake. Limnology and Oceanography 65, S134–S148 (2020).
    Google Scholar 
    Zhang, S., Hu, Z. & Wang, H. Metagenomic analysis exhibited the co-metabolism of polycyclic aromatic hydrocarbons by bacterial community from estuarine sediment. Environment International 129, 308–319 (2019).CAS 

    Google Scholar 
    Lin, Y., Wang, L., Xu, K., Li, K. & Ren, H. Revealing taxon-specific heavy metal-resistance mechanisms in denitrifying phosphorus removal sludge using genome-centric metaproteomics. Microbiome 9, 67 (2021).CAS 

    Google Scholar 
    Liu, L. et al. High-quality bacterial genomes of a partial-nitritation/anammox system by an iterative hybrid assembly method. Microbiome 8, 155 (2020).CAS 

    Google Scholar 
    Kantor, R. S. et al. Bioreactor microbial ecosystems for thiocyanate and cyanide degradation unravelled with genome-resolved metagenomics. Environmental Microbiology 17, 4929–4941 (2015).CAS 

    Google Scholar 
    Zhou, Z. et al. Gammaproteobacteria mediating utilization of methyl-, sulfur- and petroleum organic compounds in deep ocean hydrothermal plumes. The ISME Journal 14, 3136–3148 (2020).CAS 

    Google Scholar 
    Reysenbach, A. L. et al. Complex subsurface hydrothermal fluid mixing at a submarine arc volcano supports distinct and highly diverse microbial communities. Proceedings of the National Academy of Sciences 117, 32627–32638 (2020).ADS 
    CAS 

    Google Scholar 
    Hou, J. et al. Microbial succession during the transition from active to inactive stages of deep-sea hydrothermal vent sulfide chimneys. Microbiome 8, 102 (2020).CAS 

    Google Scholar 
    Campanaro, S. et al. Metagenomic analysis and functional characterization of the biogas microbiome using high throughput shotgun sequencing and a novel binning strategy. Biotechnology for Biofuels 9, 26 (2016).
    Google Scholar 
    Singleton, C. M. et al. Connecting structure to function with the recovery of over 1000 high-quality metagenome-assembled genomes from activated sludge using long-read sequencing. Nature Communications 12, 2009 (2021).CAS 

    Google Scholar 
    Diamond, S. et al. Mediterranean grassland soil C–N compound turnover is dependent on rainfall and depth, and is mediated by genomically divergent microorganisms. Nature Microbiology 4, 1356–1367 (2019).CAS 

    Google Scholar 
    Rasigraf, O. et al. Microbial community composition and functional potential in Bothnian Sea sediments is linked to Fe and S dynamics and the quality of organic matter. Limnology and Oceanography 65, S113–S133 (2020).CAS 

    Google Scholar 
    Rissanen, A. J. et al. Vertical stratification patterns of methanotrophs and their genetic controllers in water columns of oxygen-stratified boreal lakes. FEMS Microbiology Ecology 97, fiaa252 (2021).CAS 

    Google Scholar 
    Campanaro, S. et al. New insights from the biogas microbiome by comprehensive genome-resolved metagenomics of nearly 1600 species originating from multiple anaerobic digesters. Biotechnology for Biofuels 13, 25 (2020).CAS 

    Google Scholar 
    Almeida, A. et al. A unified catalog of 204,938 reference genomes from the human gut microbiome. Nature Biotechnology 39, 105–114 (2021).CAS 

    Google Scholar 
    Zhou, Z. et al. Genome- and community-level interaction insights into carbon utilization and element cycling functions of hydrothermarchaeota in hydrothermal sediment. mSystems 5 (2020).Pachiadaki, M. G. et al. Charting the complexity of the marine microbiome through single-cell genomics. Cell 179, 1623–1635.e11 (2019).CAS 

    Google Scholar 
    Martijn, J., Vosseberg, J., Guy, L., Offre, P. & Ettema, T. J. G. Deep mitochondrial origin outside the sampled alphaproteobacteria. Nature 557, 101–105 (2018).ADS 
    CAS 

    Google Scholar 
    Greenlon, A. et al. Global-level population genomics reveals differential effects of geography and phylogeny on horizontal gene transfer in soil bacteria. Proceedings of the National Academy of Sciences 116, 15200–15209 (2019).ADS 
    CAS 

    Google Scholar 
    Hervé, V. et al. Phylogenomic analysis of 589 metagenome-assembled genomes encompassing all major prokaryotic lineages from the gut of higher termites. PeerJ 8, e8614 (2020).
    Google Scholar 
    von Appen, W.J. The expedition PS114 of the research vessel POLARSTERN to the Fram Strait in 2018. Tech. Rep., Alfred Wegener Institute for Polar and Marine Research (2018).Dombrowski, N., Seitz, K. W., Teske, A. P. & Baker, B. J. Genomic insights into potential interdependencies in microbial hydrocarbon and nutrient cycling in hydrothermal sediments. Microbiome 5, 106 (2017).
    Google Scholar 
    Yu, J. et al. Dna-stable isotope probing shotgun metagenomics reveals the resilience of active microbial communities to biochar amendment in oxisol soil. Frontiers in Microbiology 11, 587972 (2020).
    Google Scholar 
    Forster, S. C. et al. A human gut bacterial genome and culture collection for improved metagenomic analyses. Nature Biotechnology 37, 186–192 (2019).CAS 

    Google Scholar 
    Gharechahi, J. et al. Metagenomic analysis reveals a dynamic microbiome with diversified adaptive functions to utilize high lignocellulosic forages in the cattle rumen. The ISME Journal 15, 1108–1120 (2021).CAS 

    Google Scholar 
    Meier, D. V., Imminger, S., Gillor, O. & Woebken, D. Distribution of mixotrophy and desiccation survival mechanisms across microbial genomes in an arid biological soil crust community. mSystems 6, e00786–20 (2021).CAS 

    Google Scholar 
    Haro-Moreno, J. M. et al. Dysbiosis in marine aquaculture revealed through microbiome analysis: reverse ecology for environmental sustainability. FEMS Microbiology Ecology 96, fiaa218 (2020).CAS 

    Google Scholar 
    Haro-Moreno, J. M. et al. Fine metagenomic profile of the Mediterranean stratified and mixed water columns revealed by assembly and recruitment. Microbiome 6, 128 (2018).
    Google Scholar 
    Dong, X. et al. Metabolic potential of uncultured bacteria and archaea associated with petroleum seepage in deep-sea sediments. Nature Communications 10, 1816 (2019).ADS 

    Google Scholar 
    Poghosyan, L. et al. Metagenomic profiling of ammonia- and methane-oxidizing microorganisms in two sequential rapid sand filters. Water Research 185, 116288 (2020).CAS 

    Google Scholar 
    Paula, D. M., Jeroen, F., Hugh, M. & Meng, M. L. & J., W.M. Wetland sediments host diverse microbial taxa capable of cycling alcohols. Applied and Environmental Microbiology 85, 00189–19 (2019).
    Google Scholar 
    Aromokeye, D. A. et al. Crystalline iron oxides stimulate methanogenic benzoate degradation in marine sediment-derived enrichment cultures. The ISME Journal 15, 965–980 (2021).CAS 

    Google Scholar 
    Borchert, E. et al. Deciphering a marine bone-degrading microbiome reveals a complex community effort. mSystems 6, e01218–20 (2021).CAS 

    Google Scholar 
    Osvatic, J. T. et al. Global biogeography of chemosynthetic symbionts reveals both localized and globally distributed symbiont groups. Proceedings of the National Academy of Sciences 118, e2104378118 (2021).CAS 

    Google Scholar 
    Boeuf, D. et al. Biological composition and microbial dynamics of sinking particulate organic matter at abyssal depths in the oligotrophic open ocean. Proceedings of the National Academy of Sciences 116, 11824–11832 (2019).ADS 
    CAS 

    Google Scholar 
    Woodcroft, B. J. et al. Genome-centric view of carbon processing in thawing permafrost. Nature 560, 49–54 (2018).ADS 
    CAS 

    Google Scholar 
    Alqahtani, M. F. et al. Enrichment of Marinobacter sp. and halophilic homoacetogens at the biocathode of microbial electrosynthesis system inoculated with Red Sea brine pool. Frontiers in Microbiology 10, 2563 (2019).
    Google Scholar 
    Haroon, M. F., Thompson, L. R., Parks, D. H., Hugenholtz, P. & Stingl, U. A catalogue of 136 microbial draft genomes from Red Sea metagenomes. Scientific Data 3, 160050 (2016).CAS 

    Google Scholar 
    Vavourakis, C. D. et al. A metagenomics roadmap to the uncultured genome diversity in hypersaline soda lake sediments. Microbiome 6, 1–18 (2018).
    Google Scholar 
    Cabello-Yeves, P. J. et al. Microbiome of the deep Lake Baikal, a unique oxic bathypelagic habitat. Limnology and Oceanography 65, 1471–1488 (2020).ADS 
    CAS 

    Google Scholar 
    Vavourakis, C. D. et al. Metagenomes and metatranscriptomes shed new light on the microbial-mediated sulfur cycle in a siberian soda lake. BMC Biology 17, 69 (2019).
    Google Scholar 
    Waterworth, S. C., Isemonger, E. W., Rees, E. R., Dorrington, R. A. & Kwan, J. C. Conserved bacterial genomes from two geographically isolated peritidal stromatolite formations shed light on potential functional guilds. Environmental Microbiology Reports 13, 126–137 (2021).CAS 

    Google Scholar 
    Huddy, R. J. et al. Thiocyanate and organic carbon inputs drive convergent selection for specific autotrophic Afipia and Thiobacillus strains within complex microbiomes. Frontiers in Microbiology 12, 643368 (2021).
    Google Scholar 
    Emerson, J. B. et al. Diverse sediment microbiota shape methane emission temperature sensitivity in Arctic lakes. Nature Communications 12, 5815 (2021).ADS 
    CAS 

    Google Scholar 
    Chiri, E. et al. Termite gas emissions select for hydrogenotrophic microbial communities in termite mounds. Proceedings of the National Academy of Sciences 118, e2102625118 (2021).CAS 

    Google Scholar 
    Gong, G., Zhou, S., Luo, R., Gesang, Z. & Suolang, S. Metagenomic insights into the diversity of carbohydrate-degrading enzymes in the yak fecal microbial community. BMC Microbiology 20, 302 (2020).
    Google Scholar 
    Zhou, S. et al. Characterization of metagenome-assembled genomes and carbohydrate-degrading genes in the gut microbiota of Tibetan pig. Frontiers in Microbiology 11, 595066 (2020).
    Google Scholar 
    Tully, B. J., Graham, E. D. & Heidelberg, J. F. The reconstruction of 2,631 draft metagenome-assembled genomes from the global oceans. Scientific Data 5, 170203 (2018).CAS 

    Google Scholar 
    Lavrinienko, A. et al. Two hundred and fifty-four metagenome-assembled bacterial genomes from the bank vole gut microbiota. Scientific Data 7, 312 (2020).CAS 

    Google Scholar 
    Peng, X. et al. Genomic and functional analyses of fungal and bacterial consortia that enable lignocellulose breakdown in goat gut microbiomes. Nature Microbiology 6, 499–511 (2021).CAS 

    Google Scholar 
    Dudek, N. K. et al. Novel microbial diversity and functional potential in the marine mammal oral microbiome. Current Biology 27, 3752–3762.e6 (2017).CAS 

    Google Scholar 
    Pinto, A. J. et al. Metagenomic evidence for the presence of comammox nitrospira-like bacteria in a drinking water system. mSphere 1, e00054–15 (2015).
    Google Scholar 
    Zaremba-Niedzwiedzka, K. et al. Asgard archaea illuminate the origin of eukaryotic cellular complexity. Nature 541, 353–358 (2017).ADS 
    CAS 

    Google Scholar 
    Nobu, M. K. et al. Catabolism and interactions of uncultured organisms shaped by eco-thermodynamics in methanogenic bioprocesses. Microbiome 8, 111 (2020).CAS 

    Google Scholar 
    Butterfield, C. N. et al. Proteogenomic analyses indicate bacterial methylotrophy and archaeal heterotrophy are prevalent below the grass root zone. PeerJ 4, e2687 (2016).
    Google Scholar 
    Castelle, C. J. et al. Protein family content uncovers lineage relationships and bacterial pathway maintenance mechanisms in DPANN Archaea. Frontiers in Microbiology 12, 660052 (2021).
    Google Scholar 
    Alteio, L. V. et al. Complementary metagenomic approaches improve reconstruction of microbial diversity in a forest soil. mSystems 5, e00768–19 (2020).
    Google Scholar 
    Shaiber, A. et al. Functional and genetic markers of niche partitioning among enigmatic members of the human oral microbiome. Genome Biology 21, 292 (2020).
    Google Scholar 
    Jungbluth, S. P., Amend, J. P. & Rappé, M. S. Metagenome sequencing and 98 microbial genomes from Juan de Fuca Ridge flank subsurface fluids. Scientific Data 4, 170037 (2017).CAS 

    Google Scholar 
    Sheik, C. S. et al. Dolichospermum blooms in Lake Superior: DNA-based approach provides insight to the past, present and future of blooms. Journal of Great Lakes Research 48, 1191–1205 (2022).CAS 

    Google Scholar 
    Barnum, T. P. et al. Genome-resolved metagenomics identifies genetic mobility, metabolic interactions, and unexpected diversity in perchlorate-reducing communities. The ISME Journal 12, 1568–1581 (2018).CAS 

    Google Scholar 
    Julian, D. et al. Coastal ocean metagenomes and curated metagenome-assembled genomes from Marsh Landing, Sapelo Island (Georgia, USA). Microbiology Resource Announcements 8, e00934–19 (2019).
    Google Scholar 
    Breister, A. M. et al. Soil microbiomes mediate degradation of vinyl ester-based polymer composites. Communications Materials 1, 101 (2020).ADS 

    Google Scholar 
    Fu, H., Uchimiya, M., Gore, J. & Moran, M. A. Ecological drivers of bacterial community assembly in synthetic phycospheres. Proceedings of the National Academy of Sciences 117, 3656–3662 (2020).ADS 
    CAS 

    Google Scholar 
    Nobu, M. K. et al. Thermodynamically diverse syntrophic aromatic compound catabolism. Environmental Microbiology 19, 4576–4586 (2017).CAS 

    Google Scholar 
    Pasolli, E. et al. Extensive unexplored human microbiome diversity revealed by over 150,000 genomes from metagenomes spanning age, geography, and lifestyle. Cell 176, 649–662 (2019).CAS 

    Google Scholar 
    Nayfach, S. et al. A genomic catalog of Earth’s microbiomes. Nature Biotechnology 39, 499–509 (2021).CAS 

    Google Scholar 
    Li, Z. et al. Deep sea sediments associated with cold seeps are a subsurface reservoir of viral diversity. The ISME Journal 15, 2366–2378 (2021).CAS 

    Google Scholar 
    Bay, S. K. et al. Trace gas oxidizers are widespread and active members of soil microbial communities. Nature Microbiology 6, 246–256 (2021).CAS 

    Google Scholar 
    Seyler, L. M., Trembath-Reichert, E., Tully, B. J. & Huber, J. A. Time-series transcriptomics from cold, oxic subseafloor crustal fluids reveals a motile, mixotrophic microbial community. The ISME Journal 15, 1192–1206 (2021).CAS 

    Google Scholar 
    Herold, M. et al. Integration of time-series meta-omics data reveals how microbial ecosystems respond to disturbance. Nature Communications 11, 5281 (2020).ADS 
    CAS 

    Google Scholar 
    Dong, X. et al. Thermogenic hydrocarbon biodegradation by diverse depth-stratified microbial populations at a Scotian Basin cold seep. Nature Communications 11, 5825 (2020).ADS 
    CAS 

    Google Scholar 
    Thompson, L. R. et al. Metagenomic covariation along densely sampled environmental gradients in the Red Sea. The ISME Journal 11, 138–151 (2017).CAS 

    Google Scholar 
    Dominik, S., Daniela, Z., Anja, P., Katharina, R. & Rolf, D. Metagenome-assembled genome sequences from different wastewater treatment stages in Germany. Microbiology Resource Announcements 10, e00504–21 (2021).
    Google Scholar 
    Langwig, M. V. et al. Large-scale protein level comparison of Deltaproteobacteria reveals cohesive metabolic groups. The ISME Journal 16, 307–320 (2022).CAS 

    Google Scholar 
    Rezaei Somee, M. et al. Distinct microbial community along the chronic oil pollution continuum of the Persian Gulf converge with oil spill accidents. Scientific Reports 11, 11316 (2021).ADS 
    CAS 

    Google Scholar 
    Gilroy, R. et al. Metagenomic investigation of the equine faecal microbiome reveals extensive taxonomic diversity. PeerJ 10, e13084 (2022).
    Google Scholar 
    Bhattarai, B., Bhattacharjee, A. S., Coutinho, F. H. & Goel, R. K. Viruses and their interactions with bacteria and archaea of hypersaline Great Salt Lake. Frontiers in Microbiology 12, 701414 (2021).
    Google Scholar 
    Liu, L. et al. Microbial diversity and adaptive strategies in the Mars-like Qaidam Basin, North Tibetan Plateau, China. Environmental Microbiology Reports (2022).Lin, H. et al. Mercury methylation by metabolically versatile and cosmopolitan marine bacteria. The ISME Journal 15, 1810–1825 (2021).CAS 

    Google Scholar 
    Martnez-Pérez, C. et al. Lifting the lid: nitrifying archaea sustain diverse microbial communities below the Ross Ice Shelf. SSRN (2020).Zhang, L. et al. Metagenomic insights into the effect of thermal hydrolysis pre-treatment on microbial community of an anaerobic digestion system. Science of The Total Environment 791, 148096 (2021).ADS 
    CAS 

    Google Scholar 
    Starr, E. P. et al. Stable-isotope-informed, genome-resolved metagenomics uncovers potential cross-kingdom interactions in rhizosphere soil. mSphere 6, e00085–21 (2021).CAS 

    Google Scholar 
    Matthew, C. et al. Archaeal and bacterial metagenome-assembled genome sequences derived from pig feces. Microbiology Resource Announcements 11, 01142–21 (2022).
    Google Scholar 
    Wang, Y., Zhao, R., Liu, L., Li, B. & Zhang, T. Selective enrichment of comammox from activated sludge using antibiotics. Water Research 197, 117087 (2021).CAS 

    Google Scholar 
    Gilroy, R. et al. Extensive microbial diversity within the chicken gut microbiome revealed by metagenomics and culture. PeerJ 9, e10941 (2021).
    Google Scholar 
    Chen, Y. H. et al. Salvaging high-quality genomes of microbial species from a meromictic lake using a hybrid sequencing approach. Communications Biology 4, 996 (2021).CAS 

    Google Scholar 
    Beach, N. K., Myers, K. S., Donohue, T. J. & Noguera, D. R. Metagenomes from 25 low-abundance microbes in a partial nitritation anammox microbiome. Microbiology Resource Announcements 11, 00212–22 (2022).CAS 

    Google Scholar 
    Solanki, V. et al. Glycoside hydrolase from the GH76 family indicates that marine Salegentibacter sp. Hel_I_6 consumes alpha-mannan from fungi. The ISME Journal 16, 1818–1830 (2022).CAS 

    Google Scholar 
    Hiraoka, S. et al. Diverse DNA modification in marine prokaryotic and viral communities. Nucleic Acids Research 50, 1531–1550 (2022).CAS 

    Google Scholar 
    Haryono, M.A.S. et al. Recovery of high quality metagenome-assembled genomes from full-scale activated sludge microbial communities in a tropical climate using longitudinal metagenome sampling. Frontiers in Microbiology 13 (2022).Rodrguez-Ramos, J.A. et al. Microbial genome-resolved metaproteomic analyses frame intertwined carbon and nitrogen cycles in river hyporheic sediments. Research Square (2021).Kim, M., Cho, H. & Lee, W. Y. Distinct gut microbiotas between southern elephant seals and Weddell seals of Antarctica. Journal of Microbiology 58, 1018–1026 (2020).CAS 

    Google Scholar 
    Voorhies, A. A. et al. Cyanobacterial life at low O2: community genomics and function reveal metabolic versatility and extremely low diversity in a Great Lakes sinkhole mat. Geobiology 10, 250–267 (2012).CAS 

    Google Scholar 
    McDaniel, E. A. et al. Tbasco: trait-based comparative ‘omics identifies ecosystem-level and niche-differentiating adaptations of an engineered microbiome. ISME Communications 2, 111 (2022).
    Google Scholar 
    Wang, W. et al. Contrasting bacterial and archaeal distributions reflecting different geochemical processes in a sediment core from the Pearl River Estuary. AMB Express 10, 16 (2020).
    Google Scholar 
    Mandakovic, D. et al. Genome-scale metabolic models of Microbacterium species isolated from a high altitude desert environment. Scientific Reports 10, 5560 (2020).ADS 
    CAS 

    Google Scholar 
    Wang, Y. et al. Seasonal prevalence of ammonia-oxidizing archaea in a full-scale municipal wastewater treatment plant treating saline wastewater revealed by a 6-year time-series analysis. Environmental Science & Technology 55, 2662–2673 (2021).ADS 
    CAS 

    Google Scholar 
    Bulzu, P. A. et al. Casting light on Asgardarchaeota metabolism in a sunlit microoxic niche. Nature Microbiology 4, 1129–1137 (2019).CAS 

    Google Scholar 
    Karen, J. et al. Hydrogen-oxidizing bacteria are abundant in desert soils and strongly stimulated by hydration. mSystems 5, e01131–20 (2020).
    Google Scholar 
    Rust, M. et al. A multiproducer microbiome generates chemical diversity in the marine sponge Mycale hentscheli. Proceedings of the National Academy of Sciences 117, 9508–9518 (2020).ADS 
    CAS 

    Google Scholar 
    Podowski, J. C., Paver, S. F., Newton, R. J. & Coleman, M. L. Genome streamlining, proteorhodopsin, and organic nitrogen metabolism in freshwater nitrifiers. mBio 13, e02379–21 (2022).
    Google Scholar 
    Coutinho, F. H. et al. New viral biogeochemical roles revealed through metagenomic analysis of Lake Baikal. Microbiome 8, 163 (2020).CAS 

    Google Scholar 
    Philippi, M. et al. Purple sulfur bacteria fix N2 via molybdenum-nitrogenase in a low molybdenum Proterozoic ocean analogue. Nature Communications 12, 4774 (2021).ADS 
    CAS 

    Google Scholar 
    Katie, S. et al. Eight metagenome-assembled genomes provide evidence for microbial adaptation in 20,000- to 1,000,000-year-old Siberian permafrost. Applied and Environmental Microbiology 87, e00972–21 (2021).
    Google Scholar 
    Mert, K. et al. Unexpected abundance and diversity of phototrophs in mats from morphologically variable microbialites in Great Salt Lake, Utah. Applied and Environmental Microbiology 86, e00165–20 (2020).
    Google Scholar 
    Patin, N. V. et al. Gulf of Mexico blue hole harbors high levels of novel microbial lineages. The ISME Journal 15, 2206–2232 (2021).CAS 

    Google Scholar 
    Wang, J., Tang, X., Mo, Z. & Mao, Y. Metagenome-assembled genomes from Pyropia haitanensis microbiome provide insights into the potential metabolic functions to the seaweed. Frontiers in Microbiology 13, 857901 (2022).
    Google Scholar 
    Burgsdorf, I. et al. Lineage-specific energy and carbon metabolism of sponge symbionts and contributions to the host carbon pool. The ISME Journal 16, 1163–1175 (2022).CAS 

    Google Scholar 
    Suarez, C. et al. Disturbance-based management of ecosystem services and disservices in partial nitritation-anammox biofilms. npj Biofilms and Microbiomes 8, 47 (2022).CAS 

    Google Scholar 
    Kumar, D. et al. Textile industry wastewaters from Jetpur, Gujarat, India, are dominated by Shewanellaceae, Bacteroidaceae, and Pseudomonadaceae harboring genes encoding catalytic enzymes for textile dye degradation. Frontiers in Environmental Science 9, 720707 (2021).ADS 

    Google Scholar 
    Seitz, V. A. et al. Variation in root exudate composition influences soil microbiome membership and function. Applied and Environmental Microbiology 88, e00226–22 (2022).
    Google Scholar 
    Lindner, B. G. et al. Toward shotgun metagenomic approaches for microbial source tracking sewage spills based on laboratory mesocosms. Water Research 210, 117993 (2022).CAS 

    Google Scholar 
    Yancey, C. E. et al. Metagenomic and metatranscriptomic insights into population diversity of microcystis blooms: Spatial and temporal dynamics of mcy genotypes, including a partial operon that can be abundant and expressed. Applied and Environmental Microbiology 88, e02464–21 (2022).
    Google Scholar 
    Liu, L. et al. Charting the complexity of the activated sludge microbiome through a hybrid sequencing strategy. Microbiome 9, 205 (2021).CAS 

    Google Scholar 
    Speth, D. R. et al. Microbial communities of Auka hydrothermal sediments shed light on vent biogeography and the evolutionary history of thermophily. The ISME Journal 16, 1750–1764 (2022).CAS 

    Google Scholar 
    Blyton, M. D. J., Soo, R. M., Hugenholtz, P. & Moore, B. D. Maternal inheritance of the koala gut microbiome and its compositional and functional maturation during juvenile development. Environmental Microbiology 24, 475–493 (2022).CAS 

    Google Scholar 
    Nuccio, E. E. et al. Niche differentiation is spatially and temporally regulated in the rhizosphere. The ISME Journal 14, 999–1014 (2020).CAS 

    Google Scholar 
    Jaffe, A. L. et al. Long-term incubation of lake water enables genomic sampling of consortia involving planctomycetes and candidate phyla radiation bacteria. mSystems 7, e00223–22 (2022).
    Google Scholar 
    Cabral, L. et al. Gut microbiome of the largest living rodent harbors unprecedented enzymatic systems to degrade plant polysaccharides. Nature Communications 13, 629 (2022).ADS 
    CAS 

    Google Scholar 
    Blyton, M. D. J., Soo, R. M., Hugenholtz, P. & Moore, B. D. Characterization of the juvenile koala gut microbiome across wild populations. Environmental Microbiology 24, 4209–4219 (2022).CAS 

    Google Scholar 
    Xu, B. et al. A holistic genome dataset of bacteria, archaea and viruses of the Pearl River estuary. Scientific Data 9, 49 (2022).MathSciNet 
    CAS 

    Google Scholar 
    Royo-Llonch, M. et al. Compendium of 530 metagenome-assembled bacterial and archaeal genomes from the polar Arctic Ocean. Nature Microbiology 6, 1561–1574 (2021).CAS 

    Google Scholar 
    Sun, J., Prabhu, A., Aroney, S. T. N. & Rinke, C. Insights into plastic biodegradation: community composition and functional capabilities of the superworm (Zophobas morio) microbiome in styrofoam feeding trials. Microbial Genomics 8, 000842 (2022).CAS 

    Google Scholar 
    Kim, M. et al. Higher pathogen load in children from Mozambique vs. USA revealed by comparative fecal microbiome profiling. ISME Communications 2, 74 (2022).ADS 

    Google Scholar 
    Kelly, J. B., Carlson, D. E., Low, J. S. & Thacker, R. W. Novel trends of genome evolution in highly complex tropical sponge microbiomes. Microbiome 10, 164 (2022).CAS 

    Google Scholar 
    Bray, M. S. et al. Phylogenetic and structural diversity of aromatically dense pili from environmental metagenomes. Environmental Microbiology Reports 12, 49–57 (2020).CAS 

    Google Scholar 
    Cabello-Yeves, P. J. et al. α-cyanobacteria possessing form IA RuBisCO globally dominate aquatic habitats. The ISME Journal 16, 2421–2432 (2022).CAS 

    Google Scholar 
    Berben, T. et al. The Polar Fox Lagoon in Siberia harbours a community of Bathyarchaeota possessing the potential for peptide fermentation and acetogenesis. Antonie van Leeuwenhoek 115, 1229–1244 (2022).CAS 

    Google Scholar 
    Tamburini, F. B. et al. Short- and long-read metagenomics of urban and rural South African gut microbiomes reveal a transitional composition and undescribed taxa. Nature Communications 13, 926 (2022).ADS 
    CAS 

    Google Scholar 
    Kantor, R. S., Miller, S. E. & Nelson, K. L. The water microbiome through a pilot scale advanced treatment facility for direct potable reuse. Frontiers in Microbiology 10, 993 (2019).
    Google Scholar 
    Muratore, D. et al. Complex marine microbial communities partition metabolism of scarce resources over the diel cycle. Nature Ecology & Evolution 6, 218–229 (2022).
    Google Scholar 
    Zhou, Y. L., Mara, P., Cui, G. J., Edgcomb, V. P. & Wang, Y. Microbiomes in the challenger deep slope and bottom-axis sediments. Nature Communications 13, 1515 (2022).ADS 
    CAS 

    Google Scholar 
    Zhang, H. et al. Metagenome sequencing and 768 microbial genomes from cold seep in South China Sea. Scientific Data 9, 480 (2022).CAS 

    Google Scholar 
    Zhuang, J. L., Zhou, Y. Y., Liu, Y. D. & Li, W. Flocs are the main source of nitrous oxide in a high-rate anammox granular sludge reactor: insights from metagenomics and fed-batch experiments. Water Research 186, e116321 (2020).
    Google Scholar 
    Shiffman, M. E. et al. Gene and genome-centric analyses of koala and wombat fecal microbiomes point to metabolic specialization for eucalyptus digestion. PeerJ 5, 4075 (2017).
    Google Scholar 
    Murphy, S. M. C., Bautista, M. A., Cramm, M. A. & Hubert, C. R. J. Diesel and crude oil biodegradation by cold-adapted microbial communities in the Labrador Sea. Applied and Environmental Microbiology 87, e00800–21 (2021).ADS 
    CAS 

    Google Scholar 
    Suarez, C. et al. Metagenomic evidence of a novel family of anammox bacteria in a subsea environment. Environmental Microbiology 24, 2348–2360 (2022).CAS 

    Google Scholar 
    Dharamshi, J.E. et al. Genomic diversity and biosynthetic capabilities of sponge-associated chlamydiae. The ISME Journal (2022).Florian, P. O., Hugo, R. & Mathieu, A. Recovery of metagenome-assembled genomes from a human fecal sample with pacific biosciences high-fidelity sequencing. Microbiology Resource Announcements 11, e00250–22 (2022).
    Google Scholar 
    Bloom, S. M. et al. Cysteine dependence of Lactobacillus iners is a potential therapeutic target for vaginal microbiota modulation. Nature Microbiology 7, 434–450 (2022).CAS 

    Google Scholar 
    Aylward, F. O. et al. Diel cycling and long-term persistence of viruses in the ocean’s euphotic zone. Proceedings of the National Academy of Sciences 114, 11446–11451 (2017).ADS 
    CAS 

    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 Research 25, 1043–1055 (2015).CAS 

    Google Scholar 
    Bowers, R. M. et al. Minimum information about a single amplified genome (MISAG) and a metagenome-assembled genome (MIMAG) of bacteria and archaea. Nature biotechnology 35, 725 (2017).CAS 

    Google Scholar 
    Chaumeil, P. A., Mussig, A. J., Hugenholtz, P. & Parks, D. H. GTDB-Tk: a toolkit to classify genomes with the Genome Taxonomy Database. Bioinformatics 36, 1925–1927 (2020).CAS 

    Google Scholar 
    Louca, S. The rates of global bacterial and archaeal dispersal. ISME Journal 16, 159–167 (2021).ADS 

    Google Scholar 
    Ondov, B. D. et al. Mash: fast genome and metagenome distance estimation using minhash. Genome Biology 17, 132 (2016).
    Google Scholar 
    Müllner, D. fastcluster: Fast hierarchical, agglomerative clustering routines for R and Python. Journal of Statistical Software 53, 1–18 (2013).
    Google Scholar 
    Kinene, T., Wainaina, J., Maina, S., Boykin, L.M. & Kliman, R.M. Methods for rooting trees, vol. 3, 489–493 (Academic Press, Oxford, 2016).Louca, S. & Doebeli, M. Efficient comparative phylogenetics on large trees. Bioinformatics 34, 1053–1055 (2018).CAS 

    Google Scholar 
    Rees, J. A. & Cranston, K. Automated assembly of a reference taxonomy for phylogenetic data synthesis. Biodiversity Data Journal 5, e12581 (2017).
    Google Scholar 
    Heck, K. et al. Evaluating methods for purifying cyanobacterial cultures by qPCR and high-throughput Illumina sequencing. Journal of Microbiological Methods 129, 55–60 (2016).CAS 

    Google Scholar 
    Cornet, L. et al. Consensus assessment of the contamination level of publicly available cyanobacterial genomes. PLOS ONE 13, e0200323 (2018).
    Google Scholar 
    Alneberg, J. et al. Genomes from uncultivated prokaryotes: a comparison of metagenome-assembled and single-amplified genomes. Microbiome 6, 173 (2018).
    Google Scholar 
    Eddy, S. R. Accelerated profile HMM searches. PLoS Computational Biology 7, e1002195 (2011).ADS 
    MathSciNet 
    CAS 

    Google Scholar 
    Buchfink, B., Xie, C. & Huson, D. H. Fast and sensitive protein alignment using DIAMOND. Nature Methods 12, 59–60 (2014).
    Google Scholar 
    Pedregosa, F. et al. Scikit-learn: Machine Learning in Python. Journal of Machine Learning Research 12, 2825–2830 (2011).MathSciNet 
    MATH 

    Google Scholar  More

  • in

    Climate-trait relationships exhibit strong habitat specificity in plant communities across Europe

    Bjorkman, A. D. et al. Plant functional trait change across a warming tundra biome. Nature 562, 57–62 (2018).ADS 
    CAS 

    Google Scholar 
    Sabatini, F. M. et al. Global patterns of vascular plant alpha diversity. Nat. Commun. 13, 4683 (2022).ADS 
    CAS 

    Google Scholar 
    Lavorel, S. & Garnier, E. Predicting changes in community composition and ecosystem functioning from plant traits: revisiting the Holy Grail. Funct. Ecol. 16, 545–556 (2002).
    Google Scholar 
    Chapin, F. S. III et al. Consequences of changing biodiversity. Nature 405, 234–242 (2000).CAS 

    Google Scholar 
    Garnier, E., Navas, M.-L. & Grigulis, K. Plant functional diversity. Organism traits, community structure, and ecosystem properties (Oxford University Press, Oxford, New York, NY, 2016).Funk, J. L. et al. Revisiting the Holy Grail: using plant functional traits to understand ecological processes. Biol. Rev. Camb. Philos. Soc. 92, 1156–1173 (2017).
    Google Scholar 
    Díaz, S. et al. The global spectrum of plant form and function. Nature 529, 167–171 (2016).ADS 

    Google Scholar 
    Adler, P. B. et al. Functional traits explain variation in plant life history strategies. Proc. Natl. Acad. Sci. U.S.A. 111, 740–745 (2014).ADS 
    CAS 

    Google Scholar 
    Wright, I. J. et al. The worldwide leaf economics spectrum. Nature 428, 821–827 (2004).ADS 
    CAS 

    Google Scholar 
    Salguero-Gómez, R. et al. Fast-slow continuum and reproductive strategies structure plant life-history variation worldwide. Proc. Natl. Acad. Sci. U. S. A. 113, 230–235 (2016).ADS 

    Google Scholar 
    Bergmann, J. et al. The fungal collaboration gradient dominates the root economics space in plants. Sci. Adv. 6, eaba3756 (2020).ADS 
    CAS 

    Google Scholar 
    Shipley, B. et al. Reinforcing loose foundation stones in trait-based plant ecology. Oecologia 180, 923–931 (2016).ADS 

    Google Scholar 
    Bruelheide, H. et al. Global trait-environment relationships of plant communities. Nat. Ecol. Evol. 2, 1906–1917 (2018).
    Google Scholar 
    McGill, B. J., Enquist, B. J., Weiher, E. & Westoby, M. Rebuilding community ecology from functional traits. Trends Ecol. Evol. 21, 178–185 (2006).
    Google Scholar 
    Miller, J. E. D., Damschen, E. I. & Ives, A. R. Functional traits and community composition: A comparison among community‐weighted means, weighted correlations, and multilevel models. Methods Ecol. Evol. 10, 415–425 (2019).
    Google Scholar 
    Guerin, G. R. et al. Environmental associations of abundance-weighted functional traits in Australian plant communities. Basic Appl. Ecol. 58, 98–109 (2021).
    Google Scholar 
    Walter, H. Vegetation of the earth and ecological systems of the geo-biosphere (Springer-Verlag, Berlin, Germany, 1985).Ordoñez, J. C. et al. A global study of relationships between leaf traits, climate and soil measures of nutrient fertility. Glob. Ecol. Biogeogr. 18, 137–149 (2009).
    Google Scholar 
    Simpson, A. H., Richardson, S. J. & Laughlin, D. C. Soil-climate interactions explain variation in foliar, stem, root and reproductive traits across temperate forests. Glob. Ecol. Biogeogr. 25, 964–978 (2016).
    Google Scholar 
    Cubino, J. P. et al. The leaf economic and plant size spectra of European forest understory vegetation. Ecography 44, 1311–1324 (2021).
    Google Scholar 
    Garnier, E. et al. Assessing the effects of land-use change on plant traits, communities and ecosystem functioning in grasslands: a standardized methodology and lessons from an application to 11 European sites. Ann. Bot. 99, 967–985 (2007).
    Google Scholar 
    Herben, T., Klimešová, J. & Chytrý, M. Effects of disturbance frequency and severity on plant traits: An assessment across a temperate flora. Funct. Ecol. 32, 799–808 (2018).
    Google Scholar 
    Linder, H. P. et al. Biotic modifiers, environmental modulation and species distribution models. J. Biogeogr. 39, 2179–2190 (2012).
    Google Scholar 
    Gross, N. et al. Linking individual response to biotic interactions with community structure: a trait-based framework. Funct. Ecol. 23, 1167–1178 (2009).
    Google Scholar 
    Ordonez, A. & Svenning, J.-C. Consistent role of Quaternary climate change in shaping current plant functional diversity patterns across European plant orders. Sci. Rep. 7, 42988 (2017).ADS 
    CAS 

    Google Scholar 
    Kemppinen, J. et al. Consistent trait–environment relationships within and across tundra plant communities. Nat. Ecol. Evol. 5, 458–467 (2021).
    Google Scholar 
    Chytrý, M. et al. European Vegetation Archive (EVA): an integrated database of European vegetation plots. Appl. Veg. Sci. 19, 173–180 (2016).
    Google Scholar 
    Karger, D. N. et al. Data from: Climatologies at high resolution for the earth’s land surface areas. EnviDat, https://doi.org/10.16904/envidat.228 (2018).Karger, D. N. et al. Climatologies at high resolution for the earth’s land surface areas. Sci. Data 4, 170122 (2017).
    Google Scholar 
    Kattge, J. et al. TRY plant trait database – enhanced coverage and open access. Glob. Change. Biol. 26, 119–188 (2020).ADS 

    Google Scholar 
    Laughlin, D. C., Leppert, J. J., Moore, M. M. & Sieg, C. H. A multi-trait test of the leaf-height-seed plant strategy scheme with 133 species from a pine forest flora. Funct. Ecol. 24, 493–501 (2010).
    Google Scholar 
    Davies, C. E., Moss, D. & Hill, M. O. EUNIS Habitat Classification Revised 2004. Report to: European Environment Agency, European Topic Centre on Nature Protection and Biodiversity, 2004.Chytrý, M. et al. EUNIS Habitat Classification: Expert system, characteristic species combinations and distribution maps of European habitats. Appl. Veg. Sci. 23, 648–675 (2020).
    Google Scholar 
    Pausas, J. G. & Bond, W. J. Humboldt and the reinvention of nature. J. Ecol. 107, 1031–1037 (2019).
    Google Scholar 
    Meng, T.-T. et al. Responses of leaf traits to climatic gradients: adaptive variation versus compositional shifts. Biogeosciences 12, 5339–5352 (2015).ADS 

    Google Scholar 
    Fang, J. et al. Precipitation patterns alter growth of temperate vegetation. Geophys. Res. Lett. 32, 81 (2005).
    Google Scholar 
    Butler, E. E. et al. Mapping local and global variability in plant trait distributions. Proc. Natl. Acad. Sci. U.S.A. 114, E10937–E10946 (2017).ADS 
    CAS 

    Google Scholar 
    Gong, H. & Gao, J. Soil and climatic drivers of plant SLA (specific leaf area). Glob. Ecol. Conserv. 20, e00696 (2019).
    Google Scholar 
    Laughlin, D. C. et al. Root traits explain plant species distributions along climatic gradients yet challenge the nature of ecological trade-offs. Nat. Ecol. Evol. 5, 1–12 (2021).
    Google Scholar 
    Carmona, C. P. et al. Fine-root traits in the global spectrum of plant form and function. Nature 597, 683–687 (2021).ADS 
    CAS 

    Google Scholar 
    Ding, J., Travers, S. K. & Eldridge, D. J. Occurrence of Australian woody species is driven by soil moisture and available phosphorus across a climatic gradient. J. Veg. Sci. 32, e13095 (2021).
    Google Scholar 
    Falster, D. S. & Westoby, M. Plant height and evolutionary games. Trends Ecol. Evol. 18, 337–343 (2003).
    Google Scholar 
    Kunstler, G. et al. Plant functional traits have globally consistent effects on competition. Nature 529, 204–207 (2016).ADS 
    CAS 

    Google Scholar 
    McLachlan, A. & Brown, A. C. Coastal Dune Ecosystems and Dune/Beach Interactions. In The Ecology of Sandy Shores (Elsevier), 251–271 (2006).Cui, E., Weng, E., Yan, E. & Xia, J. Robust leaf trait relationships across species under global environmental changes. Nat. Commun. 11, 1–9 (2020).ADS 

    Google Scholar 
    Cain, S. A. Life-Forms and Phytoclimate. Bot. Rev. 16, 1–32 (1950).
    Google Scholar 
    Yu, S. et al. Shift of seed mass and fruit type spectra along longitudinal gradient: high water availability and growth allometry. Biogeosciences 18, 655–667 (2021).ADS 

    Google Scholar 
    Murray, B. R., Brown, A. H. D., Dickman, C. R. & Crowther, M. S. Geographical gradients in seed mass in relation to climate. J. Biogeogr. 31, 379–388 (2004).
    Google Scholar 
    Metz, J. et al. Plant survival in relation to seed size along environmental gradients: a long-term study from semi-arid and Mediterranean annual plant communities. J. Ecol. 98, 697–704 (2010).
    Google Scholar 
    Tao, S., Guo, Q., Li, C., Wang, Z. & Fang, J. Global patterns and determinants of forest canopy height. Ecology 97, 3265–3270 (2016).
    Google Scholar 
    Gonzalez, P., Neilson, R. P., Lenihan, J. M. & Drapek, R. J. Global patterns in the vulnerability of ecosystems to vegetation shifts due to climate change. Glob. Ecol. Biogeogr. 19, 755–768 (2010).
    Google Scholar 
    Feeley, K. J., Bravo-Avila, C., Fadrique, B., Perez, T. M. & Zuleta, D. Climate-driven changes in the composition of New World plant communities. Nat. Clim. Chang. 10, 965–970 (2020).ADS 
    CAS 

    Google Scholar 
    Bruelheide, H. et al. sPlot—A new tool for global vegetation analyses. J. Veg. Sci. 30, 161–186 (2019).
    Google Scholar 
    Schrodt, F. et al. BHPMF—a hierarchical Bayesian approach to gap-filling and trait prediction for macroecology and functional biogeography. Glob. Ecol. Biogeogr. 24, 1510–1521 (2015).
    Google Scholar 
    Shan, H. et al. Gap filling in the plant kingdom—trait prediction using hierarchical probabilistic matrix factorization (Proceedings of the 29 th International Conference on Machine Learning, Edinburgh, Scotland, UK, 2012).Chytrý, M. et al. EUNIS-ESy, version 2021-06-01, https://doi.org/10.5281/zenodo.4812736 (2021).Wood, S. N., Pya, N. & Säfken, B. Smoothing Parameter and Model Selection for General Smooth Models. J. Am. Stat. Assoc. 111, 1548–1563 (2016).MathSciNet 
    CAS 

    Google Scholar 
    Wood, S. N. Generalized Additive Models. An Introduction with R, Second Edition (CRC Press, Portland, Oregon, USA, 2017).Nakagawa, S. & Schielzeth, H. A general and simple method for obtaining R2 from generalized linear mixed-effects models. Methods Ecol. Evol. 4, 133–142 (2013).
    Google Scholar 
    Johnson, P. C. Extension of Nakagawa & Schielzeth’s R2GLMM to random slopes models. Methods Ecol. Evol. 5, 944–946 (2014).
    Google Scholar 
    R. Core Team. R: a language and environment for statistical computing (R Foundation for Statistical Computing, 2022).Lenth, R. V. et al. emmeans: estimated marginal means, aka least-squares means; R package version 1.6.2-1 (2021).Lüdecke, D. ggeffects: tidy data frames of marginal effects from regression models. J. Open Source Softw. 3, 772 (2018).ADS 

    Google Scholar 
    Hijmans, R.J., Phillips, S., Leathwick, J. & Elith, J. dismo: species distribution modelling; R package version 1.3-3 (2020).Wickham, H. ggplot2: Elegant Graphics for Data Analysis (Springer-Verlag New York, 2016).Kambach, S. Habitat-specificity of climate-trait relationships in plant communities across Europe. github.com/StephanKambach, version 1.0; https://doi.org/10.5281/zenodo.7404176 (2022).Moles, A. T. et al. Global patterns in plant height. J. Ecol. 97, 923–932 (2009).
    Google Scholar 
    Moles, A. T. et al. Global patterns in seed size. Glob. Ecol. Biogeogr. 16, 109–116 (2007).
    Google Scholar 
    Zheng, J., Guo, Z. & Wang, X. Seed mass of angiosperm woody plants better explained by life history traits than climate across China. Sci. Rep. 7, 2741 (2017).ADS 

    Google Scholar 
    Saatkamp, A. et al. A research agenda for seed-trait functional ecology. N. Phytol. 221, 1764–1775 (2019).
    Google Scholar 
    Freschet, G. T. et al. Climate, soil and plant functional types as drivers of global fine‐root trait variation. J. Ecol. 105, 1182–1196 (2017).
    Google Scholar 
    Weigelt, A. et al. An integrated framework of plant form and function: The belowground perspective. N. Phytol. 232, 42–59 (2021).
    Google Scholar  More

  • in

    Genetic population structures of common scavenging species near hydrothermal vents in the Okinawa Trough

    Van Dover, C. L. et al. Environmental management of deep-sea chemosynthetic ecosystems: justification of and considerations for a spatially based approach. ISA Technical Study: No.9. (International Seabed Authority, 2011).Ikehata, K., Suzuki, R., Shimada, K., Ishibashi, J., & Urabe, T. Mineralogical and Geochemical Characteristics of Hydrothermal Minerals Collected from Hydrothermal Vent Fields in the Southern Mariana Spreading Center. In Subseafloor biosphere linked to hydrothermal systems: TAIGA Concept. 275–288 (Springer Tokyo, 2015).Rona, P. A. & Scott, S. D. A special issue on sea-floor hydrothermal mineralization; new perspectives; preface. Econ. Geol. 88, 1935–1976 (1993).
    Google Scholar 
    Glasby, G. P., Iizasa, K., Yuasa, M. & Usui, A. Submarine hydrothermal mineralization on the Izu-Bonin arc, south of Japan: an overview. Mar. Georesources Geotech. 18, 141–176 (2000).
    Google Scholar 
    Van Dover, C. L. Inactive sulfide ecosystems in the deep sea: a review. Front. Mar. Sci. 6, 461. https://doi.org/10.3389/fmars.2019.00461 (2019).Article 

    Google Scholar 
    Boschen, R. E., Rowde, A. A., Clark, M. R. & Gardner, J. P. Mining of deep-sea seafloor massive sulfides: a review of the deposits, their benthic communities, impacts from mining, regulatory frameworks and management strategies. Ocean Coast. Manag. 84, 54–67 (2013).
    Google Scholar 
    Washburn, T. W. et al. Ecological risk assessment for deep-sea mining. Ocean Coast. Manag. 176, 24–39 (2019).
    Google Scholar 
    Matsui, T., Sugishima, H., Okamoto, N., Igarashi, Y. Evaluation of turbidity and resedimentation through seafloor disturbance experiments for assessment of environmental impacts associated with exploitation of seafloor massive sulfides mining. Proceedings of the Twenty-eighth. International Ocean and Polar Engineering Conference. 144–151 (2018).International Seabed Authority. Recommendations for the guidance of contractors for the assessment of the possible environmental impacts arising from exploration for marine minerals in the Area. https://www.isa.org.jm/documents/isba19ltc8 (2013).Suzuki, K., Yoshida, K., Watanabe, H. & Yamamoto, H. Mapping the resilience of chemosynthetic communities in hydrothermal vent fields. Sci. Rep. 8, 9364. https://doi.org/10.1038/s41598-018-27596-7 (2018).Article 
    ADS 
    CAS 

    Google Scholar 
    Yahagi, T., Watanabe, H., Ishibashi, J. I. & Kojima, S. Genetic population structure of four hydrothermal vent shrimp species (Alvinocarididae) in the Okinawa Trough, Northwest Pacific. Mar. Ecol. Prog. Ser. 529, 159–169 (2015).ADS 

    Google Scholar 
    Mullineaux, L. S. Deep-sea hydrothermal vent communities. In Marine community ecology and conservation (eds Bertness, M. D. et al.) 383–400 (Sinauer, 2013).
    Google Scholar 
    Van Dover, C. L., German, C. R., Speer, K. G., Parson, L. M. & Vrijenhoek, R. C. Evolution and biogeography of deep-sea vent and seep invertebrates. Science 295, 1253–1257 (2002).ADS 

    Google Scholar 
    Yahagi, T., Kayama-Watanabe, H., Kojima, S. & Kano, Y. Do larvae from deep-sea hydrothermal vents disperse in surface waters?. Ecology 98, 1524–1534 (2017).
    Google Scholar 
    Hebert, P. D. & Gregory, T. R. The promise of DNA barcoding for taxonomy. Syst. Biol. 54, 852–859 (2005).
    Google Scholar 
    Iguchi, A. et al. Comparative analysis on the genetic population structures of the deep-sea whelks Buccinum tsubai and Neptunea constricta in the Sea of Japan. Mar. Biol. 151, 31–39 (2007).
    Google Scholar 
    Goode, G. B. & Bean, T. H. A catalogue of the fishes of Essex County, Massachusetts, including the fauna of Massachusetts Bay and the contiguous deep waters. Bull. Essex Inst. 11, 1–38 (1879).
    Google Scholar 
    Johnson, J. Y. Descriptions of some new genera and species of fishes obtained at Madeira. Proc. Zool. Soc. Lond. 1862, 167–180 (1862).
    Google Scholar 
    Bate, C. S. Report on the Crustacea Macrura collected by the Challenger during the years 1873–76. Report on the scientific results of the Voyage of H.M.S. Challenger during the years 1873–76. Zoology 24, 1–942 (1888).
    Google Scholar 
    Folmer, O., Black, M., Hoeh, W. R., Lutz, R. & Vrijenhoek, R. C. DNA primers for amplification of mitochondrial cytochrome c oxidase subunit I from diverse metazoan invertebrates. Mol. Mar. Biol Biotech. 3, 294–299 (1994).CAS 

    Google Scholar 
    Pilgrim, E. M., Blum, M. J., Reusser, D. A., Lee, H. & Darling, J. A. Geographic range and structure of cryptic genetic diversity among Pacific North American populations of the non-native amphipod Grandidierella japonica. Biol. Invasions 15, 2415–2428 (2013).
    Google Scholar 
    Suyama, Y. & Matsuki, Y. MIG-seq: an effective PCR-based method for genome-wide single-nucleotide polymorphism genotyping using the next-generation sequencing platform. Sci. Rep. 5, 16963. https://doi.org/10.1038/srep16963 (2015).Article 
    ADS 
    CAS 

    Google Scholar 
    R Core Team. R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. http://www.R-project.org/ (2020).Paradis, E., Claude, J. & Strimmer, K. APE: analyses of phylogenetics and evolution in R language. Bioinformatics 20, 289–290 (2004).CAS 

    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 

    Google Scholar 
    Shen, W., Le, S., Li, Y. & Hu, F. SeqKit: a cross-platform and ultrafast toolkit for FASTA/Q file manipulation. PLoS ONE 11, e0163962. https://doi.org/10.1371/journal.pone.0163962 (2016).Article 
    CAS 

    Google Scholar 
    Paradis, E. pegas: an R package for population genetics with an integrated–modular approach. Bioinformatics 26, 419–420 (2010).CAS 

    Google Scholar 
    Kumar, S., Stecher, G. & Tamura, K. MEGA7: molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 33, 1870–1874 (2016).CAS 

    Google Scholar 
    Darriba, D. et al. ModelTest-NG: a new and scalable tool for the selection of DNA and protein evolutionary models. Mol. Biol. Evol. 37, 291–294 (2020).CAS 

    Google Scholar 
    Kozlov, A. M., Darriba, D., Flouri, T., Morel, B. & Stamatakis, A. RaxML-NG: a fast, scalable and user-friendly tool for maximum likelihood phylogenetic inference. Bioinformatics 35, 4453–4455 (2019).CAS 

    Google Scholar 
    Ronquist, F. R. & Huelsenbeck, J. P. MRBAYES 3: Bayesian inference of phylogeny. Bioinformatics 19, 1572–1574 (2003).CAS 

    Google Scholar 
    Puillandre, N., Brouillet, S. & Achaz, G. ASAP: assemble species by automatic partitioning. Mol. Ecol. Resour. 21, 609–620 (2021).
    Google Scholar 
    Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet.journal 17, http://journal.embnet.org/index.php/embnetjournal/article/view/200/479 (2011).Rochette, N. C., Rivera-Colón, A. G. & Catchen, J. M. Stacks 2: Analytical methods for paired-end sequencing improve RADseq-based population genomics. Mol. Ecol. 28, 4737–4754 (2019).CAS 

    Google Scholar 
    Purcell, S. et al. PLINK: a tool set for whole-genome association and population-based linkage analyses. Am. J. Hum. Genet. 81, 559–575 (2007).CAS 

    Google Scholar 
    Jombart, T. adegenet: a R package for the multivariate analysis of genetic markers. Bioinformatics 24, 1403–1405 (2008).CAS 

    Google Scholar 
    Goudet, J. Hierfstat, a package for R to compute and test hierarchical F-statistics. Mol. Ecol. Notes 5, 184–186 (2013).
    Google Scholar 
    Oksanen, J. et al. vegan: Community Ecology Package. R package version 2.5–6. https://CRAN.R-project.org/package=vegan (2019).Dana, J. D. Synopsis of the genera of Gammaracea. Am. J. Sci. Arts 8, 135–140 (1849).
    Google Scholar 
    Hansen, H. J. Malacostraca marina Groenlandiæ occidentalis Oversigt over det vestlige Grønlands Fauna af malakostrake Havkrebsdyr. Vidensk. Meddel. Natuirist. Foren Kjobenhavn, Aaret 9, 5–226 (1888).
    Google Scholar 
    Van Dover, C. L. The ecology of deep-sea hydrothermal vents (Princeton University Press, 2000).
    Google Scholar 
    Tunnicliffe, V. The biology of hydrothermal vents: ecology and evolution. Oceanogr. Mar. Biol. Annu. Rev. 29, 319–407 (1991).
    Google Scholar 
    Priede, I. G., Bagley, P. M., Smith, A., Creasey, S. & Merrett, N. R. Scavenging deep demersal fishes of the Porcupine Seabight, north-east Atlantic: observations by baited camera, trap and trawl. J. Mar. Biol. Assoc. U. K. 74, 481–498 (1994).
    Google Scholar 
    Causse, R., Biscoito, M. & Briand, P. First record of the deep-sea eel Ilyophis saldanhai (Synaphobranchidae, Anguilliformes) from the Pacific Ocean. Cybium 29, 413–416 (2005).
    Google Scholar 
    King, N. J., Bagley, P. M. & Priede, I. G. Depth zonation and latitudinal distribution of deep-sea scavenging demersal fishes of the Mid-Atlantic Ridge, 42 to 53°N. Mar. Ecol. Prog. Ser. 319, 263–274 (2006).ADS 

    Google Scholar 
    Leitner, A. B., Durden, J. M., Smith, C. R., Klingberg, E. D. & Drazen, J. C. Synaphobranchid eel swarms on abyssal seamounts: largest aggregation of fishes ever observed at abyssal depths. Deep Sea Res. Oceanogr. Res. Part I Pap. 167, 103423. https://doi.org/10.1016/j.dsr.2020.103423 (2021).Article 

    Google Scholar 
    Fishelson, L. Comparative internal morphology of deep-sea eels, with particular emphasis on gonads and gut structure. J. Fish. Biol. 44, 75–101 (1994).
    Google Scholar 
    Bailey, D. M. et al. High swimming and metabolic activity in the deep-sea eel Synaphobranchus kaupii revealed by integrated in situ and in vitro measurements. Physiol. Biochem. Zool. 78, 335–346 (2005).
    Google Scholar 
    Trenkel, V. M. & Lorance, P. Estimating Synaphobranchus kaupii densities: contribution of fish behaviour to differences between bait experiments and visual strip transects. Deep Sea Res. Oceanogr. Res. Part I Pap. 58, 63–71 (2011).ADS 

    Google Scholar 
    Raupach, M. J. et al. Genetic homogeneity and circum-Antarctic distribution of two benthic shrimp species of the Southern Ocean, Chorismus antarcticus and Nematocarcinus lanceopes. Mar. Biol. 157, 1783–1797 (2010).CAS 

    Google Scholar 
    Dambach, J., Raupach, M. J., Leese, F., Schwarzer, J. & Engler, J. O. Ocean currents determine functional connectivity in an Antarctic deep-sea shrimp. Mar. Ecol. 37, 1336–1344 (2016).ADS 
    CAS 

    Google Scholar 
    Dambach, J., Raupach, M. J., Mayer, C., Schwarzer, J. & Leese, F. Isolation and characterization of nine polymorphic microsatellite markers for the deep-sea shrimp Nematocarcinus lanceopes (Crustacea: Decapoda: Caridea). BMC Res. Notes 6, 75. https://doi.org/10.1186/1756-0500-6-75 (2013).Article 

    Google Scholar 
    Ritchie, H., Jamieson, A. J. & Piertney, S. B. Phylogenetic relationships among hadal amphipods of the Superfamily Lysianassoidea: Implications for taxonomy and biogeography. Deep Sea Res. Part I 105, 119–131 (2015).CAS 

    Google Scholar 
    Bowen, B. W. et al. Phylogeography unplugged: comparative surveys in the genomic era. Bull. Mar. Sci. 90, 13–46 (2014).
    Google Scholar 
    Ritchie, H., Jamieson, A. J. & Piertney, S. B. Population genetic structure of two congeneric deep-sea amphipod species from geographically isolated hadal trenches in the Pacific Ocean. Deep Sea Res. Part I. 119, 50–57 (2017).
    Google Scholar 
    Iguchi, A. et al. Deep-sea amphipods around cobalt-rich ferromanganese crusts: taxonomic diversity and selection of candidate species for connectivity analysis. PLoS ONE 15, e0228483. https://doi.org/10.1371/journal.pone.0228483 (2020).Article 
    CAS 

    Google Scholar 
    Baco, A. R. et al. A synthesis of genetic connectivity in deep-sea fauna and implications for marine reserve design. Mol. Ecol. 25, 3276–3298 (2016).
    Google Scholar 
    Taylor, M. L. & Roterman, C. N. Invertebrate population genetics across Earth’s largest habitat: the deep-sea floor. Mol. Ecol. 26, 4872–4896 (2017).CAS 

    Google Scholar  More

  • in

    Intra-individual variation of hen movements is associated with later keel bone fractures in a quasi-commercial aviary

    Rufener, C. et al. Keel bone fractures are associated with individual mobility of laying hens in an aviary system. Appl. Anim. Behav. Sci. 217, 48–56 (2019).
    Google Scholar 
    Rentsch, A. K., Rufener, C. B., Spadavecchia, C., Stratmann, A. & Toscano, M. J. Laying hen’s mobility is impaired by keel bone fractures and does not improve with paracetamol treatment. Appl. Anim. Behav. Sci. 216, 19–25 (2019).
    Google Scholar 
    Rodriguez-Aurrekoetxea, A. & Estevez, I. Use of space and its impact on the welfare of laying hens in a commercial free-range system. Poult. Sci. 95, 2503–2513 (2016).CAS 

    Google Scholar 
    Fagan, W. F. et al. Spatial memory and animal movement. Ecol. Lett. 16, 1316–1329 (2013).
    Google Scholar 
    Campbell, D. L. M., Talk, A. C., Loh, Z. A., Dyall, T. R. & Lee, C. Spatial cognition and range use in free-range laying hens. Animals 8, 26 (2018).
    Google Scholar 
    de Jager, M., Weissing, F. J., Herman, P. M. J., Nolet, B. A. & van de Koppel, J. Lévy walks evolve through interaction between movement and environmental complexity. Science 1979(332), 1551–1553 (2011).
    Google Scholar 
    Krause, J., James, R. & Croft, D. P. Personality in the context of social networks. Philos. Trans. R. Soc. B Biol. Sci. 365, 4099–4106 (2010).CAS 

    Google Scholar 
    Ihwagi, F. W. et al. Poaching lowers elephant path tortuosity: Implications for conservation. J. Wildl. Manag. 83, 1022–1031 (2019).
    Google Scholar 
    Shaw, A. K. Causes and consequences of individual variation in animal movement. Mov. Ecol. 8, 1–12 (2020).
    Google Scholar 
    Matthews, S. G., Miller, A. L., Plötz, T. & Kyriazakis, I. Automated tracking to measure behavioural changes in pigs for health and welfare monitoring. Sci. Rep. 7, 1–12 (2017).CAS 

    Google Scholar 
    Berger-Tal, O. & Saltz, D. Using the movement patterns of reintroduced animals to improve reintroduction success. Curr. Zool. 60, 515–526 (2014).
    Google Scholar 
    Stuber, E. F., Carlson, B. S. & Jesmer, B. R. Spatial personalities: A meta-analysis of consistent individual differences in spatial behavior. Behav. Ecol. https://doi.org/10.1093/BEHECO/ARAB147 (2022).Article 

    Google Scholar 
    Sirovnik, J., Würbel, H. & Toscano, M. J. Feeder space affects access to the feeder, aggression, and feed conversion in laying hens in an aviary system. Appl. Anim. Behav. Sci. 198, 75–82 (2018).
    Google Scholar 
    Sirovnik, J., Voelkl, B., Keeling, L. J., Würbel, H. & Toscano, M. J. Breakdown of the ideal free distribution under conditions of severe and low competition. Behav. Ecol. Sociobiol. 75, 1–11 (2021).
    Google Scholar 
    Becot, L., Bedere, N., Burlot, T., Coton, J. & le Roy, P. Nest acceptance, clutch, and oviposition traits are promising selection criteria to improve egg production in cage-free system. PLoS ONE 16, e0251037 (2021).CAS 

    Google Scholar 
    Thompson, M. J., Evans, J. C., Parsons, S. & Morand-Ferron, J. Urbanization and individual differences in exploration and plasticity. Behav. Ecol. 29, 1415–1425 (2018).
    Google Scholar 
    Stamps, J. & Groothuis, T. G. G. The development of animal personality: Relevance, concepts and perspectives. Biol. Rev. 85, 301–325 (2010).
    Google Scholar 
    Salinas-Melgoza, A., Salinas-Melgoza, V. & Wright, T. F. Behavioral plasticity of a threatened parrot in human-modified landscapes. Biol. Conserv. 159, 303–312 (2013).
    Google Scholar 
    Stamps, J. A., Briffa, M. & Biro, P. A. Unpredictable animals: Individual differences in intraindividual variability (IIV). Anim. Behav. 83, 1325–1334 (2012).
    Google Scholar 
    Hertel, A. G., Royauté, R., Zedrosser, A. & Mueller, T. Biologging reveals individual variation in behavioural predictability in the wild. J. Anim. Ecol. 90, 723–737 (2021).
    Google Scholar 
    Biro, P. A. & Adriaenssens, B. Predictability as a personality trait: Consistent differences in intraindividual behavioral variation. Am. Nat. 182, 621–629 (2013).
    Google Scholar 
    Henriksen, R. et al. Intra-individual behavioural variability: A trait under genetic control. Int. J. Mol. Sci. 21, 8069 (2020).CAS 

    Google Scholar 
    Rufener, C. et al. Finding hens in a haystack: Consistency of movement patterns within and across individual laying hens maintained in large groups. Sci. Rep. 8, (2018).Campbell, D. L. M., Karcher, D. M. & Siegford, J. M. Location tracking of individual laying hens housed in aviaries with different litter substrates. Appl. Anim. Behav. 184, 74–79 (2016).
    Google Scholar 
    Weeks, C. A. & Nicol, C. J. Behavioural needs, priorities and preferences of laying hens. Worlds Poult. Sci. J. 62, 296–307 (2006).
    Google Scholar 
    Hartcher, K. M. & Jones, B. The welfare of layer hens in cage and cage-free housing systems. Worlds Poult. Sci. J. 73, 767–782 (2017).
    Google Scholar 
    Zeltner, E. & Hirt, H. Effect of artificial structuring on the use of laying hen runs in a free-range system. Br. Poult. Sci. 44, 533–537 (2010).
    Google Scholar 
    Stratmann, A. et al. Modification of aviary design reduces incidence of falls, collisions and keel bone damage in laying hens. Appl. Anim. Behav. Sci. 165, 112–123 (2015).
    Google Scholar 
    Vandekerchove, D., Herdt, P., Laevens, H. & Pasmans, F. Colibacillosis in caged layer hens: Characteristics of the disease and the aetiological agent. Avian Pathol. 33, 117–125 (2004).CAS 

    Google Scholar 
    Montalcini, C. M., Voelkl, B., Gómez, Y., Gantner, M. & Toscano, M. J. Evaluation of an active LF tracking system and data processing methods for livestock precision farming in the poultry sector. Sensors 22, 659 (2022).ADS 

    Google Scholar 
    Revelle, W. Procedures for psychological, psychometric, and personality research. (2021).Kaiser, H. F. The application of electronic computers to factor analysis. Educ. Psychol. Meas. 20, 141–151 (1960).
    Google Scholar 
    Rufener, C., Baur, S., Stratmann, A. & Toscano, M. J. A reliable method to assess keel bone fractures in laying hens from radiographs using a tagged visual analogue scale. Front. Vet. Sci. 5, 124 (2018).
    Google Scholar 
    Tauson, R., Kjaer, J., Maria, G. A., Cepero, R. & Holm, K.-E. The creation of a common scoring system for the integument and health of laying hens: Applied scoring of integument and health in laying hens. Final report Health from the Laywell project. https://www.laywel.eu/web/pdf/deliverables%2031-33%20health.pdf (2005).Hertel, A. G. et al. A guide for studying among-individual behavioral variation from movement data in the wild. Mov. Ecol. 8, (2020).Nakagawa, S. & Schielzeth, H. Repeatability for Gaussian and non-Gaussian data: A practical guide for biologists. Biol. Rev. 85, 935–956 (2010).
    Google Scholar 
    Dingemanse, N. J., Kazem, A. J. N., Réale, D. & Wright, J. Behavioural reaction norms: Animal personality meets individual plasticity. Trends Ecol. Evol. 25, 81–89 (2010).
    Google Scholar 
    Bates, D., Mächler, M., Bolker, B. M. & Walker, S. C. Fitting linear mixed-effects models using lme4. J Stat Softw 67, (2015).Cleasby, I. R., Nakagawa, S. & Schielzeth, H. Quantifying the predictability of behaviour: Statistical approaches for the study of between-individual variation in the within-individual variance. Methods Ecol. Evol. 6, 27–37 (2015).
    Google Scholar 
    Bürkner, P.-C. brms: An R package for bayesian multilevel models using Stan. J. Stat. Softw. 80, 1–28 (2017).
    Google Scholar 
    Vehtari, A., Gelman, A. & Gabry, J. Practical Bayesian model evaluation using leave-one-out cross-validation and WAIC. Stat. Comput. 27, 1413–1432 (2017).MathSciNet 
    MATH 

    Google Scholar 
    Gelman, A. & Rubin, D. B. Inference from iterative simulation using multiple sequences. Stat. Sci. 7, 457–472 (1992).MATH 

    Google Scholar 
    Hadfield, J. D. MCMC methods for multi-response generalized linear mixed models: The MCMCglmm R package. J. Stat. Softw. 33, 1–22 (2010).
    Google Scholar 
    Houslay, T. M. & Wilson, A. J. Avoiding the misuse of BLUP in behavioural ecology. Behav. Ecol. 28, 948–952 (2017).
    Google Scholar 
    Hertel, A. G., Niemelä, P. T., Dingemanse, N. J. & Mueller, T. Don’t poke the bear: Using tracking data to quantify behavioural syndromes in elusive wildlife. Anim. Behav. 147, 91–104 (2019).
    Google Scholar 
    Spiegel, O., Leu, S. T., Bull, C. M. & Sih, A. What’s your move? Movement as a link between personality and spatial dynamics in animal populations. Ecol. Lett. 20, 3–18 (2017).ADS 

    Google Scholar 
    Bell, A. M., Hankison, S. J. & Laskowski, K. L. The repeatability of behaviour: A meta-analysis. Anim. Behav. 77, 771–783 (2009).
    Google Scholar 
    Occhiuto, F., Vázquez-Diosdado, J. A., Carslake, C. & Kaler, J. Personality and predictability in farmed calves using movement and space-use behaviours quantified by ultra-wideband sensors. R. Soc. Open Sci. 9, (2022).Moinard, C. et al. Accuracy of laying hens in jumping upwards and downwards between perches in different light environments. Appl. Anim. Behav. Sci. 85, 77–92 (2004).
    Google Scholar 
    Baur, S., Rufener, C., Toscano, M. J. & Geissbühler, U. Radiographic evaluation of keel bone damage in laying hens—Morphologic and temporal observations in a longitudinal study. Front. Vet. Sci. 1, 129 (2020).
    Google Scholar 
    Cordiner, L. S. & Savory, C. J. Use of perches and nestboxes by laying hens in relation to social status, based on examination of consistency of ranking orders and frequency of interaction. Appl. Anim. Behav. Sci. 71, 305–317 (2001).
    Google Scholar 
    Rufener, C. & Makagon, M. M. Keel bone fractures in laying hens: A systematic review of prevalence across age, housing systems, and strains. J. Anim. Sci. 98, S36–S51 (2020).
    Google Scholar 
    Nasr, M. A. F., Nicol, C. J., Wilkins, L. & Murrell, J. C. The effects of two non-steroidal anti-inflammatory drugs on the mobility of laying hens with keel bone fractures. Vet. Anaesth. Analg. 42, 197–204 (2015).CAS 

    Google Scholar 
    Nasr, M., Murrell, J., Wilkins, L. J. & Nicol, C. J. The effect of keel fractures on egg-production parameters, mobility and behaviour in individual laying hens. Anim. Welf. 21, 127–135 (2012).CAS 

    Google Scholar 
    Koolhaas, J. M. & van Reenen, C. G. Animal behavior and well-being symposium: Interaction between coping style/personality, stress, and welfare: Relevance for domestic farm animals. J. Anim. Sci. 94, 2284–2296 (2016).CAS 

    Google Scholar 
    Coppens, C. M., de Boer, S. F. & Koolhaas, J. M. Coping styles and behavioural flexibility: Towards underlying mechanisms. Philos. Trans. R. Soc. B Biol. Sci. 365, 4021 (2010).
    Google Scholar 
    Koolhaas, J. M., de Boer, S. F., Coppens, C. M. & Buwalda, B. Neuroendocrinology of coping styles: Towards understanding the biology of individual variation. Front. Neuroendocrinol. 31, 307–321 (2010).CAS 

    Google Scholar 
    Finkemeier, M.-A., Langbein, J. & Puppe, B. Personality research in mammalian farm animals: Concepts, measures, and relationship to welfare. Front. Vet. Sci. 5, 131 (2018).
    Google Scholar 
    Martin, J. G. A., Pirotta, E., Petelle, M. B. & Blumstein, D. T. Genetic basis of between-individual and within-individual variance of docility. J. Evol. Biol. 30, 796–805 (2017).CAS 

    Google Scholar 
    Prentice, P. M., Houslay, T. M., Martin, J. G. A. & Wilson, A. J. Genetic variance for behavioural ‘predictability’ of stress response. J. Evol. Biol. 33, 642–652 (2020).
    Google Scholar  More

  • in

    Chemotaxis increases metabolic exchanges between marine picophytoplankton and heterotrophic bacteria

    Aylward, F. O. et al. Microbial community transcriptional networks are conserved in three domains at ocean basin scales. Proc. Natl Acad. Sci. USA 112, 5443–5448 (2015).Article 
    CAS 

    Google Scholar 
    Fuhrman, J. A. Microbial community structure and its functional implications. Nature 459, 193–199 (2009).Article 
    CAS 

    Google Scholar 
    Amin, S. A., Parker, M. S. & Armbrust, E. V. Interactions between diatoms and bacteria. Microbiol. Mol. Biol. Rev. 76, 667–684 (2012).Article 
    CAS 

    Google Scholar 
    Mayali, X. Metabolic interactions between bacteria and phytoplankton. Front. Microbiol. 9, 727 (2018).Article 

    Google Scholar 
    Amin, S. A. et al. Photolysis of iron–siderophore chelates promotes bacterial–algal mutualism. Proc. Natl Acad. Sci. USA 106, 17071–17076 (2009).Amin, S. A. et al. Interaction and signalling between a cosmopolitan phytoplankton and associated bacteria. Nature 522, 98 (2015).Article 
    CAS 

    Google Scholar 
    Durham, B. P. et al. Cryptic carbon and sulfur cycling between surface ocean plankton. Proc. Natl Acad. Sci. USA 112, 453 (2015).Article 
    CAS 

    Google Scholar 
    Stocker, R. Marine microbes see a sea of gradients. Science 338, 628 (2012).Article 
    CAS 

    Google Scholar 
    Bell, W. & Mitchell, R. Chemotactic and growth responses of marine bacteria to algal extracellular products. Biol. Bull. 143, 265–277 (1972).Article 

    Google Scholar 
    Azam, F. & Ammerman, J. W. in Flows of Energy and Materials in Marine Ecosystems 345–360 (Springer, 1984).Mitchell, J. G., Okubo, A. & Fuhrman, J. A. Microzones surrounding phytoplankton form the basis for a stratified marine microbial ecosystem. Nature 316, 58–59 (1985).Article 
    CAS 

    Google Scholar 
    Seymour, J. R., Amin, S. A., Raina, J.-B. & Stocker, R. Zooming in on the phycosphere: the ecological interface for phytoplankton–bacteria relationships. Nat. Microbiol. 2, 17065 (2017).Article 
    CAS 

    Google Scholar 
    Sonnenschein, E. C., Syit, D. A., Grossart, H.-P. & Ullrich, M. S. Chemotaxis of Marinobacter adhaerens and its impact on attachment to the diatom Thalassiosira weissflogii. Appl. Environ. Microbiol. 78, 6900–6907 (2012).Article 
    CAS 

    Google Scholar 
    Raina, J.-B., Fernandez, V., Lambert, B., Stocker, R. & Seymour, J. R. The role of microbial motility and chemotaxis in symbiosis. Nat. Rev. Microbiol. 17, 284–294 (2019).Article 
    CAS 

    Google Scholar 
    Seymour, J. R., Ahmed, T., Durham, W. M. & Stocker, R. Chemotactic response of marine bacteria to the extracellular products of Synechococcus and Prochlorococcus. Aquat. Microb. Ecol. 59, 161–168 (2010).Article 

    Google Scholar 
    Smriga, S., Fernandez, V. I., Mitchell, J. G. & Stocker, R. Chemotaxis toward phytoplankton drives organic matter partitioning among marine bacteria. Proc. Natl Acad. Sci. USA 113, 1576–1581 (2016).Article 
    CAS 

    Google Scholar 
    Flombaum, P., Wang, W.-L., Primeau, F. W. & Martiny, A. C. Global picophytoplankton niche partitioning predicts overall positive response to ocean warming. Nat. Geosci. 13, 116–120 (2020).Article 
    CAS 

    Google Scholar 
    Christie-Oleza, J. A., Sousoni, D., Lloyd, M., Armengaud, J. & Scanlan, D. J. Nutrient recycling facilitates long-term stability of marine microbial phototroph–heterotroph interactions. Nat. Microbiol. 2, 17100 (2017).Article 
    CAS 

    Google Scholar 
    Morris, J. J., Kirkegaard, R., Szul, M. J., Johnson, Z. I. & Zinser, E. R. Facilitation of robust growth of Prochlorococcus colonies and dilute liquid cultures by ‘helper’ heterotrophic bacteria. Appl. Environ. Microbiol. 74, 4530–4534 (2008).Article 
    CAS 

    Google Scholar 
    Sher, D., Thompson, J. W., Kashtan, N., Croal, L. & Chisholm, S. W. Response of Prochlorococcus ecotypes to co-culture with diverse marine bacteria. ISME J. 5, 1125–1132 (2011).Article 
    CAS 

    Google Scholar 
    Aharonovich, D. & Sher, D. Transcriptional response of Prochlorococcus to co-culture with a marine Alteromonas: differences between strains and the involvement of putative infochemicals. ISME J. 10, 2892–2906 (2016).Article 
    CAS 

    Google Scholar 
    Jackson, G. A. Simulating chemosensory responses of marine microorganisms. Limnol. Oceanogr. 32, 1253–1266 (1987).Article 
    CAS 

    Google Scholar 
    Gärdes, A., Iversen, M. H., Grossart, H.-P., Passow, U. & Ullrich, M. S. Diatom-associated bacteria are required for aggregation of Thalassiosira weissflogii. ISME J. 5, 436–445 (2011).Article 

    Google Scholar 
    Al-Wahaib, D., Al-Bader, D., Al-Shaikh Abdou, D. K., Eliyas, M. & Radwan, S. S. Consistent occurrence of hydrocarbonoclastic Marinobacter strains in various cultures of picocyanobacteria from the Arabian Gulf: promising associations for biodegradation of marine oil pollution. J. Mol. Microbiol. Biotechnol. 26, 261–268 (2016).CAS 

    Google Scholar 
    Raina, J.-B. et al. Subcellular tracking reveals the location of dimethylsulfoniopropionate in microalgae and visualises its uptake by marine bacteria. eLife 6, e23008 (2017).Article 

    Google Scholar 
    Brumley, D. R. et al. Cutting through the noise: bacterial chemotaxis in marine microenvironments. Front. Mar. Sci. 7, 527 (2020).Article 

    Google Scholar 
    Gärdes, A. et al. Complete genome sequence of Marinobacter adhaerens type strain (HP15), a diatom-interacting marine microorganism. Stand. Genom. Sci. 3, 97–107 (2010).Article 

    Google Scholar 
    Moore, L. R., Post, A. F., Rocap, G. & Chisholm, S. W. Utilization of different nitrogen sources by the marine cyanobacteria Prochlorococcus and Synechococcus. Limnol. Oceanogr. 47, 989–996 (2002).Article 
    CAS 

    Google Scholar 
    Wawrik, B., Callaghan, A. V. & Bronk, D. A. Use of inorganic and organic nitrogen by Synechococcus spp. and diatoms on the West Florida shelf as measured using stable isotope probing. Appl. Environ. Microbiol. 75, 6662–6670 (2009).Article 
    CAS 

    Google Scholar 
    Lambert, B. S. et al. A microfluidics-based in situ chemotaxis assay to study the behaviour of aquatic microbial communities. Nat. Microbiol. 2, 1344–1349 (2017).Article 
    CAS 

    Google Scholar 
    Raina, J.-B. et al. Chemotaxis shapes the microscale organization of the ocean’s microbiome. Nature 605, 132–138 (2022).Article 
    CAS 

    Google Scholar 
    Brumley, D. R. et al. Bacteria push the limits of chemotactic precision to navigate dynamic chemical gradients. Proc. Natl Acad. Sci. USA 116, 10792–10797 (2019).Article 
    CAS 

    Google Scholar 
    Myklestad, S. M. in Marine Chemistry (ed. Wangersky, P. J.) 111–148 (Springer Berlin Heidelberg, 2000).Ni, B., Colin, R., Link, H., Endres, R. G. & Sourjik, V. Growth-rate dependent resource investment in bacterial motile behavior quantitatively follows potential benefit of chemotaxis. Proc. Natl Acad. Sci. USA 117, 595–601 (2020).Article 
    CAS 

    Google Scholar 
    Stocker, R., Seymour, J. R., Samadani, A., Hunt, D. E. & Polz, M. F. Rapid chemotactic response enables marine bacteria to exploit ephemeral microscale nutrient patches. Proc. Natl Acad. Sci. USA 105, 4209–4214 (2008).Article 
    CAS 

    Google Scholar 
    Buitenhuis, E. et al. MAREDAT: towards a world atlas of MARine Ecosystem DATa. Earth Syst. Sci. Data 5, 227–239 (2013).Article 

    Google Scholar 
    Raina, J.-B. et al. Symbiosis in the microbial world: from ecology to genome evolution. Biol. Open 7, bio032524 (2018).Article 

    Google Scholar 
    Giardina, M. et al. Quantifying inorganic nitrogen assimilation by Synechococcus using bulk and single-cell mass spectrometry: a comparative study. Front. Microbiol. 9, 2847 (2018).Article 

    Google Scholar 
    Berges, J. A., Franklin, D. J. & Harrison, P. J. Evolution of an artificial seawater medium: improvements in enriched seawater, artificial water over the last two decades. J. Phycol. 37, 1138–1145 (2001).Article 

    Google Scholar 
    Guillard, R. R. L. in Culture of Marine Invertebrate Animals: Proceedings—1st Conference on Culture of Marine Invertebrate Animals Greenport (eds Walter, L. S. & Matoira, H. C.) 29–60 (Springer US, 1975).Kaeppel, E. C., Gärdes, A., Seebah, S., Grossart, H.-P. & Ullrich, M. S. Marinobacter adhaerens sp. nov., isolated from marine aggregates formed with the diatom Thalassiosira weissflogii. Int. J. Syst. Evolut. Microbiol. 62, 124–128 (2012).Article 
    CAS 

    Google Scholar 
    Sonnenschein, E. C. et al. Development of a genetic system for Marinobacter adhaerens HP15 involved in marine aggregate formation by interacting with diatom cells. J. Microbiol. Methods 87, 176–183 (2011).Article 
    CAS 

    Google Scholar 
    Marie, D., Partensky, F., Jacquet, S. & Vaulot, D. Enumeration and cell cycle analysis of natural populations of marine picoplankton by flow cytometry using the nucleic acid stain SYBR Green I. Appl. Environ. Microbiol. 63, 186–193 (1997).Article 
    CAS 

    Google Scholar 
    Schindelin, J. et al. Fiji: an open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).Article 
    CAS 

    Google Scholar 
    Hillion, F., Kilburn, M., Hoppe, P., Messenger, S. & Weber, P. K. The effect of QSA on S, C, O and Si isotopic ratio measurements. Geochim. Cosmochim. Acta 72, A377 (2008).
    Google Scholar 
    Popa, R. et al. Carbon and nitrogen fixation and metabolite exchange in and between individual cells of Anabaena oscillarioides. ISME J. 1, 354–360 (2007).Article 
    CAS 

    Google Scholar 
    Sumner, L. W. et al. Proposed minimum reporting standards for chemical analysis. Metabolomics 3, 211–221 (2007).Article 
    CAS 

    Google Scholar 
    Clerc, E. E., Raina, J.-B., Lambert, B. S., Seymour, J. & Stocker, R. In situ chemotaxis assay to examine microbial behavior in aquatic ecosystems. JoVE https://doi.org/10.3791/61062 (2020).Ihaka, R. & Gentleman, R. R: a language for data analysis and graphics. J. Comput. Graph. Stat. 5, 299–314 (1996).
    Google Scholar 
    Xie, L., Lu, C. & Wu, X.-L. Marine bacterial chemoresponse to a stepwise chemoattractant stimulus. Biophys. J. 108, 766–774 (2015).Article 
    CAS 

    Google Scholar 
    Son, K., Guasto, J. S. & Stocker, R. Bacteria can exploit a flagellar buckling instability to change direction. Nat. Phys. 9, 494–498 (2013).Article 
    CAS 

    Google Scholar 
    Lee, C. & Bada, J. L. Amino acids in equatorial Pacific Ocean water. Earth Planet. Sci. Lett. 26, 61–68 (1975).Article 
    CAS 

    Google Scholar 
    Yamashita, Y. & Tanoue, E. Distribution and alteration of amino acids in bulk DOM along a transect from bay to oceanic waters. Mar. Chem. 82, 145–160 (2003).Article 
    CAS 

    Google Scholar 
    Menden-Deuer, S. & Lessard, E. J. Carbon to volume relationships for dinoflagellates, diatoms, and other protist plankton. Limnol. Oceanogr. 45, 569–579 (2000).Article 
    CAS 

    Google Scholar 
    Mullin, M. M., Sloan, P. R. & Eppley, R. W. Relationship between carbon content, cell volume and area in phytoplankton. Limnol. Oceanogr. 11, 307–311 (1966).Article 

    Google Scholar  More

  • in

    Atmospheric–ocean coupling drives prevailing and synchronic dispersal patterns of marine species with long pelagic durations

    Guichard, F., Levin, S. A., Hastings, A. & Siegel, D. Toward a dynamic metacommunity approach to marine reserve theory. BioScience 54(11), 1003. https://doi.org/10.1641/0006-3568(2004)054[1003:tadmat]2.0.co;2 (2004).Article 

    Google Scholar 
    Wieters, E. A., Gaines, S. D., Navarrete, S. A., Blanchette, C. A. & Menge, B. A. Scales of dispersal and the biogeography of marine predator-prey interactions. Am. Nat. 171(3), 405–417. https://doi.org/10.1086/527492 (2008).Article 

    Google Scholar 
    Martínez-Moreno, J. et al. Global changes in oceanic mesoscale currents over the satellite altimetry record. Nat. Clim. Changehttps://doi.org/10.1038/s41558-021-01006-9 (2021).Article 

    Google Scholar 
    van Gennip, S. J. et al. Going with the flow: The role of ocean circulation in global marine ecosystems under a changing climate. Glob. Change Biol. 23(7), 2602–2617. https://doi.org/10.1111/gcb.13586 (2017).Article 
    ADS 

    Google Scholar 
    O’Connor, M. I. et al. Temperature control of larval dispersal and the implications for marine ecology, evolution, and conservation. Proc. Natl. Acad. Sci. U.S.A. 104(4), 1266–1271. https://doi.org/10.1073/pnas.0603422104 (2007).Article 
    ADS 
    CAS 

    Google Scholar 
    Cowen, R. K. & Sponaugle, S. Larval dispersal and marine population connectivity. Ann. Rev. Mar. Sci. 1(1), 443–466. https://doi.org/10.1146/annurev.marine.010908.163757 (2009).Article 

    Google Scholar 
    Ospina-Alvarez, A., Parada, C. & Palomera, I. Vertical migration effects on the dispersion and recruitment of European anchovy larvae: From spawning to nursery areas. Ecol. Model. 231, 65–79. https://doi.org/10.1016/j.ecolmodel.2012.02.001 (2012).Article 

    Google Scholar 
    Selkoe, K. A. & Toonen, R. J. Marine connectivity: A new look at pelagic larval duration and genetic metrics of dispersal. Mar. Ecol. Prog. Ser. 436, 291–305. https://doi.org/10.3354/meps09238 (2011).Article 
    ADS 

    Google Scholar 
    Siegel, D. A. et al. The stochastic nature of larval connectivity among nearshore marine populations. Proc. Natl. Acad. Sci. U.S.A. 105(26), 8974–8979. https://doi.org/10.1073/pnas.0802544105 (2008).Article 
    ADS 

    Google Scholar 
    De Lestang, S. et al. What caused seven consecutive years of low puerulus settlement in the western rock lobster fishery of Western Australia?. ICES J. Mar. Sci. 72, i49–i58. https://doi.org/10.1093/icesjms/fsu177 (2015).Article 

    Google Scholar 
    Linnane, A. et al. Evidence of large-scale spatial declines in recruitment patterns of southern rock lobster Jasus edwardsii, across south-eastern Australia. Fish. Res. 105(3), 163–171. https://doi.org/10.1016/j.fishres.2010.04.001 (2010).Article 

    Google Scholar 
    Briones-Fourzán, P., Candela, J. & Lozano-Álvarez, E. Postlarval settlement of the spiny lobster Panulirus argus along the Caribbean coast of Mexico: Patterns, influence of physical factors, and possible sources of origin. Limnol. Oceanogr. 53(3), 970–985. https://doi.org/10.4319/lo.2008.53.3.0970 (2008).Article 
    ADS 

    Google Scholar 
    Haury, L. R., McGowan, J. A. & Wiebe, P. H. Patterns and processes in the time-space scales of plankton distributions. In Spatial Pattern in Plankton Communities (ed. Steele, J. H.) 277–327 (Springer US, 1978). https://doi.org/10.1007/978-1-4899-2195-6_12.Cowen, R. K., Paris, C. B. & Srinivasan, A. Scaling of connectivity in marine populations. Science 311(5760), 522–527. https://doi.org/10.1126/science.1122039 (2006).Article 
    ADS 
    CAS 

    Google Scholar 
    Kavanaugh, M. T. et al. Seascapes as a new vernacular for pelagic ocean monitoring, management and conservation. ICES J. Mar. Sci. 73(7), 1839–1850. https://doi.org/10.1093/icesjms/fsw086 (2016).Article 

    Google Scholar 
    Ospina-Alvarez, A., Weidberg, N., Aiken, C. M. & Navarrete, S. A. Larval transport in the upwelling ecosystem of central Chile: The effects of vertical migration, developmental time and coastal topography on recruitment. Prog. Oceanogr. 168, 82–99. https://doi.org/10.1016/j.pocean.2018.09.016 (2018) http://www.sciencedirect.com/science/article/pii/S0079661117300800.Article 
    ADS 

    Google Scholar 
    Palumbi, S. Population genetics, demographic connectivity, and the design of marine reserves. Ecol. Appl. 13(1 Supplement), S146–S158 (2003).Article 

    Google Scholar 
    Barahona, M. et al. Environmental and demographic factors influence the spatial genetic structure of an intertidal barnacle in central-northern Chile. Mar. Ecol. Prog. Ser. 612, 151–165. https://doi.org/10.3354/meps12855 (2019) http://www.int-res.com/abstracts/meps/v612/p151-165/.Article 
    ADS 

    Google Scholar 
    Spanier, E. et al. A concise review of lobster utilization by worldwide human populations from prehistory to the modern era. ICES J. Mar. Sci. 72(May), i7–i21. https://doi.org/10.1093/icesjms/fsv066 (2015).Article 

    Google Scholar 
    IUCN. Palinurus elephas: Goñi, R.: The IUCN Red List of Threatened Species 2014: e.T169975A1281221. Tech. Rep., International Union for Conservation of Nature (2013). http://www.iucnredlist.org/details/169975/0. Type: dataset.Canepa, A. et al. Pelagia noctiluca in the mediterranean sea (eds Pitt, K. A. & Lucas, C. H.) In Jellyfish Blooms, Vol. 9789400770 237–266 (Springer Netherlands, 2014). https://doi.org/10.1007/978-94-007-7015-7_11.Bosch-Belmar, M. et al. Jellyfish blooms perception in Mediterranean finfish aquaculture. Mar. Policy 76, 1–7. https://doi.org/10.1016/j.marpol.2016.11.005 (2017).Article 

    Google Scholar 
    Exceltur. Impactur baleares 2014. Tech. Rep., EXCELTUR – Govern de les Illes Balears, Madrid (2014).Vignudelli, S., Gasparini, G. P., Astraldi, M. & Schiano, M. E. A possible influence of the North Atlantic Oscillation on the circulation of the Western Mediterranean Sea. Geophys. Res. Lett. 26(5), 623–626. https://doi.org/10.1029/1999GL900038 (1999).Article 
    ADS 

    Google Scholar 
    Somot, S. et al. Characterizing, modelling and understanding the climate variability of the deep water formation in the North-Western Mediterranean Sea. Clim. Dyn. 51(3), 1179–1210. https://doi.org/10.1007/s00382-016-3295-0 (2018).Article 

    Google Scholar 
    Díaz, D., Marí, M., Abelló, P. & Demestre, M. Settlement and juvenile habitat of the European spiny lobster Palinurus elephas (Crustacea: Decapoda: Palinuridae) in the Western Mediterranean Sea. Sci. Mar. 65(4), 347–356. https://doi.org/10.3989/scimar.2001.65n4347 (2001).Article 

    Google Scholar 
    Muñoz, A. et al. Exploration of the inter-annual variability and multi-scale environmental drivers of European spiny lobster, Palinurus elephas (Decapoda: Palinuridae) settlement in the NW Mediterranean. Mar. Ecol.https://doi.org/10.1111/maec.12654 (2021).Article 

    Google Scholar 
    Malej, A. & Malej, M. Population dynamics of the jellyfish Pelagia noctiluca (Forsskal, 1775) In Marine Eutrophication and Population Dynamics (eds Colombo, G., Ferrari, I., V., C. & R., R.) 215–219 (Olsen and Olsen, 1992).Ottmann, D. et al. Abundance of Pelagia noctiluca early life stages in the western Mediterranean Sea scales with surface chlorophyll. Mar. Ecol. Prog. Ser. 658, 75–88. https://doi.org/10.3354/meps13423 (2021).Article 
    ADS 
    CAS 

    Google Scholar 
    Benedetti-Cecchi, L. et al. Deterministic factors overwhelm stochastic environmental fluctuations as drivers of jellyfish outbreaks. PLoS One 10(10), 1–16. https://doi.org/10.1371/journal.pone.0141060 (2015).Article 
    CAS 

    Google Scholar 
    Licandro, P. et al. A blooming jellyfish in the northeast Atlantic and Mediterranean. Biol. Lett. 6(5), 688–691. https://doi.org/10.1098/rsbl.2010.0150 (2010).Article 
    CAS 

    Google Scholar 
    Goy, J., Morand, P. & Etienne, M. Long-term fluctuations of Pelagia noctiluca (Cnidaria, Scyphomedusa) in the western Mediterranean Sea. Prediction by climatic variables. Deep Sea Res. Part A Oceanogr. Res. Pap. 36(2), 269–279 (1989). https://doi.org/10.1016/0198-0149(89)90138-6 .Yahia, M. N. D. et al. Are the outbreaks timing of Pelagia noctiluca (Forsskal, 1775) getting more frequent in the Mediterranean basin?. ICES Cooper. Res. Rep. 300, 8–14 (2010).
    Google Scholar 
    Ferraris, M. et al. Distribution of Pelagia noctiluca (Cnidaria, Scyphozoa) in the Ligurian Sea (NW Mediterranean Sea). J. Plankton Res. 34(10), 874–885. https://doi.org/10.1093/plankt/fbs049 (2012).Article 

    Google Scholar 
    Millot, C. Circulation in the Western Mediterranean Sea. J. Mar. Syst. 20(1–4), 423–442. https://doi.org/10.1016/S0924-7963(98)00078-5 (1999).Article 

    Google Scholar 
    Galarza, J. A. et al. The influence of oceanographic fronts and early-life-history traits on connectivity among littoral fish species. Proc. Natl. Acad. Sci. 106(5), 1473–1478. https://doi.org/10.1073/pnas.0806804106 (2009).Article 
    ADS 

    Google Scholar 
    Fernández de Puelles, M. L. & Molinero, J. C. Decadal changes in hydrographic and ecological time-series in the Balearic Sea (western Mediterranean), identifying links between climate and zooplankton. ICES J. Mar. Sci. 65(3), 311–317. https://doi.org/10.1093/icesjms/fsn017 (2008).Article 

    Google Scholar 
    Arsouze, T. et al. CIESM (ed.) Sensibility analysis of the Western Mediterranean Transition inferred by four companion simulations. (ed. CIESM) EGU General Assembly Conference Abstracts, Vol. 1 of EGU General Assembly Conference Abstracts, 13073 (2013).Amores, A., Jordà, G., Arsouze, T. & Le Sommer, J. Up to what extent can we characterize ocean eddies using present-day gridded altimetric products?. J. Geophys. Res. Oceans 123(10), 7220–7236. https://doi.org/10.1029/2018JC014140 (2018).Article 
    ADS 

    Google Scholar 
    Waldman, R. et al. Impact of the mesoscale dynamics on ocean deep convection: The 2012–2013 case study in the northwestern mediterranean sea. J. Geophys. Res. Oceans 122(11), 8813–8840. https://doi.org/10.1002/2016JC012587 (2017).Article 
    ADS 

    Google Scholar 
    Lett, C. et al. A Lagrangian tool for modelling ichthyoplankton dynamics. Environ. Model. Softw. 23(9), 1210–1214. https://doi.org/10.1016/j.envsoft.2008.02.005 (2008).Article 

    Google Scholar 
    Brickman, D. & Smith, P. C. Lagrangian stochastic modeling in coastal oceanography. J. Atmos. Ocean. Technol. 19(1), 83–99. https://doi.org/10.1175/1520-0426(2002)0192.0.CO;2 (2002).Article 
    ADS 

    Google Scholar 
    Goñi, R. & Latrouite, D. Review of the biology, ecology and fisheries of Palinurus spp. species of European waters: Palinurus elephas (Fabricius, 1787) and Palinurus mauritanicus (Gruvel, 1911). Cahiers de Biol. Mar. 46(2), 127–142 (2005).
    Google Scholar 
    Bjornsson, H. & Venegas, S. A manual for EOF and SVD analyses of climatic data. Tech. Rep. CCGCR Report No. 97-1, McGill s Centre for Climate and Global Change Research (C2GCR) (1997).Herrmann, M., Somot, S., Sevault, F., Estournel, C. & Déqué, M. Modeling the deep convection in the northwestern mediterranean sea using an eddy-permitting and an eddy-resolving model: Case study of winter 1986–1987. J. Geophys. Res. Oceans 113(C4) (2008). https://doi.org/10.1029/2006JC003991.Hersbach, H. et al. ERA5 monthly averaged data on single levels from 1979 to present. Copernicus Climate Change Service (C3S) Climate Data Store (CDS). 10, 252–266 (2019). https://doi.org/10.24381/cds.f17050d7 .Bernard, P., Berline, L. & Gorsky, G. Long term (1981–2008) monitoring of the jellyfish Pelagia noctiluca (Cnidaria, Scyphozoa) on Mediterranean Coasts (Principality of Monaco and French Riviera). J. Oceanogr. Res. Data 4(1), 1–10 (2011).
    Google Scholar 
    Kough, A. S., Paris, C. B. & Butler, M. J. IV. Larval connectivity and the international management of fisheries. PLoS One 8(6), 1–12. https://doi.org/10.1371/journal.pone.0064970 (2013).Article 
    CAS 

    Google Scholar 
    Sandvik, H. et al. Modelled drift patterns of fish larvae link coastal morphology to seabird colony distribution. Nat. Commun. 7(May), 1–8. https://doi.org/10.1038/ncomms11599 (2016).Article 
    CAS 

    Google Scholar 
    Notarbartolo-Di-Sciara, G., Agardy, T., Hyrenbach, D., Scovazzi, T. & Van Klaveren, P. The Pelagos Sanctuary for Mediterranean marine mammals. Aquat. Conserv. Mar. Freshw. Ecosyst. 18(4), 367–391. https://doi.org/10.1002/aqc.855 (2008).Article 

    Google Scholar 
    Astraldi, M., Gasparini, G. P., Vetrano, a. & Vignudelli, S. Hydrographic characteristics and interannual variability of water masses in the central Mediterranean: A sensitivity test for long-term changes in the Mediterranean Sea. Deep Sea Res. Part I Oceanogr. Res. Pap. 49(4), 661–680 (2002). https://doi.org/10.1016/S0967-0637(01)00059-0 .Muffett, K. & Miglietta, M. P. Planktonic associations between medusae (classes Scyphozoa and Hydrozoa) and epifaunal crustaceans. PeerJ 9, e11281. https://doi.org/10.7717/peerj.11281 (2021) https://peerj.com/articles/11281.Article 

    Google Scholar 
    Stopar, K., Ramšak, A., Trontelj, P. & Malej, A. Lack of genetic structure in the jellyfish Pelagia noctiluca (Cnidaria: Scyphozoa: Semaeostomeae) across European seas. Mol. Phylogenet. Evol. 57(1), 417–428. https://doi.org/10.1016/j.ympev.2010.07.004 (2010).Article 
    CAS 

    Google Scholar 
    Berline, L., Zakardjian, B., Molcard, A., Ourmières, Y. & Guihou, K. Modeling jellyfish Pelagia noctiluca transport and stranding in the Ligurian Sea. Mar. Pollut. Bull. 70(1–2), 90–99. https://doi.org/10.1016/j.marpolbul.2013.02.016 (2013).Article 
    CAS 

    Google Scholar 
    Prieto, L., Macías, D., Peliz, A. & Ruiz, J. Portuguese Man-of-War (Physalia physalis) in the Mediterranean: A permanent invasion or a casual appearance? Sci. Rep. 5 (2015). https://doi.org/10.1038/srep11545.Houghton, J. D. R. et al. Identification of genetically and oceanographically distinct blooms of jellyfish. J. R. Soc. Interface 10(80), 20120920–20120920. https://doi.org/10.1098/rsif.2012.0920 (2013).Article 

    Google Scholar 
    Segura-García, I. et al. Reconstruction of larval origins based on genetic relatedness and biophysical modeling. Sci. Rep. 9(1), 1–9. https://doi.org/10.1038/s41598-019-43435-9 (2019).Article 
    ADS 
    CAS 

    Google Scholar 
    Elphie, H., Raquel, G., David, D. & Serge, P. Detecting immigrants in a highly genetically homogeneous spiny lobster population (Palinurus elephas) in the northwest Mediterranean Sea. Ecol. Evol. 2(10), 2387–2396. https://doi.org/10.1002/ece3.349 (2012).Article 

    Google Scholar 
    Babbucci, M. et al. Population structure, demographic history, and selective processes: Contrasting evidences from mitochondrial and nuclear markers in the European spiny lobster Palinurus elephas (Fabricius, 1787). Mol. Phylogenet. Evol. 56(3), 1040–1050. https://doi.org/10.1016/j.ympev.2010.05.014 (2010).Article 
    CAS 

    Google Scholar 
    Cau, A. et al. European spiny lobster recovery from overfishing enhanced through active restocking in Fully Protected Areas. Sci. Rep. 9(1) (2019). https://doi.org/10.1038/s41598-019-49553-8 .Macias, D., Garcia-Gorriz, E. & Stips, A. Deep winter convection and phytoplankton dynamics in the NW Mediterranean Sea under present climate and future (Horizon 2030) scenarios. Sci. Rep. 8(1), 1–15. https://doi.org/10.1038/s41598-018-24965-0 (2018).Article 
    CAS 

    Google Scholar  More

  • in

    A simple soil mass correction for a more accurate determination of soil carbon stock changes

    Our approach uses hypothetical 30 cm fixed depth samples taken at three successive time points (t0, t1, and t2) with prescribed changes in SOC (1.4% to 1.6%) and BD (1.5–1.1 g cm−3) over these time points (Table 1). The 30 cm soil depth is the common international standard for sampling and analysis required for SOC stock assessment and adhered to by carbon accounting and market organizations6,18. The changes we adopted (a 27% decrease in BD and a 14% increase in SOC) while relatively large, are consistent with those reported in the literature. For example, Reganold and Palmer reported a 25% decrease in BD (1.2–0.9 g cm−3) in neighboring farms with differing management practices23, and Syswerda et al. observed a 17% increase in SOC concentration (10.4–12.2 g C kg soil−1) when converting from a conventionally to organically managed row crop rotation21.Table 1 Hypothetical changes in bulk density (BD) and soil organic carbon (SOC) concentration in 30 cm fixed depth samples at time points t0, t1 and t2 along with calculated values of SOC stock and total soil mass and mineral soil mass.Full size tableIn Table 1, the total soil mass, mineral soil mass, and the SOC stock of the fixed depth samples were calculated by equations as described in the introduction from our prescribed changes in BD and SOC values.ScenariosWe compared hypothetical ESM correction scenarios with our 30 cm fixed depth sample at each time point (Table 2, Figs. 2, 3).Table 2 Hypothetical ESM scenarios showing variation with depth for bulk density (BD) and soil organic carbon (SOC) at each sampling time point, along with the sample depth intervals investigated.Full size tableFigure 2Flow chart of the definition, sampling, and SOC stock correction for a theoretical data set at time points t0, t1, and t2 for scenarios s1 with linear distributions of BD and SOC and s2 with a linear increase in BD and exponential decrease in SOC with depth. Scenario s2 is sampled at (a) 10 cm, (b) 15 cm, and (c) 30 cm intervals.Full size imageFigure 3Scenarios (S1 and S2), showing (a) bulk density variation (BD, g cm−3), and (b) soil organic carbon (SOC, %) variation by depth (0–30 cm) at each time point (t0, t1, and t2). For scenario 2, the single 30 cm depth interval was used (2c). See Table 1 and 2 for details.Full size imageScenario 1We carried out the ESM correction on a 30 cm sample and assumed that the sample was homogenous throughout the profile, with constant SOC and BD values at each time point.To correct for the error in SOC stock estimation when using fixed depth soil sampling, we used  Eqs.2a, 2b and 2c that consider changes in BD28,35. The adjusted soil depth resulting from the change in BD is calculated as:$${mathrm{M}}_{mathrm{n}}= {mathrm{M}}_{mathrm{i}}$$
    (2a)
    $${mathrm{D}}_{mathrm{a}}*{mathrm{BD}}_{mathrm{n}}*left(1-mathrm{k}*{mathrm{SOC}}_{mathrm{n}}right)={mathrm{D}}_{mathrm{i}}*{mathrm{BD}}_{mathrm{i}}*left(1-mathrm{k}*{mathrm{SOC}}_{mathrm{i}}right)$$
    (2b)
    $${mathrm{D}}_{mathrm{a}}={mathrm{D}}_{mathrm{i}}*frac{{mathrm{BD}}_{mathrm{i}}}{{mathrm{BD}}_{mathrm{n}}}*frac{1-mathrm{k}*{mathrm{SOC}}_{mathrm{i}}}{1-mathrm{k}*{mathrm{SOC}}_{mathrm{n}}}$$
    (2c)
    where Mi = Initial mineral soil mass per area (left[frac{M}{{L}^{2}}right]) , Mn = New mineral soil mass per area (left[frac{M}{{L}^{2}}right]) , Da = Adjusted soil surface depth (left[Lright]) , BDi = Initial bulk density (left[frac{M}{{L}^{3}}right]) , BDn = New bulk density (left[frac{M}{{L}^{3}}right]) , SOCi = Initial SOC as a decimal percent (left[frac{M}{M}right]) , SOCn = New SOC as a decimal percent (left[frac{M}{M}right]) , Di = Initial depth (left[Lright]).To conform with Eq. (2a), an increase in SOC over time results in a displacement of some soil mineral mass from the sample, whereas a decrease in SOC over time requires some soil mineral mass to be replaced34. Multiplying the BD by the mineral fraction of the soil (left(1-mathrm{k}*{mathrm{SOC}}right)) for each time point allowed us to compare equivalent mineral mass28. The effect of a change in SOC on mineral mass is small, with a 1% change in SOC equating to approximately a 2% change in apparent depth. This adjustment relates SOC per unit of mineral mass of the fine fraction ( 2 mm)20. The corrected apparent depth can then be used to calculate the corrected SOC stock of a single layer, fixed depth sample (Eq. 3).$$SO{C}_{stock}={D}_{a}*BD*SOC$$
    (3)
    Scenario 2In ESM correction scenarios 2a, 2b, and 2c, we imposed variable, dynamic BD and SOC values with depth over time (Table 2, Figs. 2, 3). To investigate these profiles, we determined the SOC and BD values throughout the soil depth by separating the soil into one (1) cm depth increments (i.e., 0–1 cm, 1–2 cm, etc.). We refer to this calculated incremental profile as the scenario 2 baseline. We assumed that our prescribed SOC concentration varied with depth following an exponential decay. To represent this decay, we simulated the global average distribution of SOC concentration with depth on crop land36, following the distribution from Hobley and Wilson37 (Eq. 4),$$SOCleft(dright)=SO{C}_{infty }+left(SO{C}_{o}-SO{C}_{infty }right)times {e}^{-dk}$$
    (4)
    where SOC (d) is the SOC concentration at depth (d), ({SOC}_{infty }) is the infinity SOC concentration, SOC0 is the SOC concentration at the soil surface, and k is the decay rate. We solved for the decay rate, initial SOC0, and infinity ({SOC}_{infty }) to fit the global average distribution for the 30 cm profile36 and then scaled the SOC concentration to our 30 cm fixed depth sample’s average SOC (1.4%) at t0 (Fig. 3).In scenarios 2a, 2b, and 2c, the BD increased linearly with depth38,39. At the initial time point (t0), we varied the BD values by ± 10% of the BD average over the 30 cm depth, such that for example, BD at t0 (profile average of 1.5 g cm−3) was 1.35 g cm−3 and 1.65 g cm−3 for the upper (0–1 cm) and lower (29–30 cm) depth increment, respectively. For each sequential time point, as the average BD decreased, the soil expanded. To determine the expansion, the depth of the initial sample (e.g., at t0) that filled the 30 cm depth in the subsequent sample (e.g., at t1) was calculated as the initial depth multiplied by the ratio of the average initial BD over the average new BD (e.g., 1.5/1.3 = 1.15 for t0/t1).The linear increase in BD with depth of each following time point maintained the average BD of scenario 1. We then varied the new BD by ± the percent change in the average BD between the time periods (see annotated scripts “main.R” and “functions.R” in Supplementary Material 1 for the development of the theoretical dataset). We then divided each initial BD increment (using soil mass for every 1 cm depth increment) by the new BD in the expanded increment (using soil mass for every  > 1 cm depth increment) to determine the expanded depth of each increment. The SOC value at the initial time represented the same, now expanded, ( > 1 cm) increments, as SOC is a ratio of mass. We used a linear decay rate that was twice that of the percent change in BD between time points to maintain an average BD that was consistent with scenario 1. To model the subsequent fixed depth sample, the BD and SOC concentration values of this expanded soil profile were then interpolated back to the 30 × 1 cm increments of the scenario 2 baseline depth. This calculation preserved the prescribed average BD of the new time point by only expanding the initial SOC concentration.We adjusted the SOC concentration of the next time point to maintain the average SOC concentrations of the 30 cm fixed depth sample, (see annotated scripts “main.R” and “functions.R” in Supplementary Material 1). Because the BD changed between time points and because the SOC stock in the 30 cm fixed depth sample was known, we determined the change in SOC stock between time points by subtracting the average SOC stock in the prior sample from the new sample. We then weighted this change across the 30 cm profile using the distribution of the global soil SOC in the top 30 cm to simulate SOC stratification with reduced tillage or agricultural intensification40. We then multiplied this change by the BD to convert back to SOC concentration and added the delta ((Delta )) SOC value to the prior sample. A worked example is shown in Supplementary Material 2 “Correction Example”.At each time point we split the soil profile at 10 cm and 15 cm depth intervals to create samples for scenarios 2a (3 soil intervals) and 2b (2 soil intervals), respectively. Note that scenario 2c is mathematically equivalent to scenario 1—with only one sample depth interval (30 cm) the sample contains no data on varying SOC or BD. The samples for 2a and 2b were generated by summing the total mass per area and SOC stock values from the scenario 2 baseline to produce single sample values of total soil mass per area and SOC concentration values per depth interval (as would be determined in a laboratory) and calculating BD and mineral mass.In scenario 2, any required additional mineral mass and the associated SOC values were ‘placed’ at the base of the sample to represent a soil profile that had expanded below the fixed 30 cm depth. To account for this, we calculated the increase in adjusted sample depth and accumulated additional soil mineral mass with the lowest sample depth interval of each split sample (Eqs. 5 and 6).$$mathrm{Delta D}={D}_{a}-{D}_{i}$$
    (5)
    $$SO{C}_{stock}={(D}_{1}*B{D}_{1}*SO{C}_{1}+dots + {(D}_{j}+Delta D)*B{D}_{j}*SO{C}_{j}))*10^2 (mathrm{g}/mathrm{cm}^{2})/(mathrm{Mg}/mathrm{ha})$$
    (6)
    where (mathrm{ Delta D}) is the apparent change in depth needed to generate the same mineral mass of the initial sample and the subscript j is the number of sample depth intervals from 1 to j.Varying BD linearly with depth introduces additional complexity in the calculation of the apparent depth. Each sample depth interval may expand (or contract in cases not explored here) at differing rates. Here, the over or under sampling of soil mineral mass is no longer constant with depth and the correction for apparent depth (Da) is estimated with linear interpolation using the BD of each sampling depth interval (i.e., 10 cm, 15 cm, or 30 cm). To do so, we calculated the mineral mass in each depth interval, determined their difference between the initial sample time point and new sample time point, and converted the change in mineral mass to a depth, where:$${mathrm{D}}_{mathrm{a}}={mathrm{D}}_{mathrm{i}}+frac{left(mathrm{sum}left({mathrm{D}}_{mathrm{ij}}*{mathrm{BD}}_{mathrm{ij}}*(1-mathrm{k}*{mathrm{soc}}_{mathrm{ij}}right))- mathrm{sum}left({mathrm{D}}_{mathrm{nj}}*{mathrm{BD}}_{mathrm{nj}}*(1-mathrm{k}*{mathrm{soc}}_{nmathrm{j}})right)right)}{{mathrm{BD}}_{{mathrm{nj}}_{mathrm{bottom}}}*1-mathrm{k}*{mathrm{soc}}_{n{mathrm{j}}_{mathrm{bottom}}}}$$
    (7)
    where jbottom is the lowest sample depth interval, and other terms are as previous. Using Eqs. (5), (6), and (7), with variable BD and SOC values, SOC stock can be corrected using samples split into the 10 cm and 15 cm sampling depth intervals. More

  • in

    Ecological successions throughout the desiccation of Tirez lagoon (Spain) as an astrobiological time-analog for wet-to-dry transitions on Mars

    The ecological baseline in TirezThe geology and the climate of the Tirez region favored the generation and maintenance of a type of hypersaline habitat characterized by extreme seasonality: the sulfate-chlorine waters, with sodium and magnesium cations, showed significant seasonal variations15. The alkaline pH, the low oxidant value for the redox potential of the water column and the highly reduced sediments imposed extreme conditions (see Table 1 and Supplementary Information for details). This extreme seasonality requires to define a valid representative ecological baseline to compare the ecology of the lagoon between 2002 and 2021 and, in this way, set the basis to proposing our model of ecological succession with increasing dryness as a “time-analog” for early Mars. Taxonomic data from 2002 is a snapshot of the community during one season, so we include in our discussion the results presented by Montoya et al. (2013) from a sample campaign carried out in 2005, because they16 analyzed both water and sediment and during both the wet and dry seasons.We consider here only the results obtained by Montoya et al.16 by gene cloning, since those obtained by isolation and sequencing are not comparable. At the level of large groups, no major seasonal differences were observed: Pseudomonadota, followed by Bacteroidetes, were the dominant phyla, in both water and sediments, and both in the dry and the rainy seasons; although Alphaproteobacteria was the dominant class in water, while Gammaproteobacteria was dominant in sediments (in both dry and rainy seasons). With respect to the archaeal domain, all the identified sequences were affiliated to Halobacteriales order, mainly Halorubrum (water) and Halobacterium (sediment), both within Halobacteriaceae family. We can consider these results presented in Montoya et al.16 as the “ecological baseline” for Tirez, however taken with a grain of salt, because only 43 bacterial and 35 archaeal sequences, including rainy and dry seasons and water and sediments, were considered for analysis.Prokaryotic diversity in 2002As can be expected for an extreme environment, the bacterial diversity detected in 2002 was low, although we cannot exclude the possibility that this may reflect the limitation of DNA sequencing techniques at the time. 59% of the obtained clones in the then-wet sediments corresponded to the Malaciobacter genus. Malaciobacter (previously named17 Arcobacter) is an aerotolerant Epsilonproteobacteria. Species within this genus are moderately halophilic, e.g., M. halophilus, capable to grow in up to 4% NaCl. Even though the role that Malaciobacter can play in the environment is not known, it seems to thrive in aquatic systems, like sewage, with a high organic matter content17: e.g., M. canalis, M. cloacae, or M. defluvii.After Malaciobacter-like clones, the next most numerous group belongs to the phylum Bacillota (27% of the sequenced clones; Table 2). Under stressful environmental conditions, members of the genus Virgibacillus produce endospores, a very useful property in an extreme and variable environment (ionic strength, temperature, light intensity), easy to compare with early Mars. Endospores facilitate species survival, allowing them to overcome drastic negative changes, like dry periods, and to germinate when the conditions are favorable again. The closest identified species was the halotolerant V. halodenitrificans, but with low homology, not far from other halotolerant (e.g., V. dokdonensis) or halophilic (e.g., V. marismortui) species within the same genus. The other Gram-positive clones belong to the order Clostridiales. These clones cluster in two taxonomic units related with the strictly anaerobic genus Tissierella.Despite the abundance of Pseudomonadota, their biodiversity was very low, reduced to only two genera within the Epsilon- and Delta-proteobacteria. Six sequences affiliated to Deltaproteobacteria, and clustered in one OTU (salB38, similarity 96.6% with Desulfotignum), were retrieved. Its presence in anaerobic media rich in sulfates (Table 1) seems reasonable. In fact, sulfate-reducing activity was detected using a specific enrichment assay.Finally, one taxon belonging to the phylum Spirochaetota (previously named Spirochaetes) was identified. The presence of Spirochaetota in this system is not strange because members of the genus Spirochaeta are very often found in mud and anaerobic marine environments rich in sulfates18. Moreover, the closest species to SalB63, although with a low similarity of 87%, was Spirochaeta bajacaliforniensis, a spirochete isolated19 from a microbial mat in Laguna Figueroa (Baja California), an extensive hypersaline lagoon with high gypsum content, very similar, although much bigger, than Tirez lagoon.The diversity within the domain Archaea was very low in 2002. The phylogenetic analysis of 96 clones indicate that they correspond to one specie belonging to the obligate halophile genus Methanohalophilus. Their high similarity (99.3%) with several species of Methanohalophilus, such as M. portucalensis (isolated from sediments of a solar saltern in Portugal), M. mahii (isolated from sediments of the Great Salt Lake), or M. halophilus (isolated20 from a cyanobacterial mat at Hamelin Pool, Australia), makes impossible its adscription to any particular species level. Methanohalophilus is strictly methylotrophic, which is consistent with this environment, given that the methylotrophic methanogenesis pathway, non-competitive at low-salt conditions, is predominant at high saline concentrations21. We further confirmed methanogenic activity in Tirez by the measurement of methane by gas chromatography in enrichment cultures.Prokaryotic diversity in context of other studies between 2002 and 2021It was challenging to establish a timeline for the succession of the populations involved, because the scarcity of data harvested and published so far from Tirez. However, combining our results with the few data available in Montoya et al.16 and Preston et al.22 on samplings carried out on 2005 and 2017, respectively, we can see a clear predominance of the phylum Pseudomonadota: Epsilonproteobacteria, i.e. Arcobacter-like, and Deltaproteobacteria, mainly sulfate-reducing bacteria (this work, sampling 2002), and Gammaproteobacteria16 when Tirez maintained a water film, to eventually a final predominance of Gammaproteobacteria, e.g. Chromatiales and Pseudomonadales, in the dry Tirez (this work, 2020 sampling). The Bacillales order has remained widely represented both in the wet and dry Tirez.Regarding the archaeal domain, the few references available (Refs.16,22; this work) confirm that the members of the Halobacteriaceae family are well adapted to both the humid and dry ecosystems of Tirez, being predominant in both conditions. Preston et al.22 found that the second most abundant group of archaea in the dry sediments of Tirez was the Methermicoccaceae family, within the Methanosarcinales order, Methanomicrobia class. Taking into account the results obtained in the dry Tirez (Preston et al.22; and this work, sample 2020), the methanogenic archaea have decreased drastically through time, probably due to salt stress and the competition with sulfate-reducing bacteria.Prokaryotic diversity in 2021From a metabolic point of view, most of the bacteria present today in the sediment are chemoorganotrophs, anaerobes, and halophilic or halotolerant. Scarce information is available about the predominant OTU, Candidate Division OP1. The OP1 division was one of the main bacterial phyla in a sulfur-rich sample in the deepest analyzed samples from the Red Sea sediments under brine pools23. In addition, the phylogenetically related Candidate division KB1 has been observed in deep-sea hypersaline anoxic basins at Orca Basin (Gulf of Mexico), and other hypersaline environments24. Eight of the nine genera identified show coverage greater than 1% of the sequences: i.e., Rubinisphaera, Halothiobacillus, Thiohalophilus, Anaerobacillus/Halolactibacillus, Halomonas, Halothermothrix, and Aliifodinibius are halophilic or halotolerant genera13,25.Regarding archaea, our analyses reveal archaeal groups that seem to thrive in sediments from extreme environments, e.g., marine brine pools/deep water anoxic basins or hypersaline lakes. The most abundant OTU, Thermoplasmata KTK4A, was found prominent and active in the sediment of Lake Strawbridge, a hypersaline lake in Western Australia26, and in soda-saline lakes in China27. The creation of a Candidatus Haloplasmatales, a novel order to include KTK4A-related Thermoplasmata, has been proposed27. On the other hand, both in the aforementioned soda-saline lakes in China27 and in a sulfur-rich section of the sediments from below the Red Sea brine pools23, retrieved sequences were assigned to Marine Benthic Groups B, D, and E. Finally, in the section of nitrogen-rich sediments from the aforementioned Red Sea brine pools, the unclassified lineage ST-12K10A represented the most abundant archaeal group. In the Tirez Lagoon sediment after desiccation, all Methanomicrobia readings belonged to this group.The significance and implications of an ecosystem characterized in 2021 by high diversity, high inequality, and lack of isolated representatives, resides in that Tirez is today an ecosystem in which many (most) of the species/OTUs present are dormant, and they do not play any metabolic role. Hence the high percentage of raretons, greater than 80% for both bacteria and archaea, which are actually present in the lagoon but with only one or two copies each. Only those species adapted to the conditions imposed by the extreme environment are able to actually thrive, and consequently only a few species carry out all the metabolic activity. We conclude that the microbiota in Tirez today represents an ecosystem with a high resilience capacity in the face of environmental changes that may occur.We want to clearly highlight that the technique available in 2002 to study the microbiota of the Tirez lagoon only allowed to obtain a low-resolution image, but that was the state-of-the-art procedure at the time, and the Tirez lagoon cannot be sampled again with the conditions back in 2002, which no longer exist and are not expected to return. Although we have kept in storage several samples of water and sediment from the 2002 Tirez lagoon, it is reasonable to assume that those laboratory microcosms would have chemically and microbiologically changed during the last 20 years, and as such no longer represent reliable replicas of the original lagoon, so we cannot use them for the purposes of this work. Therefore, we are aware that any comparisons of the 2002 laboratory results with the much more robust results obtained by Illumina in 2021 need to be taken with a grain of salt. With all the precautions required, in a high-level, first-order comparison, the most noticeable difference between 2002 and 2021 is a drastic change in the microbial Tirez population. Only some OTUs within Bacillales (Virgibacillus/Anaerobacillus), sulfate-reducing Deltaproteobacteria (Desulfotignum/Desulfobacteraceae-Desulfovibrio), and Spirochaetes are shared among the 2002 and 2021 samples. This comparison is enough for the purposes of this work, as we are interested in the evolution of the lagoon system as a whole to establish a “time-analog” with the wet-to-dry transition on early Mars, and not in the particular outcome of each and every OTU in Tirez. With the results at hand, we conclude that, since 2002, the lacustrine microbiota has shifted to one more adapted to the extreme conditions in the dry sediments, derived from the gradual and persistent desiccation concluding ca. 7 years ago (i.e., completely desiccated in 2015), such as lack of light, absence of oxygen, and lack of water availability. This shift has likely been triggered because organisms that were originally in the lagoon but at low abundance in 2002 became dominant as they were better adapted to desiccation, and because the incoming of new microorganisms transported by birds or wind28.Lipid biomarkers analysis of the desiccated lake sedimentsThe analysis of cell membrane-derived lipid compounds on the dry lake sediments at present allow to provide another perspective of the microbial communities inhabiting the Tirez lagoon, by contributing additional information about the ecosystem and depositional environment. It is important to note that, analyzing only the 2021 lake sediments, we cannot differentiate between lipidic biomarkers of the microorganisms inhabiting Tirez in 2002 and before from those left behind by the microorganisms living in the dried sediments today. Instead, the analyses of lipid biomarkers provide clues about the different microorganisms that have populated Tirez through time, including both older communities inhabiting the former aqueous system and also younger communities better adapted to the present dry conditions. Thus, the lipid biomarkers analysis can be considered as a time-integrative record of the microbial community inhabiting Tirez during the last decades.Based on the molecular distribution of lipid biomarkers, the presence of gram-positive bacteria was inferred from the relative abundance of the monounsaturated alkanoic acid C18:1[ω9], or iso/anteiso pairs of alkanoic acids from 12 to 17 carbons29 with dominance of i/a-C15:0 and i/a-C17:0 (Fig. 3B). In contrast, generally ubiquitous alkanoic acids such as C16:1[ω7], C18:1[ω7], or C18:2[ω6] suggested a provenance rather related to gram-negative bacteria30. The combined detection of the i/a-C15:0 and i/a-C17:0 acids, with dominance of the iso over the anteiso congeners, together with other biomarkers such as the mid-chain branched 10Me16:0, the monounsaturated C17:1, or the cyclopropyl Cy17:0 and Cy19:0 acids, may be associated with a community of SRB31 in today´s dry sediments of Tirez. Specifically, most of those alkanoic acids have been found in a variety of Deltaproteobacteria and/or Bacteroidota (previously named Bacteroidetes). The presence of archaea was deduced from the detection of prominent peaks of archaeol in the polar fraction32 (Fig. 3C), as well as squalene and relatives (dihydrosqualene and tetrahydrosqualene) in the apolar fraction33 (Fig. 3A). Squalene and a variety of unsaturated derivatives are present in the neutral lipid fractions of many archaea with high abundances in saline lakes34. The relative abundance of autotrophs over heterotrophs35 can be estimated by the ratio of the autotrophically-related pristane and phytane over the both autotrophically- and heterotrophically-produced n-C17 and n-C18 alkanes ([Pr + Ph]/[n-C17 + n-C18]). A ratio of 0.56 in the Tirez sediments suggest the presence of a relevant proportion of heterotrophs in the ancient lacustrine system.Furthermore, the lipid biomarkers analysis was able to detect compounds specific of additional microbial sources, such as cyanobacteria36 (n-C17, C17:1, or 7Me-C15 and 7Me-C17), microalgae and/or diatoms (phytosterols37; or C20:5, and C22:6 alkanoic acids30), and other photoautotrophs (phytol and potentially degradative compounds such as pristane and phytane31). A relatively higher preservation of the cell-membrane remnants (i.e., lipids) compared to the DNA-composing nucleic acids may contribute to explain the lack of detection of cyanobacteria, diatoms and microalgae, and other phototrophs by DNA analysis (a deficit in our results shared with Montoya et al.16, and Preston et al.22). Although abundant in higher plants38, sterols such as those detected here (i.e., the sterols campesterol, stigmasterol, and β-sitosterol, as well as ergosterol) are also major sterols in some microalgal classes37 (such as Bacillariophyceae, Chrysophyceae, Euglenophyceae, Eustigmatophyceae, Raphidophyceae, Xanthophyceae, and Chlorophyceae), cyanobacteria (β-sitosterol), and fungi (ergosterol39).The carbon isotopic composition of lipid biomarkers provides a rapid screening of the carbon metabolism in a system, by recognizing the principal carbon fixation pathways used by autotrophs. The range of δ13C values measured in the Tirez sediments (from − 33.9 to − 16.1‰) denotes a mixed use of different carbon assimilation pathways, involving mostly the reductive pentose phosphate (a.k.a. Calvin–Benson–Bassham or just Calvin) cycle (from − 19 to − 30‰), and in lesser extent the reductive acetyl-CoA (a.k.a. Wood–Ljungdahl) pathway (from − 28 to − 44‰), and/or the reverse tricarboxylic acid (rTCA) cycle (from − 12 to − 21‰).The lipids synthesized by microorganisms using the Calvin or reductive acetyl-CoA pathway are typically depleted relative to the bulk biomass, particularly those produced via de latter pathway. In the dry Tirez sediments, the majority of the lipid compounds are more depleted in 13C than the bulk biomass (Fig. 4). In particular, the branched alkane DiMeC18 (Fig. 4A) and the SRB-indicative 10Me16:0 acid (Fig. 4B) showed the most depleted δ13C values and suggested the use of the reductive acetyl-CoA pathway. The rest of lipid compounds showed isotopic signatures (from − 16.1 to − 31.4‰) compatible with the prevalence of the Calvin pathway. These values may directly reflect the autotrophic activity of microorganisms fixing carbon via the Calvin cycle or heterotrophic activity of microorganisms growing on their remnants. Thus, the saturated and linear alkyl chains of lipids (i.e., n-alkanes, n-alkanoic acids, and n-alkanols) showing the most negative δ13C values (e.g., alkanes n-C17 and C17:1; or acid C18:1[ω7]) reflect prokaryotic sources of Calvin-users autotrophs (e.g., cyanobacteria or purple sulfur bacteria), while the rest of compounds with slightly less negative δ13C values instead stem from the autotrophic activity of eukaryotes also users of the Calvin cycle (unsaturated fatty acids and sterols) or from the metabolism of heterotrophs such as SRB (iso/anteiso-, other branched, and cyclopropyl fatty acids) and haloarchaea (isoprenoids, phytanol, and archaeol). All in all, the compound-specific isotope composition of the dry sediments in the today´s Tirez lagoon may indirectly reflect the prevailing autotrophic mechanisms in the present lacustrine system of Tirez, by showing isotopic signatures of secondary lipids similar to their carbon source40.In addition, the use of a number of lipid molecular ratios or proxies allow further characterization of the lacustrine ecosystem and depositional environment. For example, the average chain length of the n-alkanes (24.1) suggests a relevant presence of eukaryotic biomass in the lacustrine sediments, since long-chained alkanes ( > C20) are known to originate from epicuticular leaf waxes in higher plants41. Highlighting the relevance of eukaryotes and their ecological roles is one of the major contributions of this work, because previous studies on the microbial ecology of hypersaline environments have been focused primarily on prokaryotes42.The proportion of odd n-alkanes of high molecular chain (i.e., n-C27, n-C29, and n-C31) over even n-alkanes of low molecular chain (i.e., n-C15, n-C17, and n-C19) provides an estimate of the relative abundance of terrigenous over aqueous biomass43, which in Tirez is TAR = 1.8. The Paq index may also be used to differentiate the proportion of terrigenous versus aquatic (emergent and submerged) plant biomass44. A Paq of 0.3 in the Tirez sediments from 2021 supported the relative abundance of land plants. Finally, the depositional environment in the lacustrine system of Tirez may be also characterized analyzing the ratio of pristane over phytane (Pr/Ph), which is higher than 1 when phytol degrades to pristane under oxic conditions45. Assuming that both isoprenoids in the Tirez sediments derived from phytol31, according to their similarly depleted δ13C (Fig. 4A), we can conclude that the sediments in the Tirez lagoon were deposited under predominantly oxic conditions (i.e., Pr/Ph ratio of 1.1).In summary, the lipid biomarkers study revealed useful information about the depositional environment and lacustrine ecosystem, including the presence of active or past autotrophic metabolisms involving prokaryotes (e.g., cyanobacteria and purple sulfur bacteria) and eukaryotes (plants, diatoms and other microalgae), as well as heterotrophic metabolisms of likely SRB and haloarchaea growing on Calvin-users exudates. These results are quite in agreement with the microbial community previously reported16 in sediments from the wet and dry seasons: abundant Gammaproteobacteria and Alphaproteobacteria, together with Algae and Cyanobacteria, dinoflagellates and filamentous fungi, Bacillota, Actinomycetota (previously named Actinomycetes), and a halophilic sulfate-reducing Deltaproteobacteria.Tirez as the first astrobiological “time-analog” for early MarsEarly Mars most likely had a diversity of environments in terms of pH, redox conditions, geochemistry, temperature, and so on. Field research in terrestrial analog environments contribute to understand the habitability of this diversity of environments on Mars in the past, because terrestrial analogues are places on Earth characterized by environmental, mineralogical, geomorphological, or geochemical conditions similar to those observed on present or past Mars9. Therefore, so far analogs have been referred to terrestrial locations closely similar to any of the geochemical environments that have been inferred on Mars, i.e., they are “site-analogs” that represent snapshots in time: one specific environmental condition at a very specific place and a very specific time. Because of this, each individual field analog site cannot be considered an adequate representation of the changing martian environmental conditions through time. Here we introduce the concept of astrobiological “time-analog”, referred to terrestrial analogs that may help understand environmental transitions and the related possible ecological successions on early Mars. In this sense, they should be “time-resolved analogs”: dynamic analog environments where we can analyze changes over time. To the best of our knowledge, this is the first study that looks at the environmental microbiology of a Mars astrobiological analog site over a significant and long period of change, and try to understand the ecological successions to put them in the context of martian environmental evolution.As Mars lost most of its surface water at the end of the Hesperian5,9,12, this wet-to-dry global transition can be considered the major environmental perturbation in the geological history of Mars, and therefore merits to be the first one to be assigned a “time-analog” for its better understanding and characterization. The drying of Mars was probably a stepwise process, characterized by multiple transitions between drier and wetter environments12,47, and therefore the seasonal fluctuations and eventual full desiccation of Tirez represent a suitable analog to better understand possible ecological transitions during the global desiccation of most of the Mars’s surface before the Amazonian (beginning 3.2 Ga).To introduce Tirez as the first Mars astrobiological “time-analog” of the wet-to-dry transition on early Mars, the objective of this study was threefold: first, we wanted to identify the dominant prokaryotic microorganisms in the active Tirez lagoon 20 years ago, a unique hypersaline ecosystem with an ionic composition different from that of marine environments, and therefore potentially analogous to ancient saline lacustrine environments on Mars during the Noachian and into the Hesperian46,47. Our results provide a preliminary basis to hypothesize how the microbial communities on the Noachian Mars could have developed in salty environments with dramatically fluctuating water availability. The requirement to deal with important variations in ionic strength and water availability, involving at times the complete evaporation of the water, could have represented additional constraints48 for microorganisms on early Mars.The second objective of this investigation was the identification of the microbial community inhabiting the desiccated Tirez sediments today, after all the water was lost, as a potential analog to desiccated basins on Mars at the end of the Hesperian1,3,4,47. Our results suggest that hypothetical early microbial communities on early Mars, living with relative abundance of liquid water during the Noachian, would have been forced to adapt to increasingly desiccating surface environments, characterized by extreme conditions derived from the persistent dryness and lack of water availability. Our investigation in Tirez suggest that hypothetical microorganisms at the end of the Hesperian would have needed to evolve strategies similar to those of microorganisms on Earth adapted to living at very low water activity49, to thrive in the progressively desiccating sediments.And the third objective of this investigation was the identification of the lipidic biomarkers left behind by the microbial communities in Tirez, as a guide to searching and identifying the potential leftovers of a hypothetical ancient biosphere on Mars. Lipids (i.e., fatty acids and other biosynthesized hydrocarbons) are structural components of cell membranes bearing recognized higher resistance to degradation relative to other biomolecules, thus with potential to reconstruct paleobiology in a broader temporal scale than more labile molecules50. Our results reinforce the notion that lipidic biomarkers should be preferred targets in the search for extinct and/or extant life on Mars precisely because they are so recalcitrant. More