More stories

  • in

    Mineralogy of microbially induced calcium carbonate precipitates formed using single cell drop-based microfluidics

    1.
    Dong, H. L. & Lu, A. H. Mineral-microbe interactions and implications for remediation. Elements 8, 95–100. https://doi.org/10.2113/gselements.8.2.95 (2012).
    CAS  Article  Google Scholar 
    2.
    Li, X. B. et al. Spatial patterns of carbonate biomineralization in biofilms. Appl. Environ. Microbiol. 81, 7403–7410. https://doi.org/10.1128/aem.01585-15 (2015).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    3.
    Miot, J., Benzerara, K. & Kappler, A. In Annual Review of Earth and Planetary Sciences, Vol. 42 Annual Review of Earth and Planetary Sciences (ed R. Jeanloz) 271–289 (Annual Reviews, 2014).

    4.
    Brookshaw, D., Pattrick, R., Lloyd, J. & Vaughan, D. Microbial effects on mineral–radionuclide interactions and radionuclide solid-phase capture processes. Mineral. Mag. 76, 200. https://doi.org/10.1180/minmag.2012.076.3.25 (2012).
    CAS  Article  Google Scholar 

    5.
    Li, J. et al. Biomineralization, crystallography and magnetic properties of bullet-shaped magnetite magnetosomes in giant rod magnetotactic bacteria. Earth Planet. Sci. Lett. 293, 368–376. https://doi.org/10.1016/j.epsl.2010.03.007 (2010).
    ADS  CAS  Article  Google Scholar 

    6.
    Douglas, S. & Beveridge, T. J. Mineral formation by bacteria in natural microbial communities. FEMS Microbiol. Ecol. 26, 79–88. https://doi.org/10.1111/j.1574-6941.1998.tb00494.x (1998).
    CAS  Article  Google Scholar 

    7.
    Dick, G. et al. The microbiology of deep-sea hydrothermal vent plumes: ecological and biogeographic linkages to seafloor and water column habitats. Front. Microbiol. https://doi.org/10.3389/fmicb.2013.00124 (2013).
    Article  PubMed  PubMed Central  Google Scholar 

    8.
    Hatzenpichler, R., Krukenberg, V., Spietz, R. L. & Jay, Z. J. Next-generation physiology approaches to study microbiome function at single cell level. Nat. Rev. Microbiol. https://doi.org/10.1038/s41579-020-0323-1 (2020).
    Article  PubMed  Google Scholar 

    9.
    Phillips, A. J. et al. Engineered applications of ureolytic biomineralization: a review. Biofouling 29, 715–733. https://doi.org/10.1080/08927014.2013.796550 (2013).
    CAS  Article  PubMed  Google Scholar 

    10.
    Cuthbert, M. O. et al. Controls on the rate of ureolysis and the morphology of carbonate precipitated by S. pasteurii biofilms and limits due to bacterial encapsulation. Ecol. Eng. 41, 32–40 (2012).
    Article  Google Scholar 

    11.
    De Muynck, W., Verbeken, K., De Belie, N. & Verstraete, W. Influence of urea and calcium dosage on the effectiveness of bacterially induced carbonate precipitation on limestone. Ecol. Eng. 36, 99–111. https://doi.org/10.1016/j.ecoleng.2009.03.025 (2010).
    Article  Google Scholar 

    12.
    Ferris, F. G., Fyfe, W. S. & Beveridge, T. J. Bacteria as nucleation sites for authigenic minerals in a metal-contaminated lake sediment. Chem. Geol. 63, 225–232. https://doi.org/10.1016/0009-2541(87)90165-3 (1987).
    ADS  CAS  Article  Google Scholar 

    13.
    Hammes, F. & Verstraete, W. Key roles of pH and calcium metabolism in microbial carbonate precipitation. Rev. Environ. Sci. Biotechnol. 1, 3–7. https://doi.org/10.1023/a:1015135629155 (2002).
    CAS  Article  Google Scholar 

    14.
    van Paassen, L. Biogrout, Ground Improvement by Microbial Induced Carbonate Precipitation. Doctoral Thesis. (Delft University of Technology, Delft, Netherlands, 2009).

    15.
    Zhang, W., Ju, Y., Zong, Y., Qi, H. & Zhao, K. In situ real-time study on dynamics of microbially induced calcium carbonate precipitation at a single-cell level. Environ. Sci. Technol. https://doi.org/10.1021/acs.est.8b02660 (2018).
    Article  PubMed  PubMed Central  Google Scholar 

    16.
    Dohnalkova, A. C. et al. Imaging hydrated microbial extracellular polymers: comparative analysis by electron microscopy. Appl. Environ. Microbiol. 77, 1254–1262. https://doi.org/10.1128/aem.02001-10 (2011).
    CAS  Article  PubMed  Google Scholar 

    17.
    Kaminski, T. S., Scheler, O. & Garstecki, P. Droplet microfluidics for microbiology: techniques, applications and challenges. Lab Chip 16, 2168–2187. https://doi.org/10.1039/c6lc00367b (2016).
    CAS  Article  PubMed  Google Scholar 

    18.
    Vincent, M. E., Liu, W. S., Haney, E. B. & Ismagilov, R. F. Microfluidic stochastic confinement enhances analysis of rare cells by isolating cells and creating high density environments for control of diffusible signals. Chem. Soc. Rev. 39, 974–984. https://doi.org/10.1039/b917851a (2010).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    19.
    Taheri-Araghi, S., Brown, S. D., Sauls, J. T., McIntosh, D. B. & Jun, S. In Annual Review of Biophysics, Vol. 44 (ed K. A. Dill) 123–142 (Annual Reviews, 2015).

    20.
    Pratt, S. L. et al. DropSOAC: stabilizing microfluidic drops for time-lapse quantification of single-cell bacterial physiology. Front. Microbiol. https://doi.org/10.3389/fmicb.2019.02112 (2019).
    Article  PubMed  PubMed Central  Google Scholar 

    21.
    Schmitz, C. H. J., Rowat, A. C., Köster, S. & Weitz, D. A. Dropspots: a picoliter array in a microfluidic device. Lab Chip 9, 44–49. https://doi.org/10.1039/B809670H (2009).
    CAS  Article  PubMed  Google Scholar 

    22.
    Rodriguez-Ruiz, I. et al. Transient calcium carbonate hexahydrate (ikaite) nucleated and stabilized in confined nano- and picovolumes. Cryst. Growth Des. 14, 792–802. https://doi.org/10.1021/cg401672v (2014).
    CAS  Article  Google Scholar 

    23.
    Yashina, A., Meldrum, F. & deMello, A. Calcium carbonate polymorph control using droplet-based microfluidics. Biomicrofluidics 6, 10. https://doi.org/10.1063/1.3683162 (2012).
    CAS  Article  Google Scholar 

    24.
    Liu, P. Y., Yao, J., Couples, G. D., Ma, J. S. & Iliev, O. 3-D modelling and experimental comparison of reactive flow in carbonates under radial flow conditions. Sci. Rep. 7, 10. https://doi.org/10.1038/s41598-017-18095-2 (2017).
    ADS  CAS  Article  Google Scholar 

    25.
    Connolly, J. et al. Construction of two ureolytic model organisms for the study of microbially induced calcium carbonate precipitation. J. Microbiol. Methods 94, 290–299. https://doi.org/10.1016/j.mimet.2013.06.028 (2013).
    CAS  Article  PubMed  Google Scholar 

    26.
    Duffy, D. C., McDonald, J. C., Schueller, O. J. & Whitesides, G. M. Rapid prototyping of microfluidic systems in poly(dimethylsiloxane). Anal. Chem. 70, 4974–4984. https://doi.org/10.1021/ac980656z (1998).
    CAS  Article  PubMed  Google Scholar 

    27.
    Holtze, C. et al. Biocompatible surfactants for water-in-fluorocarbon emulsions. Lab Chip 8, 1632–1639. https://doi.org/10.1039/b806706f (2008).
    CAS  Article  PubMed  Google Scholar 

    28.
    Jung, D., Biggs, H., Erikson, J. & Ledyard, P. U. New colorimetric reaction for endpoint, continuous-flow, and kinetic measurement of urea. Clin. Chem. 21, 1136–1140 (1975).
    CAS  Article  Google Scholar 

    29.
    Phillips, A. J. et al. Fracture sealing with microbially-induced calcium carbonate precipitation: a field study. Environ. Sci. Technol. 50, 4111–4117. https://doi.org/10.1021/acs.est.5b05559 (2016).
    ADS  CAS  Article  PubMed  Google Scholar 

    30.
    Yoshida, N., Higashimura, E. & Saeki, Y. Catalytic biomineralization of fluorescent calcite by the thermophilic bacterium Geobacillus thermoglucosidasius. Appl. Environ. Microbiol. 76, 7322–7327. https://doi.org/10.1128/aem.01767-10 (2010).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    31.
    Tinevez, J. Y. et al. TrackMate: an open and extensible platform for single-particle tracking. Methods 115, 80–90. https://doi.org/10.1016/j.ymeth.2016.09.016 (2017).
    CAS  Article  PubMed  Google Scholar 

    32.
    Vejborg, R. M. & Klemm, P. Cellular chain formation in Escherichia coli biofilms. Microbiology (Reading, England) 155, 1407–1417. https://doi.org/10.1099/mic.0.026419-0 (2009).
    CAS  Article  Google Scholar 

    33.
    Rodriguez-Blanco, J. D., Shaw, S. & Benning, L. G. The kinetics and mechanisms of amorphous calcium carbonate (ACC) crystallization to calcite, via vaterite. Nanoscale 3, 265–271. https://doi.org/10.1039/c0nr00589d (2011).
    ADS  CAS  Article  PubMed  Google Scholar 

    34.
    Tlili, M. M. et al. Characterization of CaCO3 hydrates by micro-Raman spectroscopy. J. Raman Spectrosc. 33, 10–16. https://doi.org/10.1002/jrs.806 (2002).
    ADS  CAS  Article  Google Scholar 

    35.
    Clarkson, J. R., Price, T. J. & Adams, C. J. Role of metastable phases in the spontaneous precipitation of calcium-carbonate. J. Chem. Soc. Faraday Trans. 88, 243–249. https://doi.org/10.1039/ft9928800243 (1992).
    CAS  Article  Google Scholar 

    36.
    Gabrielli, C., Jaouhari, R., Joiret, S., Maurin, G. & Rousseau, P. Study of the electrochemical deposition of CaCO3 by in situ Raman spectroscopy—I. Influence of the substrate. J. Electrochem. Soc. 150, C478–C484. https://doi.org/10.1149/1.1579482 (2003).
    CAS  Article  Google Scholar 

    37.
    Behrens, G., Kuhn, L. T., Ubic, R. & Heuer, A. H. Raman-spectra of vateritic calcium-carbonate. Spectrosc. Lett. 28, 983–995. https://doi.org/10.1080/00387019508009934 (1995).
    ADS  CAS  Article  Google Scholar 

    38.
    Wagner, O. et al. Biocompatible fluorinated polyglycerols for droplet microfluidics as an alternative to PEG-based copolymer surfactants. Lab Chip 16, 65–69. https://doi.org/10.1039/C5LC00823A (2016).
    CAS  Article  PubMed  Google Scholar 

    39.
    Gruner, P. et al. Controlling molecular transport in minimal emulsions. Nat. Commun. 7, 10392. https://doi.org/10.1038/ncomms10392 (2016).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    40.
    Bai, Y. et al. A double droplet trap system for studying mass transport across a droplet-droplet interface. Lab Chip 10, 1281–1285. https://doi.org/10.1039/b925133b (2010).
    CAS  Article  PubMed  Google Scholar 

    41.
    Hoffmann, R. et al. TEM preparation methods and influence of radiation damage on the beam sensitive CaCO3 shell of Emiliania huxleyi. Micron 62, 28–36. https://doi.org/10.1016/j.micron.2014.03.004 (2014).
    CAS  Article  PubMed  Google Scholar 

    42.
    Hooley, R., Brown, A. & Brydson, R. Factors affecting electron beam damage in calcite nanoparticles. Micron 120, 25–34. https://doi.org/10.1016/j.micron.2019.01.011 (2019).
    CAS  Article  PubMed  Google Scholar 

    43.
    Bentov, S., Weil, S., Glazer, L., Sagi, A. & Berman, A. Stabilization of amorphous calcium carbonate by phosphate rich organic matrix proteins and by single phosphoamino acids. J. Struct. Biol. 171, 207–215. https://doi.org/10.1016/j.jsb.2010.04.007 (2010).
    CAS  Article  PubMed  Google Scholar 

    44.
    Rautaray, D., Sanyal, A., Bharde, A., Ahmad, A. & Sastry, M. Biological synthesis of stable vaterite crystals by the reaction of calcium ions with germinating chickpea seeds. Cryst. Growth Des. 5, 399–402. https://doi.org/10.1021/cg0341858 (2005).
    CAS  Article  Google Scholar 

    45.
    Kawaguchi, T. & Decho, A. W. A laboratory investigation of cyanobacterial extracellular polymeric secretions (EPS) in influencing CaCO3 polymorphism. J. Cryst. Growth 240, 230–235. https://doi.org/10.1016/s0022-0248(02)00918-1 (2002).
    ADS  CAS  Article  Google Scholar 

    46.
    Wang, S.-S., Picker, A., Cölfen, H. & Xu, A.-W. Heterostructured calcium carbonate microspheres with calcite equatorial loops and vaterite spherical cores. Angew. Chem. Int. Ed. 52, 6317–6321. https://doi.org/10.1002/anie.201301184 (2013).
    CAS  Article  Google Scholar 

    47.
    Rao, A. & Cölfen, H. Morphology control and molecular templates in biomineralization. Biominer. Biomater. https://doi.org/10.1016/B978-1-78242-338-6.00003-X (2016).
    Article  Google Scholar 

    48.
    Aggarwal, K. P., Narula, S., Kakkar, M. & Tandon, C. Nephrolithiasis: molecular mechanism of renal stone formation and the critical role played by modulators. Biomed Res Int 2013, 292953–292953. https://doi.org/10.1155/2013/292953 (2013).
    CAS  Article  Google Scholar  More

  • in

    Acceleration of ocean warming, salinification, deoxygenation and acidification in the surface subtropical North Atlantic Ocean

    Sampling methods
    Ocean sampling at hydrostation S and BATS started in 1954 and 1988, respectively. Hydrostation S began with the pioneering efforts of Hank Stommel (Woods Hole Oceanographic Institution) and colleagues50 at a site approximately 26 km southeast of Bermuda (32° 10′N, 64° 30′W, Fig. 1). The first water–column sampling occurred on the 7th June 1954 from the 61′ R.V. Panularis with more than 1381 cruises conducted up to the present time from the R.V. Panularis II (1967–1983), R.V. Weatherbird (1983–1989); R.V. Weatherbird II (1989–2006) and R.V. Atlantic Explorer (2006–present).
    Since the arrival of the R.V. Weatherbird, Hydrostation S has been occupied at near biweekly intervals, with multiple CTD–hydrocasts through the water-column to ~2600 m, and to 4500 m at BATS (more than 450 cruises to the site). Nansen bottles were used for water sampling at first, then 5 L Niskin samplers until October 1988. Thereafter sampling has been conducted with a Seabird 9/11 CTD equipped with 12 L Niskin and Ocean Test Equipment (OTE) samplers.
    Water sampling, temperature and CTD measurements
    The sampling format has remained substantially consistent for the past 65 years, but with the introduction of CTD–hydrocast sampling in October 1988. From 1954 to 1988, reversing mercury thermometers were used for measurements of temperature until replaced by CTD measurements using a Sea-Bird 9/11 system. Bottle samples before 2012 were taken down to 2600 m at hydrostation S. Since the addition of reliable altimeters to the CTD package, sampling was extended to full ocean depth at both the hydrostation S (~3400 m) and BATS (~4500 m) sites. Numerous sensor configurations have been used on the CTD package (e.g., dual temperature, dual conductivity, dual DO sensors, transmissometer, fluorometer, PAR and altimeter). In contrast, the CTD sampling system has predominately been a Seabird 24-place rosette using 12 L Ocean Test bottles. Before profiling, the CTD is allowed to stabilise at 10 m and once stable, the CTD returns to the surface to start the profile with typical descent rates of 0.5–1.0 m s−1, depending on weather conditions. Water samples are collected on the upcast, whereby the OTE bottles are closed at the target depth after a waiting period of 45 s. The CTD is held at the target depth for another 10 s to allow the SBE35-RT sensor to take an 8 s average. The CTD continues with the upcast at an ascent rate of 0.7–1.0 m s−1. Temperature, conductivity and DO sensors are routinely returned to SeaBird every 6–9 months for routine calibration. The differences between primary and secondary temperature sensors in the deep ocean at BATS ( >3000 m) were 0.002–0.006 °C regardless of time since most recent factory calibration.
    Determination of salinity
    Salinity samples are typically taken from the OTE bottles at all depths. These samples are collected immediately following DO and CO2 sampling. Samples are taken in 125–250 ml borosilicate glass bottles (Ocean Scientific, UK) that use plastic thimbles to form a better seal. The sample remaining from the previous use is left in the bottles between cruises to prevent salt crystal buildup due to evaporation. When drawing a new sample, the old sample is first discarded over the sampling spigot, and the bottle is rinsed three times with water from the new sample. The bottle is then filled to the shoulder with the sample, and the thimble inserted into the container. The neck of the bottle and the inside of the cap are dried, then the thimble is inserted, and the cap is replaced and firmly tightened. These samples are stored in a temperature-controlled laboratory for later analysis (typically within 1–2 weeks of their collection).
    Salinity measurements have been made with a Guildline salinometer at BIOS from 1981 to present (calibrated with IAPSO standard water51) for both BATS and Hydrostation S samples. At present, samples for salinity are analysed on a Guildline Autosal 8400B laboratory salinometer using the manufacturer’s recommended techniques. All readings (10 s average) are taken using the Ocean Scientific interface box and PC software. The salinometer drift during and between successive runs tends to be zero (room temperature carefully controlled and monitored). Bottle salinities are used to calibrate the profiling CTD SBE-04 sensors, and additionally, they are also compared with the downcast CTD profiles to search for possible outliers. Deep-water samples ( >2000 m) are replicated for precision estimates (typically  More

  • in

    Microbial growth and carbon use efficiency show seasonal responses in a multifactorial climate change experiment

    1.
    Soong, J. L. et al. Microbial carbon limitation: the need for integrating microorganisms into our understanding of ecosystem carbon cycling. Glob. Chang. Biol. 00, 1–9 (2020).
    Google Scholar 
    2.
    Liang, C. & Balser, T. C. Microbial production of recalcitrant organic matter in global soils: implications for productivity and climate policy. Nat. Rev. Microbiol. 9, 75 (2011).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    3.
    Miltner, A., Bombach, P., Schmidt-Brücken, B. & Kästner, M. SOM genesis: microbial biomass as a significant source. Biogeochemistry 111, 41–55 (2012).
    CAS  Article  Google Scholar 

    4.
    Lehmann, J. & Kleber, M. The contentious nature of soil organic matter. Nature 528, 60–68 (2015).
    CAS  Article  Google Scholar 

    5.
    Manzoni, S., Taylor, P., Richter, A., Porporato, A. & Ågren, G. I. Environmental and stoichiometric controls on microbial carbon-use efficiency in soils. N. Phytol. 196, 79–91 (2012).
    CAS  Article  Google Scholar 

    6.
    Sinsabaugh, R. L., Manzoni, S., Moorhead, D. L. & Richter, A. Carbon use efficiency of microbial communities: Stoichiometry, methodology and modelling. Ecol. Lett. 16, 930–939 (2013).
    PubMed  Article  PubMed Central  Google Scholar 

    7.
    Allison, S. D. Modeling adaptation of carbon use efficiency in microbial communities. Front. Microbiol. 5, 1–9 (2014).
    Google Scholar 

    8.
    Geyer, K. M., Kyker-Snowman, E., Grandy, A. S. & Frey, S. D. Microbial carbon use efficiency: accounting for population, community, and ecosystem-scale controls over the fate of metabolized organic matter. Biogeochemistry 127, 173–188 (2016).
    CAS  Article  Google Scholar 

    9.
    Hagerty, S. B., Allison, S. D. & Schimel, J. P. Evaluating soil microbial carbon use efficiency explicitly as a function of cellular processes: implications for measurements and models. Biogeochemistry 140, 269–283 (2018).
    CAS  Article  Google Scholar 

    10.
    Manzoni, S., Schimel, J. P. & Porporato, A. Responses of soil microbial communities to water stress: results from a meta-analysis. Ecology 93, 930–938 (2011).
    Article  Google Scholar 

    11.
    Williams, M. A. & Rice, C. W. Seven years of enhanced water availability influences the physiological, structural, and functional attributes of a soil microbial community. Appl. Soil Ecol. 35, 535–545 (2007).
    Article  Google Scholar 

    12.
    Zheng, Q. et al. Growth explains microbial carbon use efficiency across soils differing in land use and geology. Soil Biol. Biochem. 128, 45–55 (2019).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    13.
    Pritchard, S. G. Soil organisms and global climate change. Plant Pathol. 60, 82–99 (2011).
    Article  Google Scholar 

    14.
    Classen, A. E. T. et al. Direct and indirect effects of climate change on soil microbial and soil microbial-plant interactions: What lies ahead? Ecosphere 6, 1–21 (2015).
    Article  Google Scholar 

    15.
    IPCC. Summary for policymakers. In Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change (eds Stocker, T. F. et al.) (Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, 2013).

    16.
    Dieleman, W. I. J. et al. Simple additive effects are rare: A quantitative review of plant biomass and soil process responses to combined manipulations of CO2 and temperature. Glob. Chang. Biol. 18, 2681–2693 (2012).
    PubMed  Article  PubMed Central  Google Scholar 

    17.
    Yue, K. et al. Influence of multiple global change drivers on terrestrial carbon storage: additive effects are common. Ecol. Lett. 20, 663–672 (2017).
    PubMed  Article  PubMed Central  Google Scholar 

    18.
    Song, J. et al. A meta-analysis of 1,119 manipulative experiments on terrestrial carbon-cycling responses to global change. Nat. Ecol. Evol. 3, 1309–1320 (2019).
    PubMed  Article  PubMed Central  Google Scholar 

    19.
    Le Quéré, C. et al. Global Carbon Budget 2018. Earth Syst. Sci. Data Discuss. 10, 2141–2194 (2018).
    Article  Google Scholar 

    20.
    Bloom, A. A., Exbrayat, J. F., Van Der Velde, I. R., Feng, L. & Williams, M. The decadal state of the terrestrial carbon cycle: Global retrievals of terrestrial carbon allocation, pools, and residence times. Proc. Natl Acad. Sci. USA 113, 1285–1290 (2016).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    21.
    Alster, C. J., Baas, P., Wallenstein, M. D., Johnson, N. G. & von Fischer, J. C. Temperature sensitivity as a microbial trait using parameters from macromolecular rate theory. Front. Microbiol. 7, 1–10 (2016).
    Article  Google Scholar 

    22.
    Bradford, M. A. Thermal adaptation of decomposer communities in warming soils. Front. Microbiol. 4, 1–16 (2013).
    Article  Google Scholar 

    23.
    Davidson, E. A. & Janssens, I. A. Temperature sensitivity of soil carbon decomposition and feedbacks to climate change. Nature 440, 165–173 (2006).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    24.
    Bárcenas-Moreno, G., Gómez-Brandón, M., Rousk, J. & Bååth, E. Adaptation of soil microbial communities to temperature: comparison of fungi and bacteria in a laboratory experiment. Glob. Chang. Biol. 15, 2950–2957 (2009).
    Article  Google Scholar 

    25.
    Fenner, N., Freeman, C. & Reynolds, B. Observations of a seasonally shifting thermal optimum in peatland carbon-cycling processes; implications for the global carbon cycle and soil enzyme methodologies. Soil Biol. Biochem. 37, 1814–1821 (2005).
    CAS  Article  Google Scholar 

    26.
    Schimel Balser, T. C. & Wallenstein, M. Microbial stress-response physiology and its implications for ecosystem function. Ecology 88, 1386–1394 (2007).
    PubMed  Article  PubMed Central  Google Scholar 

    27.
    Schimel. Life in dry soils: effects of drought on soil microbial communities and processes. Annu. Rev. Ecol. Evol. Syst. 49, 409–432 (2018).
    Article  Google Scholar 

    28.
    Bardgett, R. D., Freeman, C. & Ostle, N. J. Microbial contributions to climate change through carbon cycle feedbacks. ISME J. 2, 805–814 (2008).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    29.
    Andresen, L. C. et al. Biomass responses in a temperate European grassland through 17 years of elevated CO2. Glob. Chang. Biol. 24, 3875–3885 (2018).
    PubMed  Article  PubMed Central  Google Scholar 

    30.
    Parvin, S. et al. Free air CO2 enrichment (FACE) improves water use efficiency and moderates drought effect on N2 fixation of Pisum sativum L. Plant Soil 436, 587–606 (2019).
    CAS  Article  Google Scholar 

    31.
    Yuhui, W. et al. Effects of elevated CO2 and drought on plant physiology, soil carbon and soil enzyme activities. Pedosphere 27, 846–855 (2017).
    Article  Google Scholar 

    32.
    Eisenhauer, N., Cesarz, S., Koller, R., Worm, K. & Reich, P. B. Global change belowground: Impacts of elevated CO2, nitrogen, and summer drought on soil food webs and biodiversity. Glob. Chang. Biol. 18, 435–447 (2012).
    Article  Google Scholar 

    33.
    Nie, M., Lu, M., Bell, J., Raut, S. & Pendall, E. Altered root traits due to elevated CO2: A meta-analysis. Glob. Ecol. Biogeogr. 22, 1095–1105 (2013).
    Article  Google Scholar 

    34.
    Tucker, C. L., Bell, J., Pendall, E. & Ogle, K. Does declining carbon-use efficiency explain thermal acclimation of soil respiration with warming? Glob. Chang. Biol. 19, 252–263 (2013).
    PubMed  Article  PubMed Central  Google Scholar 

    35.
    Kuzyakov, Y., Horwath, W. R., Dorodnikov, M. & Blagodatskaya, E. Review and synthesis of the effects of elevated atmospheric CO2 on soil processes: No changes in pools, but increased fluxes and accelerated cycles. Soil Biol. Biochem. 128, 66–78 (2019).
    CAS  Article  Google Scholar 

    36.
    Wan, S., Norby, R. J., Ledford, J. & Weltzin, J. F. Responses of soil respiration to elevated CO2, air warming, and changing soil water availability in a model old-field grassland. Glob. Chang. Biol. 13, 2411–2424 (2007).
    Article  Google Scholar 

    37.
    Madhu, M. & Hatfeld, J. L. Dynamics of plant root growth under increased atmospheric carbon dioxide. Agron. J. 105, 657–669 (2013).
    CAS  Article  Google Scholar 

    38.
    Roy, J. et al. Elevated CO2 maintains grassland net carbon uptake under a future heat and drought extreme. Proc. Natl Acad. Sci. U.S.A 113, 6224–6229 (2016).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    39.
    Walker, T. W. N. et al. Microbial temperature sensitivity and biomass change explain soil carbon loss with warming. Nat. Clim. Chang. 8, 885–889 (2018).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    40.
    Hagerty, S. B. et al. Accelerated microbial turnover but constant growth efficiency with warming in soil. Nat. Clim. Chang. 4, 903–906 (2014).
    CAS  Article  Google Scholar 

    41.
    Allison, S. D., Wallenstein, M. D. & Bradford, M. A. Soil-carbon response to warming dependent on microbial physiology. Nat. Geosci. 3, 336–340 (2010).
    CAS  Article  Google Scholar 

    42.
    Lehmeier, C. A., Ballantyne, F. IV, Min, K. & Billings, S. A. Temperature-mediated changes in microbial carbon use efficiency and 13C discrimination. Biogeosciences 13, 3319–3329 (2016).
    CAS  Article  Google Scholar 

    43.
    Fuchslueger, L. et al. Microbial carbon and nitrogen cycling responses to drought and temperature in differently managed mountain grasslands. Soil Biol. Biochem. 135, 144–153 (2019).
    CAS  Article  Google Scholar 

    44.
    Alvarez, G. et al. Catalytic power of enzymes decreases with temperature: New insights for understanding soil C cycling and microbial ecology under warming. Glob. Chang. Biol. 24, 4238–4250 (2018).
    PubMed  Article  Google Scholar 

    45.
    Dijkstra, P. et al. Effect of temperature on metabolic activity of intact microbial communities: Evidence for altered metabolic pathway activity but not for increased maintenance respiration and reduced carbon use efficiency. Soil Biol. Biochem. 43, 2023–2031 (2011).
    CAS  Article  Google Scholar 

    46.
    Manzoni, S. et al. Reviews and syntheses: Carbon use efficiency from organisms to ecosystems – Definitions, theories, and empirical evidence. Biogeosciences 15, 5929–5949 (2018).
    CAS  Article  Google Scholar 

    47.
    Schmidt, S. K. et al. Biogeochemical consequences of rapid microbial turnover and seasonal succession in soil. Ecology 88, 1379–1385 (2007).
    CAS  PubMed  Article  Google Scholar 

    48.
    Blume, E. et al. Surface and subsurface microbial biomass, community structure and metabolic activity as a function of soil depth and season. Appl. Soil Ecol. 20, 171–181 (2002).
    Article  Google Scholar 

    49.
    Regan, K. M. et al. Seasonal controls on grassland microbial biogeography: Are they governed by plants, abiotic properties or both? Soil Biol. Biochem. 71, 21–30 (2014).
    CAS  Article  Google Scholar 

    50.
    Piepho Herndl, M., Pötsch, E. M. & Bahn, M. Designing an experiment with quantitative treatment factors to study the effects of climate change. J. Agron. Crop Sci. 203, 584–592 (2017).
    Article  CAS  Google Scholar 

    51.
    Liu, Y. et al. The optimum temperature of soil microbial respiration: Patterns and controls. Soil Biol. Biochem. 121, 35–42 (2018).
    CAS  Article  Google Scholar 

    52.
    Rustad, L. E. et al. A meta-analysis of the response of soil respiration, net nitrogen mineralization, and aboveground plant growth to experimental ecosystem warming. Oecologia 126, 543–562 (2001).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    53.
    Schwarz, B. et al. Warming alters energetic structure and function but not resilience of soil food webs. Nat. Clim. Chang. 7, 895–900 (2017).
    PubMed  PubMed Central  Article  Google Scholar 

    54.
    Hartley, I. P., Heinemeyer, A. & Ineson, P. Effects of three years of soil warming and shading on the rate of soil respiration: Substrate availability and not thermal acclimation mediates observed response. Glob. Chang. Biol. 13, 1761–1770 (2007).
    Article  Google Scholar 

    55.
    Drigo, B., Kowalchuk, G. A. & van Veen, J. A. Climate change goes underground: effects of elevated atmospheric CO2 on microbial community structure and activities in the rhizosphere. Biol. Fertil. Soils 44, 667–679 (2008).
    Article  Google Scholar 

    56.
    Kowalchuk, G. A. et al. Impacts of 3 years of elevated atmospheric CO2 on rhizosphere carbon flow and microbial community dynamics. Glob. Chang. Biol. 19, 621–636 (2012).
    PubMed  PubMed Central  Google Scholar 

    57.
    Luo, Y., Hui, D. & Zhang, D. Elevated CO2 stimulates net accumulations of carbon and nitrogen in land ecosystems: A meta-analysis. Ecology 87, 53–63 (2006).
    PubMed  Article  PubMed Central  Google Scholar 

    58.
    Pausch, J. & Kuzyakov, Y. Carbon input by roots into the soil: quantification of rhizodeposition from root to ecosystem scale. Glob. Chang. Biol. 24, 1–12 (2018).
    PubMed  Article  PubMed Central  Google Scholar 

    59.
    Carrillo, Y., Dijkstra, F., LeCain, D., Blumenthal, D. & Pendall, E. Elevated CO2 and warming cause interactive effects on soil carbon and shifts in carbon use by bacteria. Ecol. Lett. 21, 1639–1648 (2018).
    PubMed  Article  PubMed Central  Google Scholar 

    60.
    Adair, K. L. et al. Above and belowground community strategies respond to different global change drivers. Sci. Rep. 9, 1–11 (2019).
    CAS  Article  Google Scholar 

    61.
    Deltedesco, E. et al. Trace gas fluxes from managed grassland soil subject to multifactorial climate change manipulation. Appl. Soil Ecol. 137, 1–11 (2019).
    Article  Google Scholar 

    62.
    Vance, E. D., Brookes, P. C. & Jenkinson, D. S. An extraction method for measuring soil microbial biomass C. Soil Biol. Biochem. 19, 703–707 (1987).
    CAS  Article  Google Scholar 

    63.
    Spohn, M., Klaus, K., Wanek, W. & Richter, A. Microbial carbon use efficiency and biomass turnover times depending on soil depth – Implications for carbon cycling. Soil Biol. Biochem. 96, 74–81 (2016).
    CAS  Article  Google Scholar 

    64.
    Pinheiro, J., Bates, D., Debroy, S., Sarkar, D. & R-core Team. nlme: linear and nonlinear mixed effects models. R package version 3 (2019).

    65.
    Galecki, A. & Burzykowski, T. Linear Mixed Effects Models Using R.: A Step-by-Step Approach. https://doi.org/10.1007/978-1-4614-3900-4 (2013).

    66.
    Piepho, H. P. & Edmondson, R. N. A tutorial on the statistical analysis of factorial experiments with qualitative and quantitative treatment factor levels. J. Agron. Crop Sci. 204, 429–455 (2018).
    Article  Google Scholar 

    67.
    Box, G. E. P. & Jenkins, G. M. Time Series Analysis: Forecasting and Control. (Holden-Day, San Francisco, CA, US, 1976).

    68.
    Finerty, J. P. The Population Ecology of Cycles in Small Mammals. (Yale University Press, New Haven and London, UK, 1980).

    69.
    Nisbet, R. M. & Gurney, W. S. C. Modelling Fluctuating Populations. (John Wiley & Sons, New York, US, 1982).

    70.
    Turchin, P. & Taylor, D. Complex dynamics in ecological time series. Ecology 73, 289–305 (1992).
    Article  Google Scholar 

    71.
    Davies, N. & Chatfield, C. The Analysis of Time Series: An Introduction. The Mathematical Gazette vol. 74 (Chapman and Hall, 2007).

    72.
    Kozak, M. & Piepho, H. P. What’s normal anyway? Residual plots are more telling than significance tests when checking ANOVA assumptions. J. Agron. Crop Sci. 204, 86–98 (2018).
    Article  Google Scholar 

    73.
    Lenth, R. V. Response-surface methods in R, using rsm. J. Stat. Softw. 32, 1–21 (2010). More

  • in

    The importance of genomic variation for biodiversity, ecosystems and people

    1.
    Scheffers, B. R. et al. The broad footprint of climate change from genes to biomes to people. Science 354, aaf7671 (2016).
    PubMed  Article  CAS  PubMed Central  Google Scholar 
    2.
    Cadotte, M. W., Carscadden, K. & Mirotchnick, N. Beyond species: functional diversity and the maintenance of ecological processes and services. J. Appl. Ecol. 48, 1079–1087 (2011).
    Article  Google Scholar 

    3.
    Faith, D. P. et al. Evosystem services: an evolutionary perspective on the links between biodiversity and human well-being. Curr. Opin. Environ. Sustain. 2, 66–74 (2010).
    Article  Google Scholar 

    4.
    Mimura, M. et al. Understanding and monitoring the consequences of human impacts on intraspecific variation. Evol. Appl. 10, 121–139 (2017).
    PubMed  Article  PubMed Central  Google Scholar 

    5.
    Rudman, S. M. et al. What genomic data can reveal about eco-evolutionary dynamics. Nat. Ecol. Evol. 2, 9–15 (2018).
    PubMed  Article  PubMed Central  Google Scholar 

    6.
    Des Roches, S. et al. The ecological importance of intraspecific variation. Nat. Ecol. Evol. 2, 57–64 (2018).
    PubMed  Article  PubMed Central  Google Scholar 

    7.
    Therkildsen, N. O. et al. Contrasting genomic shifts underlie parallel phenotypic evolution in response to fishing. Science 365, 487–490 (2019).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    8.
    Crutsinger, G. M. et al. Plant genotypic diversity predicts community structure and governs an ecosystem process. Science 313, 966–968 (2006).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    9.
    Leigh, D. M., Hendry, A. P., Vázquez-Domínguez, E. & Friesen, V. L. Estimated six per cent loss of genetic variation in wild populations since the industrial revolution. Evol. Appl. 12, 1505–1512 (2019).
    PubMed  PubMed Central  Article  Google Scholar 

    10.
    Boeuf, G. Marine biodiversity characteristics. C. R. Biol. 334, 435–440 (2011).
    PubMed  Article  Google Scholar 

    11.
    Loss, S. R., Terwilliger, L. A. & Peterson, A. C. Assisted colonization: integrating conservation strategies in the face of climate change. Biol. Conserv. 144, 92–100 (2011).
    Article  Google Scholar 

    12.
    Aitken, S. N. & Whitlock, M. C. Assisted gene flow to facilitate local adaptation to climate change. Annu. Rev. Ecol. Evol. Syst. 44, 367–388 (2013).
    Article  Google Scholar 

    13.
    Witzenberger, K. A. & Hochkirch, A. Ex situ conservation genetics: a review of molecular studies on the genetic consequences of captive breeding programmes for endangered animal species. Biodivers. Conserv. 20, 1843–1861 (2011).
    Article  Google Scholar 

    14.
    Novak, B. J. De-extinction. Genes 9, 548 (2018).
    PubMed Central  Article  CAS  PubMed  Google Scholar 

    15.
    Muir, W. M. et al. Genome-wide assessment of worldwide chicken SNP genetic diversity indicates significant absence of rare alleles in commercial breeds. Proc. Natl Acad. Sci. USA 105, 17312–17317 (2008).
    CAS  PubMed  Article  Google Scholar 

    16.
    Beck, M. W. et al. The global flood protection savings provided by coral reefs. Nat. Commun 9, 2186 (2018).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    17.
    Millennium Ecosystem Assessment. Ecosystems and Human Well-Being: Synthesis (Island Press, 2005).

    18.
    Díaz, S. et al. The IPBES conceptual framework — connecting nature and people. Curr. Opin. Environ. Sustain. 14, 1–16 (2015).
    Article  Google Scholar 

    19.
    Hendry, A. P. Eco-evolutionary Dynamics (Princeton Univ. Press, 2017).

    20.
    Whitham, T. G. et al. Community and ecosystem genetics: a consequence of the extended phenotype. Ecology 84, 559–573 (2003).
    Article  Google Scholar 

    21.
    Larkin, A. A. & Martiny, A. C. Microdiversity shapes the traits, niche space, and biogeography of microbial taxa. Environ. Microbiol. Rep. 9, 55–70 (2017).
    CAS  PubMed  Article  Google Scholar 

    22.
    Rodríguez-Verdugo, A., Buckley, J. & Stapley, J. The genomic basis of eco-evolutionary dynamics. Mol. Ecol. 26, 1456–1464 (2017).
    PubMed  Article  CAS  Google Scholar 

    23.
    Chen, E., Huang, X., Tian, Z., Wing, R. A. & Han, B. The genomics of oryza species provides insights into rice domestication and heterosis. Annu. Rev. Plant. Biol. 70, 639–665 (2019).
    CAS  PubMed  Article  Google Scholar 

    24.
    Bailey, J. K. et al. Beavers as molecular geneticists: a genetic basis to the foraging of an ecosystem engineer. Ecology 85, 603–608 (2004).
    Article  Google Scholar 

    25.
    Whitham, T. G. et al. A framework for community and ecosystem genetics: from genes to ecosystems. Nat. Rev. Genet. 7, 510–523 (2006).
    CAS  PubMed  Article  Google Scholar 

    26.
    Lee, S. M., Jellison, T. & Alper, H. S. Systematic and evolutionary engineering of a xylose isomerase-based pathway in Saccharomyces cerevisiae for efficient conversion yields. Biotechnol. Biofuels 7, 1–8 (2014).
    Article  CAS  Google Scholar 

    27.
    Gleizer, S. et al. Conversion of Escherichia coli to generate all biomass carbon from CO2. Cell 179, 1255–1263.e12 (2019).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    28.
    Zhu, Y. et al. Genetic diversity and disease control in rice. Nature 406, 718–722 (2000).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    29.
    King, K. C. & Lively, C. M. Does genetic diversity limit disease spread in natural host populations. Heredity 109, 199–203 (2012).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    30.
    Robinson, S. J., Samuel, M. D., Johnson, C. J., Adams, M. & McKenzie, D. I. Emerging prion disease drives host selection in a wildlife population. Ecol. Appl. 22, 1050–1059 (2012).
    PubMed  Article  Google Scholar 

    31.
    Springbett, A. J., MacKenzie, K., Woolliams, J. A. & Bishop, S. C. The contribution of genetic diversity to the spread of infectious diseases in livestock populations. Genetics 165, 1465–1474 (2003).
    CAS  PubMed  PubMed Central  Google Scholar 

    32.
    McGranahan, N. & Swanton, C. Clonal heterogeneity and tumor evolution: past, present, and the future. Cell 168, 613–628 (2017).
    CAS  PubMed  Article  Google Scholar 

    33.
    Heap, I. M. The occurrence of herbicide-resistant weeds worldwide. Pestic. Sci. 51, 235–243 (1997).
    CAS  Article  Google Scholar 

    34.
    Whalon, M. E., Mota-Sanchez, D. & Hollingworth, R. M. Global Pesticide Resistance in Arthropods (CABI, 2008).

    35.
    Hartley, C. J. et al. Amplification of DNA from preserved specimens shows blowflies were preadapted for the rapid evolution of insecticide resistance. Proc. Natl Acad. Sci. USA 103, 8757–8762 (2006).
    CAS  PubMed  Article  Google Scholar 

    36.
    Dunlop, E. S., Eikeset, A. M. & Stenseth, N. C. From genes to populations: how fisheries-induced evolution alters stock productivity. Ecol. Appl. 25, 1860–1868 (2015).
    PubMed  Article  Google Scholar 

    37.
    Waples, R. S. & Audzijonyte, A. Fishery-induced evolution provides insights into adaptive responses of marine species to climate change. Front. Ecol. Environ. 14, 217–224 (2016).
    Article  Google Scholar 

    38.
    Food and Agriculture Organization of the United Nations. Review of the state of world marine fishery resources (FAO, 2011).

    39.
    Darimont, C. T. et al. Human predators outpace other agents of trait change in the wild. Proc. Natl Acad. Sci. USA 106, 952–954 (2009).
    CAS  PubMed  Article  Google Scholar 

    40.
    Philipp, D. P. et al. Fisheries-induced evolution in Largemouth Bass: linking vulnerability to angling, parental care, and fitness. Am. Fish. Soc. Symp. 82, 223–234 (2015).
    Google Scholar 

    41.
    Philipp, D. P. et al. Selection for vulnerability to angling in largemouth bass. Trans. Am. Fish. Soc. 138, 189–199 (2009).
    Article  Google Scholar 

    42.
    Pigeon, G., Festa-Bianchet, M., Coltman, D. W. & Pelletier, F. Intense selective hunting leads to artificial evolution in horn size. Evol. Appl. 9, 521–530 (2016).
    PubMed  PubMed Central  Article  Google Scholar 

    43.
    Faith, D. P. Conservation evaluation and phylogenetic diversity. Biol. Conserv. 61, 1–10 (1992).
    Article  Google Scholar 

    44.
    Carlson, S. M., Cunningham, C. J. & Westley, P. A. H. Evolutionary rescue in a changing world. Trends Ecol. Evol. 29, 521–530 (2014).
    PubMed  Article  PubMed Central  Google Scholar 

    45.
    Hendry, A. P., Schoen, D. J., Wolak, M. E. & Reid, J. M. The contemporary evolution of fitness. Annu. Rev. Ecol. Evol. Syst. 49, 457–476 (2018).
    Article  Google Scholar 

    46.
    Dakos, V. et al. Ecosystem tipping points in an evolving world. Nat. Ecol. Evol. 3, 355–362 (2019).
    PubMed  Article  PubMed Central  Google Scholar 

    47.
    Souza, F. F. C. et al. Uncovering prokaryotic biodiversity within aerosols of the pristine Amazon forest. Sci. Total Environ. 688, 83–86 (2019).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    48.
    Suffredini, I. B., Barradas Paciencia, M. L., Varella, A. D. & Younes, R. N. Antibacterial activity of Brazilian Amazon plant extracts. Braz. J. Infect. Dis. 10, 400–402 (2006).
    PubMed  Article  PubMed Central  Google Scholar 

    49.
    Blanco-Salas, J., Gutiérrez-García, L., Labrador-Moreno, J. & Ruiz-Téllez, T. Wild plants potentially used in human food in the protected area ‘Sierra Grande de Hornachos’ of Extremadura (Spain). Sustainability 11, 456 (2019).
    Article  Google Scholar 

    50.
    Sam Ma, Z., Li, L. & Zhang, Y. P. Defining individual-level genetic diversity and similarity profiles. Sci. Rep. 10, 5805 (2020).
    Article  CAS  Google Scholar 

    51.
    Avolio, M. L., Beaulieu, J. M., Lo, E. Y. Y. & Smith, M. D. Measuring genetic diversity in ecological studies. Plant. Ecol. 213, 1105–1115 (2012).
    Article  Google Scholar 

    52.
    Günther, T. & Coop, G. Robust identification of local adaptation from allele frequencies. Genetics 195, 205–220 (2013).
    PubMed  PubMed Central  Article  Google Scholar 

    53.
    Booker, T. R., Jackson, B. C. & Keightley, P. D. Detecting positive selection in the genome. BMC Biol. 15, 1–10 (2017).
    Article  CAS  Google Scholar 

    54.
    Dawkins, R. The Extended Phenotype – The Gene as the Unit of Selection (Oxford Univ. Press, 1983).

    55.
    Shuster, S. M., Lonsdorf, E. V., Wimp, G. M., Bailey, J. K. & Whitham, T. G. Community heritability measures the evolutionary consequences of indirect genetic effects on community structure. Evolution 60, 991–1003 (2006).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    56.
    Lynch, M. & Walsh, B. Genetics and Analysis of Quantitative Traits (Sinauer Associates, 1998).

    57.
    Doudna, J. A. & Charpentier, E. The new frontier of genome engineering with CRISPR-Cas9. Science 346, 1258096 (2014).
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    58.
    Knott, G. J. & Doudna, J. A. CRISPR-Cas guides the future of genetic engineering. Science 361, 866–869 (2018).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    59.
    Skovmand, L. H. et al. Keystone genes. Trends Ecol. Evol. 33, 689–700 (2018).
    PubMed  Article  PubMed Central  Google Scholar 

    60.
    Pregitzer, C. C., Bailey, J. K., Hart, S. C. & Schweitzer, J. A. Soils as agents of selection: feedbacks between plants and soils alter seedling survival and performance. Evol. Ecol. 24, 1045–1059 (2010).
    Article  Google Scholar 

    61.
    Bailey, J. K. et al. From genes to ecosystems: a synthesis of the effects of plant genetic factors across levels of organization. Phil. Trans. R. Soc. B 364, 1607–1616 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    62.
    Davies, C., Ellis, C. J., Iason, G. R. & Ennos, R. A. Genotypic variation in a foundation tree (Populus tremula L.) explains community structure of associated epiphytes. Biol. Lett. 10, 20140190 (2014).
    PubMed  PubMed Central  Article  Google Scholar 

    63.
    Thompson, T. Q. et al. Anthropogenic habitat alteration leads to rapid loss of adaptive variation and restoration potential in wild salmon populations. Proc. Natl Acad. Sci. USA 116, 177–186 (2019).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    64.
    Ford, M. D. et al. Reviewing and synthesizing the state of the science regarding associations between adult run timing and specific genotypes in Chinook salmon and steelhead (US Department of Commerce, 2020).

    65.
    Leroy, C. J. et al. Salmon carcasses influence genetic linkages between forests and streams. Can. J. Fish. Aquat. Sci. 73, 910–920 (2016).
    Article  Google Scholar 

    66.
    Crutsinger, G. M. et al. Testing a ‘genes-to-ecosystems’ approach to understanding aquatic-terrestrial linkages. Mol. Ecol. 23, 5888–5903 (2014).
    PubMed  Article  PubMed Central  Google Scholar 

    67.
    Lewontin, R. C. The Genetic Basis of Evolutionary Change (Columbia Univ. Press, 1974).

    68.
    Csilléry, K., Rodríguez-Verdugo, A., Rellstab, C. & Guillaume, F. Detecting the genomic signal of polygenic adaptation and the role of epistasis in evolution. Mol. Ecol. 27, 606–612 (2018).
    PubMed  Article  PubMed Central  Google Scholar 

    69.
    Zytynska, S. E., Fleming, S., Tétard-Jones, C., Kertesz, M. A. & Preziosi, R. F. Community genetic interactions mediate indirect ecological effects between a parasitoid wasp and rhizobacteria. Ecology 91, 1563–1568 (2010).
    PubMed  Article  PubMed Central  Google Scholar 

    70.
    Carroll, S. P., Dingle, H. & Famula, T. R. Rapid appearance of epistasis during adaptive divergence following colonization. Proc. R. Soc. Lond. B 270, S80–S83 (2003).
    Article  Google Scholar 

    71.
    Carroll, S. P. et al. And the beak shall inherit – evolution in response to invasion. Ecol. Lett. 8, 944–951 (2005).
    Article  Google Scholar 

    72.
    Doust, A. N. et al. Beyond the single gene: how epistasis and gene-byenvironment effects influence crop domestication. Proc. Natl Acad. Sci. USA 111, 6178–6183 (2014).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    73.
    Wellenreuther, M., Mérot, C., Berdan, E. & Bernatchez, L. Going beyond SNPs: the role of structural genomic variants in adaptive evolution and species diversification. Mol. Ecol. 28, 1203–1209 (2019).
    PubMed  Article  PubMed Central  Google Scholar 

    74.
    Ayala, D. et al. Association mapping desiccation resistance within chromosomal inversions in the African malaria vector Anopheles gambiae. Mol. Ecol. 28, 1333–1342 (2019).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    75.
    Christmas, M. J. et al. Chromosomal inversions associated with environmental adaptation in honeybees. Mol. Ecol. 28, 1358–1374 (2019).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    76.
    Kess, T. et al. A migration-associated supergene reveals loss of biocomplexity in Atlantic cod. Sci. Adv. 5, eaav2461 (2019).
    PubMed  PubMed Central  Article  Google Scholar 

    77.
    Berg, P. R. et al. Trans-oceanic genomic divergence of Atlantic cod ecotypes is associated with large inversions. Heredity 119, 418–428 (2017).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    78.
    Frank, K. T., Petrie, B., Choi, J. S. & Leggett, W. C. Ecology: trophic cascades in a formerly cod-dominated ecosystem. Science 308, 1621–1623 (2005).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    79.
    Prunier, J. et al. Gene copy number variations involved in balsam poplar (Populus balsamifera L.) adaptive variations. Mol. Ecol. 28, 1476–1490 (2019).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    80.
    Youngson, N. A. & Whitelaw, E. Transgenerational epigenetic effects. Annu. Rev. Genomics Hum. Genet. 9, 233–257 (2008).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    81.
    Miura, K. et al. OsSPL14 promotes panicle branching and higher grain productivity in rice. Nat. Genet. 42, 545–549 (2010).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    82.
    Ong-Abdullah, M. et al. Loss of Karma transposon methylation underlies the mantled somaclonal variant of oil palm. Nature 525, 533–537 (2015).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    83.
    Le Luyer, J. et al. Parallel epigenetic modifications induced by hatchery rearing in a Pacific salmon. Proc. Natl Acad. Sci. USA 114, 12964–12969 (2017).
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    84.
    Baerwald, M. R. et al. Migration-related phenotypic divergence is associated with epigenetic modifications in rainbow trout. Mol. Ecol. 25, 1785–1800 (2016).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    85.
    Oke, K. B. et al. Recent declines in salmon body size impact ecosystems and fisheries. Nat. Commun. 11, 4155 (2020).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    86.
    Davies, T. J., Urban, M. C., Rayfield, B., Cadotte, M. W. & Peres-Neto, P. R. Deconstructing the relationships between phylogenetic diversity and ecology: a case study on ecosystem functioning. Ecology 97, 2212–2222 (2016).
    PubMed  Article  PubMed Central  Google Scholar 

    87.
    Cadotte, M. W. Phylogenetic diversity-ecosystem function relationships are insensitive to phylogenetic edge lengths. Funct. Ecol. 29, 718–723 (2015).
    Article  Google Scholar 

    88.
    Cadotte, M. W. Experimental evidence that evolutionarily diverse assemblages result in higher productivity. Proc. Natl Acad. Sci. USA 110, 8996–9000 (2013).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    89.
    MacIvor, J. S. et al. Manipulating plant phylogenetic diversity for green roof ecosystem service delivery. Evol. Appl. 11, 2014–2024 (2018).
    PubMed  PubMed Central  Article  Google Scholar 

    90.
    Clark, J. S., Scher, C. L. & Swift, M. The emergent interactions that govern biodiversity change. Proc. Natl Acad. Sci. USA 117, 17074–17083 (2020).
    PubMed  Article  Google Scholar 

    91.
    Crutsinger, G. M. A community genetics perspective: opportunities for the coming decade. N. Phytol. 210, 65–70 (2016).
    Article  Google Scholar 

    92.
    Zuppinger-Dingley, D. et al. Selection for niche differentiation in plant communities increases biodiversity effects. Nature 515, 108–111 (2014).
    CAS  PubMed  Article  Google Scholar 

    93.
    van Moorsel, S. J. et al. Community evolution increases plant productivity at low diversity. Ecol. Lett. 21, 128–137 (2018).
    PubMed  Article  Google Scholar 

    94.
    Wade, M. J. The co-evolutionary genetics of ecological communities. Nat. Rev. Genet. 8, 185–195 (2007).
    CAS  PubMed  Article  Google Scholar 

    95.
    Genung, M. A., Bailey, J. K. & Schweitzer, J. A. Welcome to the neighbourhood: Interspecific genotype by genotype interactions in Solidago influence above- and belowground biomass and associated communities. Ecol. Lett. 15, 65–73 (2012).
    PubMed  Article  Google Scholar 

    96.
    Genung, M. A., Bailey, J. K. & Schweitzer, J. A. The afterlife of interspecific indirect genetic effects: genotype interactions alter litter quality with consequences for decomposition and nutrient dynamics. PLoS ONE 8, e53718 (2013).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    97.
    Lankau, R. A. & Nodurft, R. N. An exotic invader drives the evolution of plant traits that determine mycorrhizal fungal diversity in a native competitor. Mol. Ecol. 22, 5472–5485 (2013).
    PubMed  Article  Google Scholar 

    98.
    Lankau, R. A., Nuzzo, V., Spyreas, G. & Davis, A. S. Evolutionary limits ameliorate the negative impact of an invasive plant. Proc. Natl Acad. Sci. USA 107, 1253 (2010).
    CAS  Article  Google Scholar 

    99.
    Lankau, R. A. Coevolution between invasive and native plants driven by chemical competition and soil biota. Proc. Natl Acad. Sci. USA 109, 11240–11245 (2012).
    CAS  PubMed  Article  Google Scholar 

    100.
    Lankau, R. A., Bauer, J. T., Anderson, M. R. & Anderson, R. C. Long-term legacies and partial recovery of mycorrhizal communities after invasive plant removal. Biol. Invasions 16, 1979–1990 (2014).
    Article  Google Scholar 

    101.
    Miller, E. T., Svanbäck, R. & Bohannan, B. J. M. Microbiomes as metacommunities: understanding host-associated microbes through metacommunity ecology. Trends Ecol. Evol. 33, 926–935 (2018).
    PubMed  Article  PubMed Central  Google Scholar 

    102.
    Pearse, D. E., Miller, M. R., Abadía-Cardoso, A. & Garza, J. C. Rapid parallel evolution of standing variation in a single, complex, genomic region is associated with life history in steelhead/rainbow trout. Proc. R. Soc. B 281, 20140012 (2014).
    PubMed  Article  PubMed Central  Google Scholar 

    103.
    Narum, S. R., Genova, A. D., Micheletti, S. J. & Maass, A. Genomic variation underlying complex life-history traits revealed by genome sequencing in Chinook salmon. Proc. R. Soc. B 285, 20180935 (2018).
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    104.
    Prince, D. J. et al. The evolutionary basis of premature migration in Pacific salmon highlights the utility of genomics for informing conservation. Sci. Adv. 3, e1603198 (2017).
    PubMed  PubMed Central  Article  Google Scholar 

    105.
    Rey, O. et al. Linking epigenetics and biological conservation: towards a conservation epigenetics perspective. Funct. Ecol. 34, 414–427 (2020).
    Article  Google Scholar 

    106.
    Hu, J. & Barrett, R. D. H. Epigenetics in natural animal populations. J. Evol. Biol. 30, 1612–1632 (2017).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    107.
    Herrera, C. M., Medrano, M., Pérez, R., Bazaga, P. & Alonso, C. Within-plant heterogeneity in fecundity and herbivory induced by localized DNA hypomethylation in the perennial herb Helleborus foetidus. Am. J. Bot. 106, 798–806 (2019).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    108.
    Cohen, S. N., Chang, A. C. Y., Boyer, H. W. & Helling, R. B. Construction of biologically functional bacterial plasmids in vitro (R factor/restriction enzyme/transformation/endonuclease/antibiotic resistance). Proc. Natl Acad. Sci. USA 70, 3240–3244 (1973).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    109.
    Porteus, M. H. & Carroll, D. Gene targeting using zinc finger nucleases. Nat. Biotechnol. 23, 967–973 (2005).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    110.
    Urnov, F. D., Rebar, E. J., Holmes, M. C., Zhang, H. S. & Gregory, P. D. Genome editing with engineered zinc finger nucleases. Nat. Rev. Genet. 11, 636–646 (2010).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    111.
    Joung, J. K. & Sander, J. D. TALENs: a widely applicable technology for targeted genome editing. Nat. Rev. Mol. Cell Biol. 14, 49–55 (2013).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    112.
    Burt, A. Site-specific selfish genes as tools for the control and genetic engineering of natural populations. Proc. R. Soc. Lond. B 270, 921–928 (2003).
    CAS  Article  Google Scholar 

    113.
    Zhang, Y., Massel, K., Godwin, I. D. & Gao, C. Applications and potential of genome editing in crop improvement. Genome Biol. 19, 210 (2018).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    114.
    Charu, V. & Kaplan, D. L. Silk as a biomaterial. Prog. Polym. Sci. 100, 130–134 (2012).
    Google Scholar 

    115.
    Mosa, K. A., Saadoun, I., Kumar, K., Helmy, M. & Dhankher, O. P. Potential biotechnological strategies for the cleanup of heavy metals and metalloids. Front. Plant Sci. 7, 303 (2016).
    PubMed  PubMed Central  Article  Google Scholar 

    116.
    Champer, J., Buchman, A. & Akbari, O. S. Cheating evolution: engineering gene drives to manipulate the fate of wild populations. Nat. Rev. Genet. 17, 146–159 (2016).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    117.
    Rode, N. O., Estoup, A., Bourguet, D., Courtier-Orgogozo, V. & Débarre, F. Population management using gene drive: molecular design, models of spread dynamics and assessment of ecological risks. Conserv. Genet. 20, 671–690 (2019).
    CAS  Article  Google Scholar 

    118.
    Esvelt, K. M. & Gemmell, N. J. Conservation demands safe gene drive. PLoS Biol. 15, 1–8 (2017).
    Article  CAS  Google Scholar 

    119.
    Phuc, H. et al. Late-acting dominant lethal genetic systems and mosquito control. BMC Biol. 5, 11 (2007).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    120.
    Campbell, K. J. et al. in Island Invasives: Scaling up to Meet the Challenge (eds Veitch, C. R., Clout, M. N., Martin, A. R., Russel, J. C. & West, C. J.) 6–14 (IUCN, 2019).

    121.
    Sherkow, J. S. & Greely, H. T. What if extinction is not forever? Science 340, 32–33 (2013).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    122.
    Otoupal, P. B., Cordell, W. T., Bachu, V., Sitton, M. J. & Chatterjee, A. Multiplexed deactivated CRISPR-Cas9 gene expression perturbations deter bacterial adaptation by inducing negative epistasis. Commun. Biol. 1, 129 (2018).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    123.
    Kraft, K. et al. Deletions, inversions, duplications: engineering of structural variants using CRISPR/Cas in mice. Cell Rep. 10, 833–839 (2015).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    124.
    Springer, N. M. & Schmitz, R. J. Exploiting induced and natural epigenetic variation for crop improvement. Nat. Rev. Genet. 18, 563–575 (2017).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    125.
    Reinders, J. et al. Compromised stability of DNA methylation and transposon immobilization in mosaic Arabidopsis epigenomes. Genes Dev. 23, 939–950 (2009).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    126.
    Carrière, Y., Crowder, D. W. & Tabashnik, B. E. Evolutionary ecology of insect adaptation to Bt crops. Evol. Appl. 3, 561–573 (2010).
    PubMed  PubMed Central  Article  Google Scholar 

    127.
    Fish, D. & Carpenter, S. R. Leaf litter and larval mosquito dynamics in tree-hole ecosystems. Ecology 63, 283–288 (1982).
    Article  Google Scholar 

    128.
    Kraus, J. M. & Vonesh, J. R. Fluxes of terrestrial and aquatic carbon by emergent mosquitoes: a test of controls and implications for cross-ecosystem linkages. Oecologia 170, 1111–1122 (2012).
    PubMed  Article  PubMed Central  Google Scholar 

    129.
    Sheehan, S. & Song, Y. S. Deep learning for population genetic inference. PLoS Comput. Biol. 12, e1004845 (2016).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    130.
    Schrider, D. R. & Kern, A. D. Supervised machine learning for population genetics: a new paradigm. Trends Genet. 34, 301–312 (2018).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    131.
    Christin, S., Hervet, É. & Lecomte, N. Applications for deep learning in ecology. Methods Ecol. Evol. 10, 1632–1644 (2019).
    Article  Google Scholar 

    132.
    Desjardins-Proulx, P., Laigle, I., Poisot, T. & Gravel, D. Ecological interactions and the Netflix problem. PeerJ 2017, e3644 (2017).
    Article  Google Scholar 

    133.
    Ruffley, M., Peterson, K., Week, B., Tank, D. C. & Harmon, L. J. Identifying models of trait-mediated community assembly using random forests and approximate Bayesian computation. Dep. Biol. Sci. https://doi.org/10.1002/ece3.5773 (2019).
    Article  Google Scholar 

    134.
    Laikre, L. et al. Neglect of genetic diversity in implementation of the convention on biological diversity: conservation in practice and policy. Conserv. Biol. 24, 86–88 (2010).
    PubMed  Article  PubMed Central  Google Scholar 

    135.
    Hoban, S. et al. Genetic diversity targets and indicators in the CBD post-2020 Global Biodiversity Framework must be improved. Biol. Conserv. 248, 108654 (2020).
    Article  Google Scholar 

    136.
    Meyer, P. et al. Endogenous and environmental factors influence 35S promoter methylation of a maize A1 gene construct in transgenic petunia and its colour phenotype. Mol. Gen. Genet. 231, 345–352 (1992).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    137.
    Morandin, L. A. & Winston, M. L. Wild bee abundance and seed production in conventional, organic, and genetically modified canola. Ecol. Appl. 15, 871–881 (2005).
    Article  Google Scholar 

    138.
    Axelsson, E. P. et al. Leaf litter from insect-resistant transgenic trees causes changes in aquatic insect community composition. J. Appl. Ecol. 48, 1472–1479 (2011).
    Article  Google Scholar 

    139.
    Axelsson, E. P., Hjältén, J. & LeRoy, C. J. Performance of insect-resistant Bacillus thuringiensis (Bt)-expressing aspens under semi-natural field conditions including natural herbivory in Sweden. For. Ecol. Manage. 264, 167–171 (2012).
    Article  Google Scholar 

    140.
    Sundström, L. F., Lõhmus, M., Tymchuk, W. E. & Devlin, R. H. Gene-environment interactions influence ecological consequences of transgenic animals. Proc. Natl Acad. Sci. USA 104, 3889–3894 (2007).
    PubMed  Article  PubMed Central  Google Scholar 

    141.
    Sundström, L. F., Lôhmus, M., Johnsson, J. I. & Devlin, R. H. Growth hormone transgenic salmon pay for growth potential with increased predation mortality. Proc. R. Soc. Lond. B 271, 350–352 (2004).
    Article  Google Scholar 

    142.
    Bodbyl Roels, S. A. & Kelly, J. K. Rapid evolution caused by pollinator loss in Mimulus guttatus. Evolution 65, 2541–2552 (2011).
    PubMed Central  Article  Google Scholar 

    143.
    Cheptou, P. O., Carrue, O., Rouifed, S. & Cantarel, A. Rapid evolution of seed dispersal in an urban environment in the weed Crepis sancta. Proc. Natl Acad. Sci. USA 105, 3796–3799 (2008).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    144.
    Polymenakou, P. N. Atmosphere: a source of pathogenic or beneficial microbes? Atmosphere 3, 87–102 (2012).
    Article  Google Scholar 

    145.
    Collins, S. Many possible worlds: expanding the ecological scenarios in experimental evolution. Evol. Biol. 38, 3–14 (2011).
    Article  Google Scholar 

    146.
    Archer, D. et al. Atmospheric lifetime of fossil fuel carbon dioxide. Annu. Rev. Earth Planet. Sci. 37, 117–134 (2009).
    CAS  Article  Google Scholar 

    147.
    Sunday, J. M. et al. Evolution in an acidifying ocean. Trends Ecol. Evol. 29, 117–125 (2014).
    PubMed  Article  PubMed Central  Google Scholar 

    148.
    Harmon, L. J. et al. Evolutionary diversification in stickleback affects ecosystem functioning. Nature 458, 1167–1170 (2009).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    149.
    Hairston, N. G. et al. Rapid evolution revealed by dormant eggs. Nature 401, 446–446 (1999).
    Article  Google Scholar 

    150.
    Bothe, H. & Słomka, A. Divergent biology of facultative heavy metal plants. J. Plant Physiol. 219, 45–61 (2017).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    151.
    Reusch, T. B. H., Ehlers, A., Hammerli, A. & Worm, B. Ecosystem recovery after climatic extremes enhanced by genotypic diversity. Proc. Natl Acad. Sci. USA 102, 2826–2831 (2005).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    152.
    Crutsinger, G. M., Souza, L. & Sanders, N. J. Intraspecific diversity and dominant genotypes resist plant invasions. Ecol. Lett. 11, 16–23 (2008).
    PubMed  PubMed Central  Google Scholar 

    153.
    Pelz, H. J. et al. The genetic basis of resistance to anticoagulants in rodents. Genetics 170, 1839–1847 (2005).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    154.
    National Research Council. Materials Research to Meet 21st Century Defense Needs (National Academies Press, 2003).

    155.
    Hutchison, W. D. et al. Areawide suppression of European corn borer with Bt maize reaps savings to non-Bt maize growers. Science 330, 222–225 (2010).
    CAS  PubMed  Article  Google Scholar 

    156.
    Leale, A. M. & Kassen, R. The emergence, maintenance, and demise of diversity in a spatially variable antibiotic regime. Evol. Lett. 2, 134–143 (2018).
    PubMed  PubMed Central  Article  Google Scholar 

    157.
    Grant, P. R. & Grant, B. R. Unpredictable evolution in a 30-year study of Darwin’s finches. Science 296, 707–711 (2002).
    CAS  PubMed  Article  Google Scholar 

    158.
    Grant, P. R. & Grant, B. R. Evolution of character displacement in Darwin’ s finches. Science 313, 224–226 (2006).
    CAS  PubMed  Article  Google Scholar 

    159.
    Lamichhaney, S. et al. Evolution of Darwin’s finches and their beaks revealed by genome sequencing. Nature 518, 371–375 (2015).
    CAS  PubMed  Article  Google Scholar 

    160.
    Constantino, V. Instinct extinct: the great pacific flyway. Leonardo 52, 5–11 (2018).
    Article  Google Scholar 

    161.
    Lewis, B., Grant, W. S., Brenner, R. E. & Hamazaki, T. Changes in size and age of chinook salmon Oncorhynchus tshawytscha returning to Alaska. PLoS ONE 10, 132872 (2015).
    Google Scholar 

    162.
    Schweitzer, J. A. et al. From genes to ecosystems: the genetic basis of condensed tannins and their role in nutrient regulation in a Populus model system. Ecosystems 11, 1005–1020 (2008).
    CAS  Article  Google Scholar 

    163.
    Díaz, S. et al. Assessing nature’s contributions to people. Science 359, 270–272 (2018). Introduction to the key concept of NCP.
    PubMed  Article  Google Scholar  More

  • in

    A bacterial endophyte exploits chemotropism of a fungal pathogen for plant colonization

    Fungal, yeast, and bacterial strains and culture conditions
    Rahnella aquatilis strain 36 (Ra36) was previously isolated from chickpea roots23. Escherichia coli and Ra36 were routinely grown at 37 °C or 28 °C, respectively, in Luria Bertani (LB) medium (5 g/L yeast extract, 10 g/L peptone, 10 g/L NaCl, pH 7.0) containing kanamycin sulfate (50 µg/mL), ampicillin (250 µg/mL) or carbenicillin (100 µg/mL), where appropriate. Saccharomyces cerevisiae was grown in yeast extract peptone dextrose (YPD; 1% Bacto yeast extract, 2% bacto peptone, and 2% dextrose), except for the selection of recombinants which was carried out in yeast nitrogen base medium (Sigma-Aldrich, Madrid, Spain) containing the appropriate amino-acid supplements.
    F. oxysporum f. sp. lycopersici race 2 isolate 4287 (FGSC9935; Fol) or previously reported Fol mutants either lacking the ste2 gene6 or constitutively expressing the GFP7 were used throughout the work. For microconidia production, cultures were grown in potato dextrose broth (PDB) at 28 °C with shaking at 170 rpm38. For phenotypic analysis of Fol colony growth, serial dilutions (106, 105, 104, and 103 mL−1) of freshly obtained microconidia were spotted onto YPD agar plates supplemented or not with 0.8% (w/v) GlcA (Sigma-Aldrich), incubated at 28 °C for 3 d and imaged. Experiments included three replicates and were performed at least three times with similar results. Microconidial and bacterial cell suspensions were routinely stored at −80 °C with 30% glycerol.
    Bioinformatic analysis
    Identification of gcd and fliC gene orthologs from different R. aquatilis strains was performed with the BLAST algorithm39, using the R. aquatilis HX2 gcd and fliC genes as baits. Nucleotide sequences encompassing the entire gcd and fliC genes and comprising 1.5-kb upstream and downstream of the retrieved ORFs were aligned using the BioEdit 7.0.0 software (Ibis Bioscience, Carlsbad, CA, USA). Primer pairs Gcd7/Gcd8 and Flic7/Flic8 were designed on conserved upstream and downstream gene regions and used for amplification and sequencing of the two genes from Ra36. A complete list of the primer sequences used in this study is provided in Supplementary Table 1.
    Nucleic acid manipulations
    Isolation of genomic and plasmid DNA from yeast and bacterial cells was performed with the Puragene Yeast/Bact. Kit B and QIAprep Spin Miniprep Kit (Qiagen, Hilden, Germany) according to the manufacturer’s instructions. Total genomic DNA from tomato seedlings was extracted by using the DNeasy Plant Mini Kit (Qiagen) according to the manufacturer’s instructions. Routine nucleic acid manipulations were performed according to standard protocols40.
    Rahnella aquatilis mutant generation
    The Ra36 gcd− and fliC− mutants were generated by targeted replacement with the kanamycin resistance marker. DNA fragments flanking the gcd and fliC coding regions were amplified from gDNA of Ra36 using primer pairs Gcd1+Gcd2, Gcd3+Gcd4 and Flic1+Flic2, Flic3+Flic4, respectively. The kanamycin resistance marker was amplified from the pET28-c (+) plasmid (Novagen, Merck, Darmstadt, Germany) using primers Kanf+Kanr. The S. cerevisiae recombination cloning method41 was used to assemble the complete deletion cassette in the pRS246 vector42 linearized with EcoRI/HindIII, by co-transformation with 1 µg of each of the four DNA fragments into the yeast strain FY83443. The resulting plasmids pRS426.1 and pRS426.2 were isolated from yeast and cloned and propagated in the E. coli TOP10 strain (Invitrogen, Barcelona, Spain) using standard protocols40. The gcd− and fliC− deletion alleles were amplified from pRS426.1 and pRS426.2, respectively, with primers PRS426f+PRS426r and used to transform cells of the Ra36 wild-type strain. Kanamycin-resistant transformants showing homologous insertion in the gcd and fliC coding regions were identified by PCR of gDNA with primers Gcd5+Gcd6 and Flic5+Flic6, respectively. Allelic replacement of the gcd and fliC genes was further confirmed by DNA sequencing of a PCR fragment amplified from gDNA of the transformants with primers Gcd7+Gcd8 and Flic7+Flic8, respectively.
    For fluorescence microscopy analysis, a red fluorescent strain of Ra36 was obtained by transforming bacterial cells with 5 µg of plasmid FPB-31-444 (ATUM Bio, Newark, CA, USA) carrying the red fluorescent protein Rudolph-RFP and an ampicillin resistance cassette, using a Bio-Rad Gene Pulser electroporator (Bio-Rad, Madrid, Spain) set at 1.8 kV, 200 W and 25 μF and 0.1 cm cuvettes.
    Extracellular pH determination and GlcA quantification
    For visual evaluation of Ra36 acidifying activity, 5 µL drops containing 2.5 × 106Ra36 or Ra36 gcd− colony forming unit (CFU) were spot-inoculated onto plates containing minimal medium with urea (MMU; 15 g L−1 glucose, 0.5 g L−1 MgSO4·7(H2O), 0.5 g L−1 KCl, 1.0 g L−1 KH2PO4, 1.5 g L−1 urea, 20 g L−1 agar) adjusted to pH 7. After 2 days of incubation at 28 °C, plates were overlaid with a 0.833 μM solution of the pH indicator bromocresol green or purple and imaged. Experiments were performed at least four times, with two replicates each.
    To assess a possible inhibitory effect of Ra36 on Fol colony growth and to visualize Ra36-mediated alkalinisation of Fol colony margins, 5 µL drops containing 5 × 104Fol microconidia and 2.5 × 106 CFU mL−1Ra36 or Ra36 gcd− were spot-inoculated at a distance of 30 mm on MMU plates adjusted to pH 7.0. Visual analysis of pH change was performed by supplementing media with bromocresol purple (0.833 μM). Plates were incubated at 28 °C for 4–7 d before imaging. All fungus–bacteria confrontation assays were performed in three or more independent experiments with three replicates each.
    For plate visualization of rhizosphere pH dynamics12 during the Fol-Ra36–plant interaction, 2-week-old tomato seedlings (Solanum lycopersicum cv. Moneymaker; used throughout the study) were either left untreated or inoculated by dipping roots into a suspension of Ra36 or Ra36 gcd− cells in water (1 × 109 CFU mL−1). Seedlings were then transferred onto 0.5% water agar plates supplemented with bromocresol purple (0.833 μM) and containing either Fol microconidia (1.25 × 106 mL−1) or water (negative control). Plates were incubated at 25 °C with 14 h light and 10 h dark up to 3 d before imaging. Experiments included three replicate plates and were performed at least three times with similar results.
    To record pH dynamics in liquid media, Fol microconidia (5 × 105 conidia mL−1) were pre-germinated 14 h in MMU medium with shaking (170 rpm) at 28 °C, washed with sterile ddH2O and shifted to fresh MMU medium adjusted to pH 7. Then, 5 × 106 CFU mL−1Ra36 or Ra36 gcd− were added and cultures were incubated at 28 °C and 170 rpm. Each hour, aliquots of culture supernatant were withdrawn, and pH was measured in a pH meter. Cultures containing either Fol or Ra36 alone were used as controls. Experiments were performed on three to four independent occasions with three replicates each.
    To monitor pH dynamics in liquid media in the presence of the plant, 5 × 105 mL−1Fol microconidia and 5 × 106 CFU mL−1Ra36 or Ra36 gcd− were inoculated in sterile H2O. Then, a 2-week-old tomato plant was added to keep only the root submerged in the liquid. The culture was incubated, aliquots of culture supernatant were collected, and pH was measured as described above. Cultures containing either Fol, Ra36, or tomato plants alone were used as controls. Experiments were performed on three to four independent occasions with three replicates each.
    To investigate rhizosphere pH dynamics in soil, 3-week-old tomato seedlings were dip-inoculated with a suspension of Ra36 or Ra36 gcd− cells (1 × 109 CFU mL−1), transferred to non-sterilized or sterilized (autoclaved 20 min at 120 °C) horticultural soil obtained from a tomato field located in Riccia, Molise (Italy) and maintained in a growth chamber (14⁄10 h light⁄dark cycle) at 28 °C. After 2 weeks, 10 plants per treatment were collected, roots were cleaned from the soil, submerged in 5 mL sterile ddH2O for 12 hours at 28 °C and extracellular pH was measured in a pH meter. Uninoculated plants were used as controls. Experiments were performed on four independent occasions with three replicates each.
    D-GlcA concentrations in culture supernatants or tomato root exudates described above was determined using the K-GATE detection kit (Megazyme International, Bray, Ireland) according to the manufacturer’s instructions. For all experiments, GlcA measurements were performed on three to four independent occasions with at least three replicates per treatment.
    Virulence-related assays
    Assays for the evaluation of Fol invasive growth on living plant tissue7 were performed using tomato fruits (cultivar Rovente). In brief, the epidermis of surface sterilized tomato fruits was punctured with a pipette tip and 10 µL of bacterial suspension (109 CFU mL−1) or water (control) was injected into the fruit tissue. After 30 min, 10 µL of a suspension of Fol microconidia (5 × 108 mL−1) or water (control) was injected into the same inoculation site. Fruits were incubated at 28 °C in a humid chamber and infection sites were imaged 5 days after inoculation (DAI). Experiments were performed on three independent occasions with three replicates each.
    Cellophane invasion assays9 were performed on MMU plates. In brief, 100 µL of a Ra36 cell suspension (OD600 of 0.1) or water (control) was uniformly spread on the culture medium before placing an autoclaved cellophane sheet on the plate (colorless; Manipulados Margok, Zizurkil, Spain). Next, 2 × 105Fol microconidia were spot-inoculated at the center of the plate. After 4 days at 28 °C, plates were imaged (before), then the cellophane membrane was carefully removed, and plates were incubated for one additional day at 28 °C and imaged (after). MMU was either unbuffered or buffered to pH 5 or 7 with 100 mm 2-(N-morpholino) ethanesulfonic acid (MES). Experiments included three replicate plates and were performed three times with similar results.
    For analysis of hyphal aggregation, fungal (2.5 × 106 conidia mL−1) and bacterial (108 CFU mL−1) strains were co-cultivated for 48 h in liquid MMU at 28 °C and 170 rpm. Experiments included four replicates and were performed at least three times with similar results.
    For root adhesion assays7, entire roots of 2-week-old tomato seedlings were dip-inoculated with Ra36 as described above, placed in Erlenmeyer flasks containing a suspension of 107 mL−1Fol microconidia and incubated 3 days at 28 °C and 120 rpm. Fungal aggregates in liquid medium or on tomato roots were imaged in a Leica binocular microscope (Leica Microsistemas S.L.U., Barcelona, Spain) using a Leica DC 300 F digital camera. Experiments included four replicates and were performed at least three times with similar results.
    Plant infection assays
    Roots of 2-week-old tomato seedlings were dipped for 30 min into a suspension of Fol microconidia in water (5 × 106 conidia mL−1). For co-inoculation assays, plant roots were immersed for 2 h in a suspension of Ra36 cells (1 × 109 CFU mL−1) and then dip-inoculated with Fol microconidia (5 × 106 conidia mL−1). To test the effect of inoculum density, suspensions of Ra36 cells at different concentrations (1 × 1010; 1 × 109; 1 × 108; 1 × 107, and 1 × 106 CFU mL−1) were used. After root inoculation, tomato seedlings were planted in vermiculite or in non-sterilized or sterilized horticultural soil (see above), maintained in a growth chamber (14⁄10 h light⁄dark cycle) at 28 °C and irrigated either with unbuffered water or with a solution of 1 mM MES buffer adjusted to pH 5.0 or pH 7.0.
    Soil inoculation was performed by mixing non-sterilized or sterilized horticultural soil with Ra36 cells and/or Fol microconidia at concentrations of 1 × 108 CFU and 4 × 105 conidia per g of soil, respectively.
    Plant survival was recorded daily up to 55 days, calculated by the Kaplan–Meier method, and compared among groups using the log-rank test. All infection assays included fifteen plants per treatment and were performed at least three times with similar results. Data were plotted using the GraphPad Prism 5 software (GraphPad Software, La Jolla, CA, USA).
    Detection of Fol and Ra36 in tomato plant tissue
    Qualitative assessment of fungal and bacterial colonization of tomato plants was performed 2 weeks after dip-inoculation of 2-week-old tomato roots with a suspension of Ra36 cells (1 × 109 CFU mL−1) and/or Fol microconidia in water (5 × 106 conidia mL−1) as described above. Seedlings were removed from the vermiculite and the roots and stems were surface sterilized by submerging them into 2% sodium hypochlorite for 2 min. Plant tissues were rinsed twice with sterile distilled water and cut into 1 cm long sections, which were transferred onto MMU plates with or without 0.833 μM of the pH indicator bromocresol purple. Fungal or bacterial growth was recorded after 5 days of incubation at 28 °C. Experiments included three replicate plates per treatment and were performed at least three times with similar results.
    Quantification of fungal and bacterial biomass in Fol- and/or Ra36-inoculated tomato plants grown in vermiculite was carried out by real-time qPCR19, using total gDNA extracted from tomato roots 15 days after dip-inoculation with a suspension of Ra36 cells (1 × 109 CFU mL−1) and/or Fol microconidia (5 × 106 conidia mL−1). To measure the effect of Fol in soil on the colonization ability of Ra36, 2-week-old tomato seedlings were planted in horticultural soil containing 2 × 104Fol microconidia g−1 soil. After 3 days, 50 µl of a suspension of Ra36 cells containing 8 × 108 CFU mL−1 were added at a distance of 5 cm from the hypocotyl. After 10 days, total gDNA was extracted from the tomato roots and biomass of Ra36 was measured by real-time qPCR.
    Real-time qPCR reactions were performed with the SYBR® Premix Ex Taq™ (Takara Bio, Inc., Otsu, Japan) in a Eppendorf Mastercycler ep gradient S system (Eppendorf, Milan, Italy), using the primer pairs Gcd9 + Gcd10 (located outside of the deleted region of the Ra36 gcd gene), ACT2 + ACTQ6 (Fol actin gene) and GADPH1 + GADPH2 (tomato gadph gene) (Supplementary Table 1). Relative amounts of fungal and bacterial genomic DNA were calculated by comparative ΔΔCt with the tomato gadph gene. DNA concentrations in each sample were extrapolated from standard curves obtained by plotting the logarithm of known concentrations (10-fold dilution series from 10 ng to 1 pg/25 μL reaction) of Fol and Ra36 gDNA against the Ct values. To normalize the serially diluted DNA samples, 100 ng gDNA from non-inoculated plants was added to each sample of the dilution series. Real-time qPCR data represent the mean ± SE from three independent experiments, each with five plants per treatment.
    Collection of exudates from tomato roots and Fol hyphae
    To obtain tomato root exudate6, uninoculated or Ra36-inoculated tomato roots (1 × 109 CFU/mL) were placed in sterile ddH2O in the absence or presence of 5 × 105 microconidia mL−1 of Fol. After 48 h at 25 °C, the supernatant was sterilized by filtration through a 0.22-μm membrane (Merck Millipore) and stored at −20 °C until use.
    To obtain Fol hyphal exudate4, 1 × 107 microconidia mL−1 were pre-germinated for 16 h in 50 mL diluted PDB (1:50; v:v in H2O) at 28 °C with shaking at 170 rpm. Germlings were washed twice with sterile ddH2O and incubated for 48 h in 5 mL sterile ddH2O at 28 °C and 170 rpm. The supernatant was sterilized by filtration through a 0.22-μm membrane (Merck Millipore) and stored at −20 °C until use. When required, the pH was adjusted to 5.8 with 0.1 N HCl.
    Assays for fungal chemotropism and bacterial chemotaxis
    Plate preparation, chemoattractant application, and scoring of Fol germ tube redirectioning toward gradients of root exudates obtained from untreated or Ra36-inoculated plants were performed by using a hyphal chemotropism assay6. In brief, 2.5 × 106Fol microconidia mL−1 were embedded in 4 ml water agar (WA; 0.5%, w/v) (Oxoid) and poured into a 9 cm Petri dish. Then, two parallel wells, each at 5 mm distance from the scoring line, were filled with 40 μL of the test compound or the solvent control solution. Scoring was done on five independent batches of cells (n = 100 cells per batch) for each test compound. Experiments were performed at least three times with similar results.
    Chemotaxis capillary assays32 were carried out as follows. Ra36 wild-type, Ra gcd−, and Ra fliC− strains were grown in LB overnight at 28 °C, washed either with sterile ddH2O, PBS, tomato root exudate or Fol hyphal exudate depending on the experiment, and diluted in the same medium to an OD600 of 0.1. Aliquots of 250 μL bacterial suspension were added to individual wells of a 96-well microtiter plate together with a 10-µL capillary containing the test compound or the solvent control. For competing gradient assays, two capillaries containing the different test compounds were added to the well. Plates were incubated for 60 min at 28 °C, capillaries were carefully lifted, the content was serially diluted and plated onto LA medium, and CFUs were counted 48 h after incubation at 28 °C. The following chemoattractant compounds and concentrations were tested: glutamine (Gln), tryptophan (Trp), all at 295 mM; glucose (Gluc), galactose (Gal), all at 50 mM. The chemotaxis ratio was calculated by dividing the number of bacteria in the tube containing the test compound by the number of bacteria in the tube containing the solvent control. All experiments included two replicates and were performed at least three times with similar results.
    Flagellum-dependent swimming motility assays44 were performed as follows. In brief, LB plates (0.3% w/v agarose) were spot-inoculated in the center with 5 μl of an overnight culture of the Ra36 wild-type or fliC− strain, incubated 24 h at 30 °C and colony radial growth was imaged. Experiments were performed three times, with four replicates each.
    Fluorescence microscopy
    For microscopic observation in tomato roots of GFP-tagged Fol and RFP-tagged Ra36 strains, 2-week-old tomato seedlings were dip-inoculated into suspensions of Fol microconidia (5 × 106 conidia mL−1) and/or Ra36 cells (1 × 109 CFU mL−1), planted in moist vermiculite and maintained in a growth chamber (14⁄10 h light⁄dark photoperiod) at 28 °C. Two and four DAI, roots were gently washed to remove the adhering vermiculite, incubated in a 95% perfluorodecalin solution (Sigma-Aldrich), stained for 2 min with 0.005% (w/v) calcofluor white (CFW; Sigma-Aldrich) to visualize the plant cell wall and imaged. Experiments were performed three times, with two replicates each.
    To observe Ra36 swarming toward Fol hyphae, 2 µL drops containing 5 × 104Fol-GFP microconidia or 2.5 × 106 CFU Ra36-RFP were spot-inoculated at a distance of 5 mm on a 1 × 1 cm square pad of soft (0.25% w/v) agarose MMU medium placed on top of a microscope glass slide. Images were recorded every 30 min up to 4 h post inoculation. Experiments were performed three times, with two replicates each.
    To study bacterial movement along fungal hyphae, 2 µL drops containing 5 × 104Fol-GFP microconidia or 2.5 × 106Ra36-RFP CFU were spot-inoculated at a distance of 30 mm on MMU medium plates. Microscopic observation of Fol hyphae was performed 48 h after fungal and bacterial colonies had merged. Experiments were performed four times, with two replicates each.
    To determine the movement of Ra36 and Fol hyphae across a medium-free space, 5 × 104Fol-GFP microconidia and 2.5 × 106Ra36-RFP CFU were spot-inoculated either individually or together on soft (0.7% w/v) agarose MMU plates. Next, a 2-day-old tomato seedling was placed at a distance of 15 mm, and a 5-mm wide medium-free gap was created between the inoculation point and the tomato root by removing the medium with a sterile spatula. Microscopic observation was performed daily up to 3 DAI to follow Ra36 and Fol dynamics over time. To qualitatively assess the presence of Ra36 and Fol on the plant roots, tomato seedlings were gently removed from the plate at 2 DAI, rinsed under sterile H2O and stained with CFW, as described above. Experiments were performed at least three times on two or more separate days.
    Low-resolution imaging was performed using a SteREO Lumar.V12 fluorescence stereomicroscope (Zeiss, Barcelona, Spain). Wide-field fluorescence imaging was performed using a Zeiss Axio Imager M2 Dual Cam microscope (Zeiss) equipped with a Photometrics Evolve EM512 digital camera (Photometrics Technology, Tucson, AZ, USA). Examination using epifluorescence (×400 magnification) was performed with the following filter blocks: CFW staining (G 365, FT 395, LP 420), RFP (BP 560/40, FT 585, BP 630/75), GFP (BP 450/490, FT 510, LP 515). Images were captured and processed using Axiovision 4.8, ZEN lite 2.3 (both from Zeiss) or ImageJ (v1.52)45.
    Statistical analysis
    Percentage of plant survival was compared among treatment groups using the log-rank test. For multiple-group comparisons, a one-way analysis of variance was used for testing no differences among the group means. Post hoc comparisons were adjusted using Dunnett’s or Tukey’s corrections (chemotaxis and real-time qPCR data). Comparisons between two groups were carried out using a two-tailed unpaired Student’s t test (chemotaxis data). A Yates’ corrected Chi-squared test (two-sided) was used to determine significant differences between the observed frequencies of fungal germ tubes pointing toward the chemoattractant or the solvent control (chemotropism data).
    Real-time qPCR data are presented as mean ± SE. Data from chemotaxis and chemotropism experiments and from pH and GlcA measurements are presented as mean ± SD.
    Statistical analyses were performed using GraphPad Prism 5 software (GraphPad Software). In all cases a value of P  More

  • in

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

    23.
    Vrijenhoek, R. DNA primers for amplification of mitochondrial cytochrome c oxidase subunit I from diverse metazoan invertebrates. Mol. Mar. Biol. Biotechnol. 3, 294–299 (1994).
    PubMed  PubMed Central  Google Scholar 

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

  • in

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

  • in

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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