More stories

  • in

    Internal tides can provide thermal refugia that will buffer some coral reefs from future global warming

    1.
    Heron, S. F., Maynard, J. A., van Hooidonk, R. & Eakin, C. M. Warming trends and bleaching stress of the world’s coral reefs 1985–2012. Sci. Rep.6, 38402 (2016).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 
    2.
    Hughes, T. P. et al. Spatial temporal patterns of mass bleaching of corals in the Anthropocene. Science359, 80–83 (2018).
    ADS  CAS  PubMed  Google Scholar 

    3.
    Hoegh-Guldberg, O. Climate change, coral bleaching and the future of the world’s coral reefs. Mar. Freshw. Res.50, 839–866 (1999).
    Google Scholar 

    4.
    Donner, S. D., Skirving, W. J., Little, C. M., Oppenheimer, M. & Hoegh-Guldberg, O. Global assessment of coral bleaching and required rates of adaptation under climate change. Glob. Change Biol.11, 2251–2265 (2005).
    ADS  Google Scholar 

    5.
    Goreau, T. J. & Hayes, R. Coral bleaching and ocean “hot spots.”. Ambio23, 176–180 (1994).
    Google Scholar 

    6.
    Veron, J. E. N. et al. The coral reef crisis: the critical importance of More

  • in

    A shift in sulfur-cycle manipulation from atmospheric emissions to agricultural additions

    1.
    Lamarque, J.-F. et al. Multi-model mean nitrogen and sulfur deposition from the Atmospheric Chemistry and Climate Model Intercomparison Project (ACCMIP): evaluation of historical and projected future changes. Atmos. Chem. Phys. 13, 6247–6294 (2013).
    Google Scholar 
    2.
    Galloway, J. N. Anthropogenic mobilization of sulphur and nitrogen: immediate and delayed consequences. Annu. Rev. Energy Env. 21, 261–292 (1996).
    Google Scholar 

    3.
    Cowling, E. B. Acid precipitation in historical perspective. Environ. Sci. Technol. 16, 110A–123A (1982).
    Google Scholar 

    4.
    Gorham, E. On the acidity and salinity of rain. Geochim. Cosmochim. Acta 7, 231–239 (1955).
    Google Scholar 

    5.
    Likens, G. E. & Bormann, F. H. Acid rain: a serious regional environmental problem. Science 184, 1176–1179 (1974).
    Google Scholar 

    6.
    Goyer, R. A. et al. Potential human health effects of acid rain: report of a workshop. Environ. Health Perspect. 60, 355–368 (1985).
    Google Scholar 

    7.
    Likens, G. E., Driscoll, C. T. & Buso, D. C. Long-term effects of acid rain: response and recovery of a forest ecosystem. Science 272, 244–246 (1996).
    Google Scholar 

    8.
    Johnson, A. H. & Siccama, T. G. Acid deposition and forest decline. Environ. Sci. Technol. 17, 294A–305A (1983).
    Google Scholar 

    9.
    Schulze, E.-D. Air pollution and forest decline in a spruce (Picea abies) forest. Science 244, 776–783 (1989).
    Google Scholar 

    10.
    Driscoll, C. T. et al. Acidic deposition in the northeastern United States: sources and inputs, ecosystem effects and management strategies. BioScience 51, 180–198 (2001).
    Google Scholar 

    11.
    Mitchell, M. J. & Likens, G. E. Watershed sulfur biogeochemistry: shift from atmospheric deposition dominance to climatic regulation. Environ. Sci. Technol. 45, 5267–5271 (2011).
    Google Scholar 

    12.
    EPA Air Emissions Data (EPA, accessed 14 April 2020); https://go.nature.com/3fiYt3p

    13.
    Klimont, Z., Smith, S. J. & Cofala, J. The last decade of global anthropogenic sulfur dioxide: 2000–2011 emissions. Environ. Res. Lett. 8, 014003 (2013).
    Google Scholar 

    14.
    Learn More About Sulphur (The Sulphur Institute, 2020); https://go.nature.com/32OHX87

    15.
    China Statistical Yearbook 2017 (National Bureau of Statistics of China, accessed 1 March 2019); https://go.nature.com/2E7z6E2

    16.
    Thompson, J. F. Sulfur metabolism in plants. Annu. Rev. Plant Physiol. 18, 59–84 (1967).
    Google Scholar 

    17.
    Anderson, J. W. in The Biochemistry of Plants Vol. 16 (ed. Miflin, B. J.) 327–381 (Academic Press, 1990).

    18.
    Canfield, D. E. & Raiswell, R. The evolution of the sulfur cycle. Am. J. Sci. 299, 697–723 (1999).
    Google Scholar 

    19.
    Jackson, G. D. Effects of nitrogen and sulfur on canola yield and nutrient uptake. Agron. J. 92, 644–649 (2000).
    Google Scholar 

    20.
    Ma, B.-L. et al. Growth, yield, and yield components of canola as affected by nitrogen, sulfur, and boron application. J. Plant Nutr. Soil Sci. 178, 658–670 (2015).
    Google Scholar 

    21.
    Clark, N., Orloff, S. & Ottman, M. Fertilizing high yielding alfalfa in California and Arizona. Better Crops with Plant Food 101, 21–23 (2017).
    Google Scholar 

    22.
    Haneklaus, S., Bloem, E., Schnug, E., de Kok, L. J. & Stulen, I. in Handbook of Plant Nutrition (eds Barker, A. V. & Pilbeam, D. J.) Ch. 7 (CRC Press, 2006).

    23.
    Chien, S. H. et al. Agronomic effectiveness of granular nitrogen/phosphorus fertilizers containing elemental sulfur with and without ammonium sulfate: a review. Agron. J. 108, 1203–1213 (2016).
    Google Scholar 

    24.
    Dick, W. A., Kost, D. & Chen, L. in Sulfur: A Missing Link Between Soils, Crops, and Nutrition (ed. Jez, J.) Ch. 5 (ASA, CSSA, SSSA, 2008).

    25.
    Schnug, E. & Evans, E. J. Monitoring of the sulfur supply of agricultural crops in northern Europe. Phyton 32, 119–122 (1992).
    Google Scholar 

    26.
    Gaspar, A. P., Laboski, C. A. M., Naeve, S. L. & Conley, S. P. Secondary and micronutrient uptake, partitioning, and removal across a wide range of soybean seed yield levels. Agron. J. 110, 1328–1338 (2008).
    Google Scholar 

    27.
    Fernández, F. G., Ebelhar, S., Greer, K. & Brown, H. Corn response to sulfur in Illinois FREC 2011 Report (FREC, 2012); https://go.nature.com/32PDORh

    28.
    Steinke, K., Rutan, J. & Thurgood, L. Corn response to nitrogen at multiple sulfur rates. Agron. J. 107, 1347–1354 (2015).
    Google Scholar 

    29.
    Sutradhar, A. K., Kaiser, D. E. & Fernández, F. G. Does total nitrogen/sulfur ratio predict nitrogen or sulfur requirement for corn? Soil Sci. Soc. Am. J. 81, 564–577 (2017).
    Google Scholar 

    30.
    Kurbondski, A. J., Kaiser, D. E., Rosen, C. J. & Sutradhar, A. K. Does irrigated corn require multiple applications of sulfur? Soil Sci. Soc. Am. J. 83, 1124–1136 (2019).
    Google Scholar 

    31.
    Ketterings, Q. M. et al. Soil and tissue testing for sulfur management of alfalfa in New York State. Soil Sci. Soc. Am. J. 76, 298–306 (2012).
    Google Scholar 

    32.
    Haupt, G., Lauzon, J. & Hall, B. Sulfur fertilization: improving alfalfa yield and quality. Crops Soils 48, 26–30 (2015).
    Google Scholar 

    33.
    Data and Statistics (USDA NASS, accessed 20 May 2019); https://go.nature.com/3hxxAcK

    34.
    California Pesticide Information Portal (CalPIP) (California Department of Pesticide Regulation, accessed 20 May 2019); https://calpip.cdpr.ca.gov/main.cfm

    35.
    Orem, W. et al. Sulfur in the South Florida ecosystem: distribution, sources, biogeochemistry, impacts, and management for restoration. Crit. Rev. Environ. Sci. Technol. 41, 249–288 (2011).
    Google Scholar 

    36.
    Gabriel, M., Redfield, G. & Rumbold, D. Sulfur as a regional water quality concern in South Florida 2008 South Florida Environmental Report, Appendix 3B-2 (South Florida Water Management District, 2008).

    37.
    Shainberg, I. et al. in Advances in Soil Science (ed. Stewart, B. A.) 1–111 (Springer, 1989).

    38.
    DeSutter, T. M. & Cihacek, L. J. Potential agricultural uses of flue gas desulfurization gypsum in the Northern Great Plains. Agron. J. 101, 817–825 (2009).
    Google Scholar 

    39.
    Ritchey, K. D., Feldhake, C. M., Clark, R. B. & de Sousa, D. M. G. in Agricultural Utilization of Urban and Industrial By-Products Vol. 58 (eds Karlen, D. L. et al.) Ch. 8 (ASA, CSSA, SSSA, 1995).

    40.
    Driscoll, C. T., Driscoll, K. M., Fakhraei, H. & Civerolo, K. Long-term temporal trends and spatial patterns in the acid-base chemistry of lakes in the Adirondack region of New York in response to decreases in acidic deposition. Atmos. Environ. 146, 5–14 (2016).
    Google Scholar 

    41.
    Rice, K. C., Scanlon, T. M., Lynch, J. A. & Cosby, B. J. Decreased atmospheric sulfur deposition across the southeastern U.S.: when will watersheds release stored sulfate. Environ. Sci. Technol. 48, 10071–10078 (2014).
    Google Scholar 

    42.
    Beaton, J. D. Sulfur requirements of cereals, tree fruits, vegetables, and other crops. Soil Sci. 101, 267–282 (1966).
    Google Scholar 

    43.
    Rehm, G. W. & Clapp, J. G. in Sulfur: A Missing Link between Soils, Crops, and Nutrition (ed. Jez, J.) Ch. 9 (ASA, CSSA, SSSA, 2008).

    44.
    Kaiser, D. E. & Kim, K.-I. Soybean response to sulfur fertilizer applied as a broadcast or starter using replicated strip trials. Agron. J. 105, 1189–1198 (2013).
    Google Scholar 

    45.
    David, M. B., Gentry, L. E. & Mitchell, C. A. Riverine response of sulfate to declining atmospheric sulfur deposition in agricultural watersheds. J. Environ. Qual. 45, 1313–1319 (2016).
    Google Scholar 

    46.
    Wine (Agricultural Marketing Resource Center, 2019); https://go.nature.com/2WO6eHl

    47.
    Hinckley, E. L. S. & Matson, P. A. Transformations, transport, and potential unintended consequences of high sulfur inputs to Napa Valley vineyards. Proc. Natl Acad. Sci. USA 108, 14005–14010 (2011).
    Google Scholar 

    48.
    Williams, J. S. & Cooper, R. M. The oldest fungicide and newest phytoalexin – a reappraisal of the fungitoxicity of elemental sulphur. Plant Pathol. 53, 263–279 (2004).
    Google Scholar 

    49.
    Grape Acreage Reports Listing (USDA National Agricultural Statistics Service, accessed 21 May 2019); https://go.nature.com/38pHlXb

    50.
    US Drought Portal (NIDIS, accessed 21 May 2019); https://go.nature.com/39pyo0w

    51.
    Rice, R. W., Gilbert, R. A. & McCray, J. M. Nutritional requirements for Florida sugarcane Sugarcane Cultural Practices (Sugarcane Handbook), UF-IFAS Extension SS-AGR-228 (Univ. of Florida, 2006).

    52.
    McCray, J. M. Elemental sulfur recommendations for sugarcane on Florida organic soils Sugarcane Cultural Practices (Sugarcane Handbook), UF-IFAS Extension SS-AGR-429 (Univ. of Florida, 2019); http://edis.ifas.ufl.edu/ag429

    53.
    National Research Council Progress Toward Restoring the Everglades: The Fifth Biennial Review: 2014 (The National Academies Press, 2014).

    54.
    Schueneman, T. J. Characterization of sulfur sources in the EAA. Annu. Proc. Soil Crop Sci. Soc. Florida 60, 49–52 (2001).
    Google Scholar 

    55.
    Lanning, M. et al. Intensified vegetation water use under acid deposition. Sci. Adv. 5, eaav5168 (2019).
    Google Scholar 

    56.
    Lu, X. et al. Plant acclimation to long-term high nitrogen deposition in an N-rich tropical forest. Proc. Natl Acad. Sci. USA 115, 5187–5192 (2018).
    Google Scholar 

    57.
    Podar, M. et al. Global prevalence and distribution of genes and microorganisms involved in mercury methylation. Sci. Adv. 1, e1500675 (2015).
    Google Scholar 

    58.
    Driscoll, C. T., Mason, R. P., Chan, H. M., Jacob, D. J. & Pirrone, N. Mercury as a global pollutant: Sources, pathways, and effects. Environ. Sci. Technol. 47, 4967–4983 (2013).
    Google Scholar 

    59.
    Schmeltz, D. et al. MercNet: a national monitoring network to assess responses to changing mercury emissions in the United States. Ecotoxicology 20, 1713–1725 (2011).
    Google Scholar 

    60.
    US Environmental Protection Agency 2011 National Listing of Fisheries Advisories EPA-820-F-13-058 (EPA, 2013).

    61.
    Gilmour, C. C. et al. Methylmercury concentrations and production rates across a trophic gradient in the northern Everglades. Biogeochemistry 40, 327–345 (1998).
    Google Scholar 

    62.
    Bailey, L. T. et al. Influence of porewater sulfide on methylmercury production and partitioning in sulfate-impacted lake sediments. Sci. Total Environ. 580, 1197–1204 (2017).
    Google Scholar 

    63.
    Wasik, J. K. C. et al. The effects of hydrologic fluctuation and sulfate regeneration on mercury cycling in an experimental peatland. J. Geophys. Res. Biogeosciences 120, 1697–1715 (2015).
    Google Scholar 

    64.
    Benoit, J. M. et al. in Biogeochemistry of Environmentally Important Trace Elements (eds Cai, Y. & Braids, O. C.) 262–297 (ACS, 2002).

    65.
    Chen, C. Y., Driscoll, C. T. & Kamman, N. C. in Mercury in the Environment: Pattern and Process (ed. Bank, M.) 143–166 (Univ. of California Press, 2012).

    66.
    Robinson, A., Richey, A., Slotton, D., Collins, J. & Davis, J. North Bay Mercury Biosentinel Project 2016–2017 Contribution # 868 (San Francisco Estuary Institute, Aquatic Science Center, 2018).

    67.
    Marvin-DiPasquale, M., Agee, J. L., Bouse, R. M. & Jaffe, B. E. Microbial cycling of mercury in contaminated pelagic and wetland sediments of San Pablo Bay, California. Environ. Geol. 43, 260–267 (2003).
    Google Scholar 

    68.
    Wiener, J. G., Evers, D. C., Gay, D. A., Morrison, H. A. & Williams, K. A. Mercury contamination in the Laurentian Great Lakes region: introduction and overview. Environ. Pollut. 161, 243–251 (2012).
    Google Scholar 

    69.
    Smolders, A. J. P., Lamers, L. P. M., Lucassen, E. C. H. E. T., Van Dervelde, G. & Roelofs, J. G. M. Internal eutrophication: how it works and what to do about it–a review. Chem. Ecol. 22, 93–111 (2006).
    Google Scholar 

    70.
    Caraco, N. F., Cole, J. J. & Likens, G. E. Evidence for sulphate-controlled phosphorus release from sediments of aquatic systems. Nature 341, 316–318 (1989).
    Google Scholar 

    71.
    Smolders, A. J. P., Lucassen, E. C. H. E. T., Bobbink, R., Roelofs, J. G. M. & Lamers, L. P. M. How nitrate leaching from agricultural lands provokes phosphate eutrophication in groundwater fed wetlands: the sulphur bridge. Biogeochemistry 98, 1–7 (2010).
    Google Scholar 

    72.
    van der Welle, M. E. W., Roelofs, J. G. M. & Lamers, L. P. M. Multi-level effects of sulphur–iron interactions in freshwater wetlands in The Netherlands. Sci. Total Environ. 406, 426–429 (2008).
    Google Scholar 

    73.
    De Kok, L. J., Durenkamp, M., Yang, L. & Stulen, I. in Sulfur in Plants, An Ecological Perspective (eds Hawkesford, M. J. & De Kok, L. J.) Ch. 5 (Springer, 2007).

    74.
    Lamers, L. P. M. et al. Sulfide as a soil phytotoxin—a review. Front. Plant Sci. 4, 268 (2013).
    Google Scholar 

    75.
    Koch, M. S., Mendelssohn, I. A. & McKee, K. L. Mechanism for the hydrogen sulfide‐induced growth limitation in wetland macrophytes. Limnol. Oceanogr. 35, 399–408 (1990).
    Google Scholar 

    76.
    Gao, S., Tanji, K. K. & Scardaci, S. C. Impact of rice straw incorporation on soil redox status and sulfide toxicity. Agron. J. 96, 70–76 (2004).
    Google Scholar 

    77.
    Lamers, L. P. M., Tomassen, H. B. M. & Roelofs, J. G. M. Sulfate-induced eutrophication and phytotoxicity in freshwater wetlands. Environ. Sci. Technol. 32, 199–205 (1998).
    Google Scholar 

    78.
    Li, S., Mendelssohn, I. A., Chen, H. & Orem, W. H. Does sulphate enrichment promote the expansion of Typha domingensis (cattail) in the Florida Everglades? Freshw. Biol. 54, 1909–1923 (2009).
    Google Scholar 

    79.
    Ye, M., Beach, J., Martin, J. & Senthilselvan, A. Occupational pesticide exposures and respiratory health. Int. J. Environ. Res. Public Health 10, 6442–6471 (2013).
    Google Scholar 

    80.
    Hoppin, J. A., Umbach, D. M., London, S. J., Alavanja, M. C. R. & Sandler, D. P. Chemical predictors of wheeze among farmer pesticide applicators in the Agricultural Health Study. Am. J. Respir. Crit. Care Med. 165, 683–689 (2002).
    Google Scholar 

    81.
    Degryse, F., Ajiboye, B., Baird, R., da Silva, R. C. & McLaughlin, M. J. Oxidation of elemental sulfur in granular fertilizers depends on the soil-exposed surface area. Soil Sci. Soc. Am. J. 80, 294–305 (2016).
    Google Scholar 

    82.
    Guo, J. H. et al. Significant acidification in major Chinese croplands. Science 327, 1008–1010 (2010).
    Google Scholar 

    83.
    Clark, M. & Tilman, D. Comparative analysis of environmental impacts of agricultural production systems, agricultural input efficiency, and food choice. Environ. Res. Lett. 12, 064016 (2017).
    Google Scholar 

    84.
    Galloway, J. N. et al. The nitrogen cascade. BioScience 53, 341–356 (2003).
    Google Scholar  More

  • in

    Dual clumped isotope thermometry resolves kinetic biases in carbonate formation temperatures

    Samples
    Devils Hole vein calcite: A Holocene vein calcite (DHC2-8) that precipitated 4.5–16.9 ka before present, was collected from the Devils Hole cave #2 in Nevada, USA (36.427138 N, 116.291172 W). It is postulated that DHC2-8 precipitated at extremely slow rates, i.e., 0.1–0.8 μm year−1, at a constant temperature of 33.7(±0.8) °C30. For this study, we prepared a ca 0.5 × 0.5 × 1.5 cm slab of calcite. First, we abraded the surface ca 0.5 mm of the slab with a slow-speed hand-held drill to remove impurities. Then, the slab was cleaned in de-ionised water in an ultrasonic bath for 5 min and dried in a vacuum oven at 30 °C before it was crushed and homogenised using an agate mortar and pestle. Material from the same vein was analysed for clumped isotopes in other studies5,32.
    Cryogenic cave carbonate: A coarsely crystalline cryogenic cave calcite (MSK 2b) was obtained from Mitterschneidkar Eishöhle in the Austrian Alps (47.1165 N, 11.7407 E). The cave opens at 2258 m above sea level and contained perennial ice in the near-entrance part until 2007. Today the cave is ice-free, and the mean annual air temperature in the interior of the cave is 0.23 °C33. Coarsely crystalline cryogenic cave carbonates generally precipitate slowly and very close to 0 °C, otherwise powder-like fine-crystalline cryogenic cave carbonates form50. Cryogenic cave carbonates occur in several places in the inner part of the cave, and U–Th dating of these carbonates suggests the presence of perennial ice up to about 2600 years before present33. The sample crystals were crushed and homogenised using an agate mortar and pestle and were subsequently dried in a vacuum oven at 30 °C before isotope analysis. Additional information on the potential equilibrium nature of this sample is found in Supplementary Fig. 1.
    Cave pool carbonate: A 3.5-cm thick subaqueous calcite sample (Obi 87-i) was collected in 2008 from a perennial pool (Silbersee) in Rasslsystem cave, which is part of the Obir Caves (46.5102 N, 14.5480 E), a series of karst caves in the Southern Alps of Austria, located at approximately 1100 m above sea level. The Obir Caves are hypogene in origin51, i.e., they were not connected to the surface and hence had a very stable microclimate until discovered during mining activities in the 1870s. The Silbersee pool, located in the inner part of Rasslsystem cave, has a surface area of 7 × 4 m and is on average ca 1 m deep. The pool water temperature between 1998 and 2002 was 5.4(±0.1) °C, closely corresponding to the long-term mean annual air temperature outside the cave at this elevation52. The sample analysed in this study is a 4 mm wide subsample retrieved from 2.7 cm above the base of Obi 87, and is estimated to have formed at about 1500 years before present, based on the U–Th dating of a lower layer in Obi 87 (a layer estimated to have formed about 3800 years before present at 1.5 cm above the base of Obi 87) and assuming a constant calcite growth rate of 5.3 μm year−1 (unpublished data, C Spötl). Although the water temperature about 1500 years before present is not precisely known, it likely corresponded to the mean annual air temperature outside the cave at that time in a similar way as the modern pool temperature does. Various temperature proxy data for the Alps suggest that the mean annual air temperature fluctuated by up to ±1.5 °C in the last two millennia before the industrial revolution53. Considering the ca 1.5 °C warming in the Alps during the past century, we estimate the water temperature of Silbersee pool from which Obi 87-i precipitated ca 1500 years ago to be 4.0(±1.5) °C. Experiments demonstrated that subaqueous pool carbonates can precipitate in oxygen isotope equilibrium with water54. Prior to isotope analyses, Obi 87-i was cleaned in de-ionised water in an ultrasonic bath, crushed and homogenised using an agate mortar and pestle, and dried in a vacuum oven at 30 °C.
    Synthetic speleothem carbonate: A calcite sample (MHD1) was precipitated in a laboratory, under cave-like conditions55. Solutions super-saturated relative to calcite were pumped to flow down an inclined, sandblasted glass plate in a thin solution film (0.1 mm in thickness), precipitating CaCO3 along the flow path. The experiments were performed in an enclosed space, which allowed control of all surrounding conditions, such as pCO2, temperature, and relative humidity. Specifically, sample MHD1 was precipitated at 30.7(±0.3) °C, with an atmospheric pCO2 of 1007(±47) ppm and a relative humidity of 97.5(±1.2)%. The average δ13C and δ18O values of the atmospheric CO2 were −44.7(±0.8)‰ and −10.6(±0.6)‰ VPDB, respectively. The experimental solution was prepared by dissolving 5 mmol CaCO3 powder in high-purity water while bubbling tank CO2 through the water column. After the complete dissolution of CaCO3 powder, i.e., when there were no visible particles in the solution, the solution was stored for five days at the experimental temperature to obtain isotopic equilibrium among all dissolved inorganic carbon species. This resulted in an initial solution composition of pH = 6.34, [DIC] = 18.19 mM, δ13CHCO3− ≈ −31.9(±1.3)‰, and δ18OHCO3− ≈ −8.69(±0.11)‰ VPDB. After being exposed to lower pCO2 in the climate box, it took ca 18 s for the solution to reach chemical equilibrium with the atmospheric CO2, which increased the solution pH and led to super-saturation with respect to calcite. The calcite sample was scratched off the glass plate after the experiment was completed and corresponded to the first 5 cm of flow, i.e., the first 24 s of CaCO3 precipitation.
    Stalagmite: A calcite sample (SPA121-02) was retrieved from a stalagmite in Spannagel Cave in the Austrian Alps (47.0803 N, 11.6717 E), an extensive cave system with the main entrance at 2523 m above sea level. SPA121-02 is a 4-mm-thin layer within SPA121, a stalagmite that records a long growth history from about 240 to 76 ka. SPA121-02 was formed at about 225 ka during Marine Isotope Stage (MIS) 7.4 when this high-alpine cave was buried beneath a warm-based glacier preventing the cave from freezing56. The growth of this stalagmite during MIS 7 likely occurred at constant temperatures around freezing point, i.e., 0(±2) °C. The relatively high δ13C values of SPA121-02 (about 7‰ VPDB, Supplementary Table 1) was attributed to the disequilibrium isotope effects during peak cold times56. A 3 × 6 × 4 mm large piece was cut out from the axial part of the stalagmite SPA121 using a diamond-coated band saw. The piece was then cleaned in an ultrasonic bath in de-ionised water, dried, and crushed and homogenised with an agate mortar and pestle before isotope analysis.
    Cold-water coral: A scleractinian, azooxanthellate coral Desmophyllum pertusum (formerly known as Lophelia pertusa) (JR) was collected alive at Traenadjupet, Norwegian Sea (66.973333 N, 11.108833 E) at a water depth of 300 m during cruise POS325-356/1. The annual mean seawater temperature at the collection location is 7.2(±1.0) °C57. With a hand-held drill, a corallite was cut from the axis of the colony, and the septa were removed, i.e., only the theca walls were sampled. The sample was cleaned in an ultrasonic bath using de-ionised water and dried in a vacuum oven at 30 °C before being crushed and homogenised using an agate mortar and pestle.
    Warm-water coral: A scleractinian, zooxanthellate coral Porites lutea (PC1_2005) was collected at the Rashdoo Atoll, Maldives (4.293776 N, 72.977115 E) at a water depth of ca 1 m. For isotope analysis, a ca 0.5 cm thick section was cut from the coral core. Based on sclerochronological analysis, this section corresponded to the growth in the year 2005 when the annual mean temperature at this location was 29.3(±1.0) °C58. The mean annual extension rate of the coral is ca 10 mm year−1. To remove material that may have been thermally altered when the section was initially cut from the colony, the surface 0.5 mm was scraped away. Then, the section was cleaned in an ultrasonic bath using de-ionised water and dried in a vacuum oven at 30 °C before being crushed and homogenised using an agate mortar and pestle.
    Modern brachiopod shell: A terebratulid brachiopod Magellania venosa (Mv143-b) was collected from Punta Gruesa, Chile (42.409833 S, 72.424333 W) from 20 m below sea level. The annual mean temperature at the collection location is 11.4(±1.7) °C12. Magellania venosa is one of the fastest-growing modern brachiopods, with growth rates ranging from 3.8 mm year−1 (adult) to 17.3 mm year−1 (juvenile)59. For this study, we sampled a ca 2 × 3 cm area in the middle part of the ventral valve. First, we abraded the primary layer of the shell using a slow-speed hand-held drill and a diamond drill bit, cleaned the shell in an ultrasonic bath with de-ionised water, dried it in a vacuum oven at 30 °C, and finally homogenised the material using an agate mortar and pestle. The anterior part of the same specimen showed the largest offset from equilibrium in ∆47 values in a previous study12.
    Cretaceous belemnite: A belemnite Belemnopsis sp. (66-4.65) was retrieved from DSDP Site 511 at the Falkland Plateau (51.004667 S, 46.971667 W). The investigated rostrum solidum shows excellent preservation based on cathodoluminescence, and trace element analyses44,60. Burial temperatures remained below 100 °C for this core, which makes the solid-state alteration of the clumped isotope composition of this sample unlikely61,62. The same sample in this study was analysed for its ∆47 values, together with other belemnites, to reconstruct Early Cretaceous southern high latitude palaeotemperatures44.
    Mass spectrometry
    We performed the CO2 clumped isotope analyses of sample carbonates on a Thermo Scientific 253 Plus gas source isotope ratio mass spectrometer during April 2019–March 2020, in three measurement sessions (April–August 2019, September–December 2019, and January–March 2020), following the method of Fiebig et al.25. Samples were measured in 5–10 replicates. Each replicate analysis consists of 13 acquisitions (10 cycles of reference and samples comparisons in each acquisition and 20 s integration time during each cycle). The raw clumped isotope values (indicated by subscript “raw” on the ∆ symbol) were calculated using the [Brand]/IUPAC isotopic parameters29,63.
    Data correction for the reference carbonates
    In order to assign the long-term ∆47 (CDES90) and ∆48 (CDES90) values of the ETH 1, ETH 2, and ETH 3 carbonate reference materials finally used for sample correction (Table 1, see the next section), we followed the correction approach outlined by Fiebig et al.25. using equilibrated gases only (subscript “CDES90” on the ∆ symbol indicates that the ∆47 and ∆48 values of these carbonate reference materials are reported on the Carbon Dioxide Equilibrium Scale at a reaction temperature of 90 °C). A total of 36 aliquots of CO2 gases equilibrated at 25 °C and 54 aliquots equilibrated at 1000 °C were considered for the April–August 2019 period (Supplementary Data 1). Data correction for the reference carbonates consisted of two steps: correction for non-linearity followed by correction for scale compression25,64,65. These two steps are detailed below.
    Correction for non-linearity: Within errors, the two sets of equilibrium gases, equilibrated either at 1000 °C or 25 °C, had identical slopes in ∆47 (raw) vs δ47 (Supplementary Fig. 3a) and ∆48 (raw) vs δ48 (Supplementary Fig. 3b) spaces, respectively, when the negative m/z 47.5 intensity is directly subtracted from measured m/z 47–49 intensities (scaling factor of −1, see below and in Fiebig et al.25). We, therefore, considered the slopes displayed by the merged data sets for the correction of non-linearity. In ∆47 (raw) vs δ47 space, the equilibrium gases showed a flat line such that non-linearity correction needs not be applied. In ∆48 (raw) vs δ48 space, the slope displayed by the merged data set was −0.0040(±0.0002).
    Correction for scale compression: The intercepts for the 1000 °C and the 25 °C gases displayed in ∆47 (raw) vs δ47 and ∆48 (raw) vs δ48 spaces were compared to the corresponding theoretical values66 to constrain empirical transfer functions (Supplementary Data 1).
    Finally, we combined the ∆47 (CDES90) and ∆48 (CDES90) values of ETH 1, ETH 2, and ETH 3 determined during the April–August 2019 period (Supplementary Data 1) with those reported in Fiebig et al.25 to calculate the long-term values listed in Table 1 (Supplementary Fig. 4). Shapiro-Wilks tests show that the combined ∆47 (CDES90) and ∆48 (CDES90) data set of the carbonate reference materials have a normal distribution with W-values  > 0.95 and p-values  > > 0.05.
    Data correction for the carbonate samples
    Unlike the method described in Fiebig et al.25, we did not make use of equilibrated gases to correct the samples but used our long-term ∆47 (CDES90) and ∆48 (CDES90) values obtained for ETH 1, ETH 2, and ETH 3 instead (Supplementary Data 2–4). This purely carbonate-based correction approach follows the principle of identical treatment of sample and reference materials and allows identification of subtle temporal drifts in the acid reaction environment and correction for them67,68,69. Correction of the sample data consisted of three steps: correction for non-linearity followed by correction for scale compression, and finally correction for variations in the reaction environment. These three steps are detailed below.
    Correction for non-linearity: The negative background causing the non-linearities in ∆47 (raw) vs δ47, ∆48 (raw) vs δ48, and ∆49 (raw) vs δ49 spaces was corrected using Easotope70 by subtracting the intensities measured on the m/z 47.5 cup from the intensities measured on the m/z 47–49 cups, after multiplying the former by respective scaling factors. These scaling factors were determined empirically and enable one to calculate accurate negative backgrounds below m/z 47, m/z 48, and m/z 49 from the measured m/z 47.5 intensity. For the three periods of sample analyses, i.e., April–August 2019, September–December 2019, and January–March 2020, we determined the scaling factors in a way that no residual slopes remained between the respective measured values of the frequently analysed ETH 1 and ETH 2 standards in the corresponding ∆ vs δ spaces (Supplementary Figs. 5–7). The uniformity of the measured long-term ∆47 (CDES90) and ∆48 (CDES90) values of ETH 1 and ETH 2, also supported by experimental data71, allowed us to assume that they have identical ∆47 and ∆48 values (Table 1). Consequently, for the April–August 2019 period of sample analyses, scaling factors of −0.988, −0.906, and −0.648, respectively, were applied to correct m/z 47, m/z 48, and m/z 49 intensities based on the monitored m/z 47.5 intensity. For the September–December 2019 period, the corresponding scaling factors were −1.003, −0.938, and −0.581, respectively. For the January–March 2020 period, factors of −1.010, −0.92326, and −0.555, respectively, were applied.
    Correction for scale compression: According to the principles outlined above, we used our long-term ∆47 (CDES90) and ∆48 (CDES90) values of ETH 1, ETH 2, and ETH 3 (Table 1) to project the non-linearity corrected, raw clumped isotope values of the carbonate samples to the CDES. We determined empirical transfer functions based on a comparison of our long-term mean ∆47 (CDES90) and ∆48 (CDES90) values of ETH 1, ETH 2, and ETH 3 (Table 1) with their corresponding, non-linearity corrected ∆47 (raw) and ∆48 (raw) averages over the three periods of sample analysis (Supplementary Figs. 5a–b, 6a–b, 7a–b). The application of these transfer functions to non-linearity corrected sample ∆47 (raw) and ∆48 (raw) values yields the ∆47 (CDES90,uc) and ∆48 (CDES90,uc) values of the samples (Supplementary Data 2–4).
    Correction for subtle long-term variations in the acid reaction environment: When residuals between the accepted long-term and the measured ∆ (CDES90,uc) for all ETH 1, ETH 2, and ETH 3 replicate analyses are plotted against time, small but systematic temporal variations become detectable. For ∆47 (CDES90,uc), these residuals are on the order of ≤0.010‰ (Supplementary Figs. 8a, 9a, 10a), and for ∆48 (CDES90,uc) they are on the order of ≤0.030‰ (Supplementary Figs. 8b,  9b, 10b). We determined a residual vs measurement time function (Supplementary Data 2–4) and used it to further correct the ∆47 (CDES90,uc) and ∆48 (CDES90,uc) values in order to obtain the final clumped isotope compositions of the investigated carbonate samples (Table 2).
    We used the non-linearity corrected ∆49 (raw) values of the carbonate-derived CO2 and the presumably uncontaminated equilibrated CO2 gases to check for potential contamination in the analyte. All ∆49 (raw) values of the carbonates fall within the range of the ∆49 (raw) values of the equilibrated gases, indicating no contamination of the investigated solids (Supplementary Figs. 11a–b, 12a–b, 13a–b). In addition, the lack of correlation between ∆48 (raw) and ∆49 (raw) of the measured analytes further argues that there is no contamination on ∆49 that would influence ∆48 (Supplementary Figs. 11c, 12c, 13c). All measured values can be found in Supplementary Data 1–4.
    Acid fractionation factors
    To be able to compare the experimentally measured clumped isotope compositions of a carbonate, i.e., the ∆47 (CDES90) and ∆48 (CDES90) values of the CO2 gas derived from the phosphoric acid digestion of that carbonate, with its theoretically predicted composition, i.e., the ∆63 and ∆64 values of the carbonate, we determined25 the clumped isotope fractionation factors associated with the 90 °C acid fractionation during our analysis. These are based on the long-term ∆47 (CDES90) and ∆48 (CDES90) values of ETH 1 and ETH 2 standards which were both potentially equilibrated at 600 °C68. The theoretically predicted calcite ∆63 and ∆64 values at 600 °C are 0.018‰ and 0.002‰28, respectively. These, combined with our experimentally measured ∆47 (CDES90) values of 0.212(±0.002)‰ and ∆48 (CDES90) of 0.140(±0.005)‰, yield acid fractionation factors of 0.194(±0.002)‰ for ∆63–∆47 and 0.138(±0.005)‰ for ∆64–∆48.
    Numerical modelling
    We used numerical models to simulate the evolution of the isotopic composition of the DIC during (1) CO2 absorption, i.e., the key process involved in coral calcification26, and (2) the laboratory carbonate precipitation of the synthetic speleothem27 (Supplementary Data 5).
    (1) CO2 absorption simulations were constructed using the IsoDIC model to mimic the internal calcification environment of scleractinian corals26. Specifically, the modelled calcification environment consisted of an aqueous solution ([DIC] = 2 mM, δ13CDIC = 0, and pH = 8.8 for cold-water corals and pH = 8.5 for warm-water corals), which was exposed to a CO2-containing atmosphere (pCO2 = 1100 ppm72 and δ13CCO2 = −15‰73). The temperature of the modelled calcification environment corresponded to the mean growth temperatures of the cold- and warm-water corals, i.e., 7.2 °C and 28.9 °C, respectively. The catalytic enhancement of the inter-conversion between CO2 (aq) and HCO3− by carbonic anhydrase during coral calcification is simulated by increasing the rate constants of CO2 (aq) (de)hydration reactions26. The initial oxygen and clumped isotope compositions of both the DIC and air CO2 were assumed to be in isotopic equilibrium with the water (δ18OH2O = 0 VSMOW) at the above described temperatures.
    (2) To model the isotopic composition of the synthetic speleothem, simulations were constructed using the IsoCave model27, based on the conditions of the laboratory experiment55 (T = 30.7 °C, pCO2 = 1007 ppm, water film thickness of 100 μm, δ13CCO2 = −44.7‰, δ18OCO2 = −10.6‰ VPDB, δ13CCaCO3 = −6‰, δ18OH2O = −9‰ VSMOW, see above as well) and yielded an initial solution composition of pH = 6.3, [DIC] = 18.1 mM, [Ca2+] = 4.9 mM, δ13CHCO3− = −31.2‰, and δ18OHCO3− = −9.0‰ VPDB, which are close to the experimentally determined values (pH = 6.34, [DIC] = 18.2 mM, [Ca2+] = 5 mM, δ13CHCO3− ≈ −31.9(±1.3)‰, and δ18OHCO3− ≈ −8.69(±0.11)‰ VPDB, see above). More

  • in

    De novo transcriptome assembly, functional annotation, and expression profiling of rye (Secale cereale L.) hybrids inoculated with ergot (Claviceps purpurea)

    1.
    Geiger, H. & Miedaner, T. Rye breeding. Cereals3, 157–181 (2009).
    Google Scholar 
    2.
    Bushuk, W. Rye: Production, Chemistry, and Technology (American Association of Cereal Chemists, Saint Paul, 2001).
    Google Scholar 

    3.
    Bauer, E. et al. Towards a whole-genome sequence for rye (Secale cereale L.). Plant J.89, 853–869 (2017).
    CAS  PubMed  Google Scholar 

    4.
    Tudzynski, P., Correia, T. & Keller, U. Biotechnology and genetics of ergot alkaloids. Appl. Microbiol. Biotechnol.57, 593–605 (2001).
    CAS  PubMed  Google Scholar 

    5.
    Streit, E. et al. Current situation of mycotoxin contamination and co-occurrence in animal feed—focus on Europe. Toxins4, 788–809 (2012).
    CAS  PubMed  PubMed Central  Google Scholar 

    6.
    Tudzynski, P. & Scheffer, J. Claviceps purpurea: molecular aspects of a unique pathogenic lifestyle. Mol. Plant Pathol.5, 377–388 (2004).
    CAS  PubMed  Google Scholar 

    7.
    Wang, Z., Gerstein, M. & Snyder, M. RNA-Seq: a revolutionary tool for transcriptomics. Nat. Rev. Genet.10, 57–63 (2009).
    CAS  PubMed  PubMed Central  Google Scholar 

    8.
    Brown, J. W., Calixto, C. P. & Zhang, R. High-quality reference transcript datasets hold the key to transcript-specific RNA-sequencing analysis in plants. New Phytol.213, 525–530 (2017).
    CAS  PubMed  Google Scholar 

    9.
    Watson-Lazowski, A. et al. Plant adaptation or acclimation to rising CO2? Insight from first multigenerational RNA-Seq transcriptome. Glob. Change Biol.22, 3760–3773 (2016).
    ADS  Google Scholar 

    10.
    O’Rourke, J. A. et al. An RNA-Seq transcriptome analysis of orthophosphate-deficient white lupin reveals novel insights into phosphorus acclimation in plants. Plant Physiol.161, 705–724 (2013).
    PubMed  Google Scholar 

    11.
    Chakraborty, S., Britton, M., Martínez-García, P. & Dandekar, A. M. Deep RNA-Seq profile reveals biodiversity, plant–microbe interactions and a large family of NBS-LRR resistance genes in walnut (Juglans regia) tissues. Amb. Express6, 12 (2016).
    PubMed  PubMed Central  Google Scholar 

    12.
    Costa, V., Angelini, C., De Feis, I. & Ciccodicola, A. Uncovering the complexity of transcriptomes with RNA-Seq. BioMed Res. Int. https://doi.org/10.1155/2010/853916 (2010).
    Article  Google Scholar 

    13.
    McIntyre, L. M. et al. RNA-seq: technical variability and sampling. BMC Genom.12, 293 (2011).
    CAS  Google Scholar 

    14.
    Bolger, A. M., Lohse, M. & Usadel, B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics30, 2114–2120 (2014).
    CAS  PubMed  PubMed Central  Google Scholar 

    15.
    Wang, S. & Gribskov, M. Comprehensive evaluation of de novo transcriptome assembly programs and their effects on differential gene expression analysis. Bioinformatics33, 327–333 (2017).
    CAS  PubMed  Google Scholar 

    16.
    Grabherr, M. G. et al. Full-length transcriptome assembly from RNA-Seq data without a reference genome. Nat. Biotechnol.29, 644–652 (2011).
    CAS  PubMed  PubMed Central  Google Scholar 

    17.
    Haas, B. J. et al. De novo transcript sequence reconstruction from RNA-Seq: reference generation and analysis with Trinity. Nat. Protoc.8, 1494 (2013).
    CAS  PubMed  Google Scholar 

    18.
    Smith-Unna, R., Boursnell, C., Patro, R., Hibberd, J. M. & Kelly, S. TransRate: reference-free quality assessment of de novo transcriptome assemblies. Genome Res.26, 1134–1144 (2016).
    CAS  PubMed  PubMed Central  Google Scholar 

    19.
    Fu, L., Niu, B., Zhu, Z., Wu, S. & Li, W. CD-HIT: accelerated for clustering the next-generation sequencing data. Bioinformatics28, 3150–3152 (2012).
    CAS  PubMed  PubMed Central  Google Scholar 

    20.
    Simão, F. A., Waterhouse, R. M., Ioannidis, P., Kriventseva, E. V. & Zdobnov, E. M. BUSCO: assessing genome assembly and annotation completeness with single-copy orthologs. Bioinformatics31, 3210–3212 (2015).
    PubMed  Google Scholar 

    21.
    Mirdita, V. & Miedaner, T. Resistance to Ergot in Self-incompatible Germplasm Resources of Winter Rye. J. Phytopathol.157, 350–355 (2009).
    Google Scholar 

    22.
    Mirdita, V., Dhillon, B., Geiger, H. & Miedaner, T. Genetic variation for resistance to ergot (Claviceps purpurea [Fr.] Tul.) among full-sib families of five populations of winter rye (Secale cereale L.). Theor. Appl. Genet.118, 85–90 (2008).
    CAS  PubMed  Google Scholar 

    23.
    Anna, K. et al. Ergot infection in winter rye hybrids shows differential contribution of male and female genotypes and environment. Euphytica216, 65 (2020).
    Google Scholar 

    24.
    Lindner, R. & Friedel, C. C. A comprehensive evaluation of alignment algorithms in the context of RNA-seq. PLoS ONE7, e52403 (2012).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    25.
    Nookaew, I. et al. A comprehensive comparison of RNA-Seq-based transcriptome analysis from reads to differential gene expression and cross-comparison with microarrays: a case study in Saccharomyces cerevisiae. Nucleic Acids Res.40, 10084–10097 (2012).
    CAS  PubMed  PubMed Central  Google Scholar 

    26.
    Benjamin, A. M., Nichols, M., Burke, T. W., Ginsburg, G. S. & Lucas, J. E. Comparing reference-based RNA-Seq mapping methods for non-human primate data. BMC Genom.15, 570 (2014).
    Google Scholar 

    27.
    Engelke, T. Ansätze für eine integrierte Bekämpfung des Mutterkorns (Claviceps purpurea [Fr.] Tul.) im Roggen (Cuvillier, Göttingen, 2002).
    Google Scholar 

    28.
    Miedaner, T., Mirdita, V., Rodemann, B., Drobeck, T. & Rentel, D. Genetic variation of winter rye cultivars for their ergot (Claviceps purpurea) reaction tested in a field design with minimized interplot interference. Plant Breed.129, 58–62 (2010).
    Google Scholar 

    29.
    Smart, M. & Roden, L. C. A small-scale RNA isolation protocol useful for high-throughput extractions from recalcitrant plants. S. Afr. J. Bot.76, 375–379 (2010).
    CAS  Google Scholar 

    30.
    30Brown, C. T., Howe, A., Zhang, Q., Pyrkosz, A. B. & Brom, T. H. A reference-free algorithm for computational normalization of shotgun sequencing data. arXiv preprint https://arxiv.org/abs/1203.4802 (2012).

    31.
    Pevzner, P. A., Tang, H. & Waterman, M. S. An Eulerian path approach to DNA fragment assembly. Proc. Natl. Acad. Sci.98, 9748–9753 (2001).
    ADS  MathSciNet  CAS  PubMed  MATH  Google Scholar 

    32.
    Zhao, Q.-Y. et al. Optimizing de novo transcriptome assembly from short-read RNA-Seq data: a comparative study. BMC Bioinform.12, S2 (2011).
    CAS  Google Scholar 

    33.
    Conesa, A. et al. Blast2GO: a universal tool for annotation, visualization and analysis in functional genomics research. Bioinformatics21, 3674–3676 (2005).
    CAS  Google Scholar 

    34.
    Kanehisa, M. & Goto, S. KEGG: kyoto encyclopedia of genes and genomes. Nucleic Acids Res.28, 27–30 (2000).
    CAS  PubMed  PubMed Central  Google Scholar 

    35.
    Robinson, M. D., McCarthy, D. J. & Smyth, G. K. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics26, 139–140 (2010).
    CAS  PubMed  PubMed Central  Google Scholar 

    36.
    Biselli, C. et al. Comparative transcriptome profiles of near-isogenic hexaploid wheat lines differing for effective alleles at the 2DL FHB resistance QTL. Frontiers Plant Sci.9, 37 (2018).
    Google Scholar 

    37.
    Ma, X., Keller, B., McDonald, B. A., Palma-Guerrero, J. & Wicker, T. Comparative transcriptomics reveals how wheat responds to infection by Zymoseptoria tritici. Mol. Plant Microbe Interact.31, 420–431 (2018).
    CAS  PubMed  Google Scholar 

    38.
    Li, Y. et al. Transcriptome analysis identifies candidate genes and functional pathways controlling the response of two contrasting barley varieties to powdery mildew infection. Int. J. Mol. Sci.21, 151 (2020).
    CAS  Google Scholar 

    39.
    Haseneyer, G. et al. From RNA-seq to large-scale genotyping-genomics resources for rye (Secale cereale L.). BMC Plant Biol.11, 131 (2011).
    CAS  PubMed  PubMed Central  Google Scholar 

    40.
    Hill, C. B. et al. De novo transcriptome assembly and analysis of differentially expressed genes of two barley genotypes reveal root-zone-specific responses to salt exposure. Sci. Rep.6, 31558 (2016).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    41.
    Liu, Y.-J. et al. Transcriptome analysis of wheat seedling and spike tissues in the hybrid Jingmai 8 uncovered genes involved in heterosis. Planta247, 1307–1321 (2018).
    CAS  PubMed  Google Scholar 

    42.
    Duan, J., Xia, C., Zhao, G., Jia, J. & Kong, X. Optimizing de novo common wheat transcriptome assembly using short-read RNA-Seq data. BMC Genom.13, 392 (2012).
    CAS  Google Scholar 

    43.
    Vicente, R. et al. De novo transcriptome analysis of durum wheat flag leaves provides new insights into the regulatory response to elevated CO2 and high temperature. Frontiers Plant Sci.10, 1605 (2019).
    Google Scholar 

    44.
    Kang, S.-H., Lee, J.-Y., Lee, T.-H., Park, S.-Y. & Kim, C.-K. novo transcriptome assembly of the Chinese pearl barley, adlay, by full-length isoform and short-read RNA sequencing. PLoS ONE13, e0208344 (2018).
    CAS  PubMed  PubMed Central  Google Scholar 

    45.
    Fox, S. E. et al. De novo transcriptome assembly and analyses of gene expression during photomorphogenesis in diploid wheat Triticum monococcum. PLoS ONE9, e96855 (2014).
    ADS  PubMed  PubMed Central  Google Scholar 

    46.
    Gullner, G., Komives, T., Király, L. & Schröder, P. Glutathione S-transferase enzymes in plant-pathogen interactions. Frontiers in plant science9, 1836 (2018).
    PubMed  PubMed Central  Google Scholar 

    47.
    Bienert, M. D., Baijot, A. & Boutry, M. Plant ABC Transporters 137–162 (Springer, Berlin, 2014).
    Google Scholar 

    48.
    Xin, M. et al. Cloning and characterization of a novel UDP-glycosyltransferase gene induced by DON from wheat. J. Integr. Agric.14, 830–838 (2015).
    Google Scholar 

    49.
    Oh, B.-J. et al. A cytochrome P450 gene is differentially expressed in compatible and incompatible interactions between pepper (Capsicum annuum) and the anthracnose fungus Colletotrichum gloeosporioides. Mol. Plant-Microbe Interact.12, 1044–1052 (1999).
    CAS  PubMed  Google Scholar 

    50.
    Silveira, R. L., Stoyanov, S. R., Gusarov, S., Skaf, M. S. & Kovalenko, A. Plant biomass recalcitrance: effect of hemicellulose composition on nanoscale forces that control cell wall strength. J. Am. Chem. Soc.135, 19048–19051 (2013).
    CAS  PubMed  Google Scholar 

    51.
    Volpi, C. et al. Claviceps purpurea expressing polygalacturonases escaping PGIP inhibition fully infects PvPGIP2 wheat transgenic plants but its infection is delayed in wheat transgenic plants with increased level of pectin methyl esterification. Plant Physiol. Biochem.73, 294–301 (2013).
    CAS  PubMed  Google Scholar 

    52.
    An, S. H. et al. Pepper pectin methylesterase inhibitor protein CaPMEI1 is required for antifungal activity, basal disease resistance and abiotic stress tolerance. Planta228, 61–78 (2008).
    CAS  PubMed  PubMed Central  Google Scholar 

    53.
    Volpi, C. et al. The ectopic expression of a pectin methyl esterase inhibitor increases pectin methyl esterification and limits fungal diseases in wheat. Mol. Plant Microbe Interact.24, 1012–1019 (2011).
    CAS  PubMed  Google Scholar 

    54.
    Lionetti, V. et al. Three pectin methylesterase inhibitors protect cell wall integrity for Arabidopsis immunity to Botrytis. Plant Physiol.173, 1844–1863 (2017).
    CAS  PubMed  PubMed Central  Google Scholar 

    55.
    Lionetti, V. et al. Overexpression of pectin methylesterase inhibitors in Arabidopsis restricts fungal infection by Botrytis cinerea. Plant Physiol.143, 1871–1880 (2007).
    CAS  PubMed  PubMed Central  Google Scholar 

    56.
    Tundo, S. et al. Pyramiding PvPGIP2 and TAXI-III but not PvPGIP2 and PMEI enhances resistance against Fusarium graminearum. Mol. Plant Microbe Interact.29, 629–639 (2016).
    CAS  PubMed  Google Scholar 

    57.
    Prabhu, S. A. et al. Immuno-affinity purification of PglPGIP1, a polygalacturonase-inhibitor protein from pearl millet: studies on its inhibition of fungal polygalacturonases and role in resistance against the downy mildew pathogen. Mol. Biol. Rep.42, 1123–1138 (2015).
    CAS  PubMed  Google Scholar 

    58.
    Oeser, B., Heidrich, P. M., Müller, U., Tudzynski, P. & Tenberge, K. B. Polygalacturonase is a pathogenicity factor in the Claviceps purpurea/rye interaction. Fungal Genet. Biol.36, 176–186 (2002).
    CAS  PubMed  Google Scholar 

    59.
    Lionetti, V., Cervone, F. & Bellincampi, D. Methyl esterification of pectin plays a role during plant–pathogen interactions and affects plant resistance to diseases. J. Plant Physiol.169, 1623–1630 (2012).
    CAS  PubMed  Google Scholar 

    60.
    Malinovsky, F. G., Fangel, J. U. & Willats, W. G. The role of the cell wall in plant immunity. Frontiers Plant Sci.5, 178 (2014).
    Google Scholar 

    61.
    Bacete, L., Mélida, H., Miedes, E. & Molina, A. Plant cell wall-mediated immunity: cell wall changes trigger disease resistance responses. Plant J.93, 614–636 (2018).
    CAS  PubMed  Google Scholar 

    62.
    Hückelhoven, R. Cell wall–associated mechanisms of disease resistance and susceptibility. Annu. Rev. Phytopathol.45, 101–127 (2007).
    PubMed  Google Scholar 

    63.
    Li, S. et al. Arabidopsis COBRA-LIKE 10, a GPI-anchored protein, mediates directional growth of pollen tubes. Plant J.74, 486–497 (2013).
    CAS  PubMed  Google Scholar 

    64.
    Tudzynski, P., Tenberge, K. & Oeser, B. Claviceps purpurea. Pathog. Host Specif. Plant Dis. Histopathol. Biochem. Genet. Mol. Bases2, 161–187 (1995).
    Google Scholar 

    65.
    Miedaner, T. & Geiger, H. H. Biology, genetics, and management of ergot (Claviceps spp.) in rye, sorghum, and pearl millet. Toxins7, 659–678 (2015). More

  • in

    Repeatable ecological dynamics govern the response of experimental communities to antibiotic pulse perturbation

    1.
    Smith, M. D., Knapp, A. K. & Collins, S. L. A framework for assessing ecosystem dynamics in response to chronic resource alterations induced by global change. Ecology 90, 3279–3289 (2009).
    PubMed  Google Scholar 
    2.
    Rockstrom, J. et al. A safe operating space for humanity. Nature 461, 472–475 (2009).
    PubMed  Google Scholar 

    3.
    Morley, V. J., Woods, R. J. & Read, A. F. Bystander selection for antimicrobial resistance: implications for patient health. Trends Microbiol. 27, 864–877 (2019).
    CAS  PubMed  Google Scholar 

    4.
    Vrancken, G., Gregory, A. C., Huys, G. R. B., Faust, K. & Raes, J. Synthetic ecology of the human gut microbiota. Nat. Rev. Microbiol. 17, 754–763 (2019).
    CAS  PubMed  Google Scholar 

    5.
    Lässig, M., Mustonen, V. & Walczak, A. M. Predicting evolution. Nat. Ecol. Evol. 1, 0077 (2017).
    Google Scholar 

    6.
    Robinson, J. I. et al. Metabolomic networks connect host-microbiome processes to human Clostridioides difficile infections. J. Clin. Invest. 130, 3792–3806 (2019).
    Google Scholar 

    7.
    Oh, S., Choi, D. & Cha, C. J. Ecological processes underpinning microbial community structure during exposure to subinhibitory level of triclosan. Sci. Rep. 9, 4598 (2019).
    PubMed  PubMed Central  Google Scholar 

    8.
    Zorner, P., Farmer, S. & Alibek, K. Quantifying crop rhizosphere microbiome ecology: the next frontier in enhancing the commercial utility of agricultural microbes. Ind. Biotechnol. 14, 116–119 (2018).
    Google Scholar 

    9.
    Lozada-Gobilard, S. et al. Environmental filtering predicts plant-community trait distribution and diversity: kettle holes as models of meta-community systems. Ecol. Evol. 9, 1898–1910 (2019).
    PubMed  PubMed Central  Google Scholar 

    10.
    Li, J. et al. Shared molecular targets confer resistance over short and long evolutionary timescales. Mol. Biol. Evol. 36, 691–708 (2019).
    CAS  PubMed  PubMed Central  Google Scholar 

    11.
    Vázquez-García, I. et al. Clonal heterogeneity influences the fate of new adaptive mutations. Cell Rep. 21, 732–744 (2017).
    PubMed  PubMed Central  Google Scholar 

    12.
    Illingworth, C. J. R., Parts, L., Schiffels, S., Liti, G. & Mustonen, V. Quantifying selection acting on a complex trait using allele frequency time series data. Mol. Biol. Evol. 29, 1187–1197 (2012).
    CAS  PubMed  Google Scholar 

    13.
    Leibler, S. & Kussell, E. Individual histories and selection in heterogeneous populations. Proc. Natl Acad. Sci. USA 107, 13183–13188 (2010).
    CAS  PubMed  Google Scholar 

    14.
    Des Roches, S. et al. The ecological importance of intraspecific variation. Nat. Ecol. Evol. 2, 57–64 (2018).
    PubMed  Google Scholar 

    15.
    Wistrand-Yuen, E. et al. Evolution of high-level resistance during low-level antibiotic exposure. Nat. Commun. 9, 1599 (2018).
    PubMed  PubMed Central  Google Scholar 

    16.
    Schoener, T. W. The newest synthesis: understanding the interplay of evolutionary and ecological dynamics. Science 331, 426–429 (2011).
    CAS  PubMed  Google Scholar 

    17.
    Scheuerl, T. et al. Bacterial adaptation is constrained in complex communities. Nat. Commun. 11, 754 (2020).
    CAS  PubMed  PubMed Central  Google Scholar 

    18.
    Roodgar, M. et al. Longitudinal linked read sequencing reveals ecological and evolutionary responses of a human gut microbiome during antibiotic treatment. Preprint at bioRxiv https://doi.org/10.1101/2019.12.21.886093 (2019).

    19.
    Dethlefsen, L. & Relman, D. A. Incomplete recovery and individualized responses of the human distal gut microbiota to repeated antibiotic perturbation. Proc. Natl Acad. Sci. USA 108, 4554–4561 (2011).
    CAS  PubMed  Google Scholar 

    20.
    Palleja, A. et al. Recovery of gut microbiota of healthy adults following antibiotic exposure. Nat. Microbiol. 3, 1255–1265 (2018).
    CAS  PubMed  Google Scholar 

    21.
    Jernberg, C., Lofmark, S., Edlund, C. & Jansson, J. K. Long-term ecological impacts of antibiotic administration on the human intestinal microbiota. ISME J. 1, 56–66 (2007).
    CAS  PubMed  PubMed Central  Google Scholar 

    22.
    Yassour, M. et al. Natural history of the infant gut microbiome and impact of antibiotic treatment on bacterial strain diversity and stability. Sci. Transl. Med. 8, 343ra81 (2016).
    PubMed  PubMed Central  Google Scholar 

    23.
    Mack, I. et al. Antimicrobial resistance following azithromycin mass drug administration: potential surveillance strategies to assess public health impact. Clin. Infect. Dis. 70, 1501–1508 (2019).
    Google Scholar 

    24.
    Tamburini, S., Shen, N., Wu, H. C. & Clemente, J. C. The microbiome in early life: implications for health outcomes. Nat. Med. 22, 713–722 (2016).
    CAS  PubMed  Google Scholar 

    25.
    Langdon, A., Crook, N. & Dantas, G. The effects of antibiotics on the microbiome throughout development and alternative approaches for therapeutic modulation. Genome Med. 8, 39 (2016).
    PubMed  PubMed Central  Google Scholar 

    26.
    Brown, J. H. & Kodricbrown, A. Turnover rates in insular biogeography: effect of immigration on extinction. Ecology 58, 445–449 (1977).
    Google Scholar 

    27.
    Patel, R. & DuPont, H. L. New approaches for bacteriotherapy: prebiotics, new-generation probiotics, and synbiotics. Clin. Infect. Dis. 60, 108–121 (2015).
    Google Scholar 

    28.
    Lawley, T. D. et al. Targeted restoration of the intestinal microbiota with a simple, defined bacteriotherapy resolves relapsing Clostridium difficile disease in mice. PLoS Pathog. 8, e1002995 (2012).
    CAS  PubMed  PubMed Central  Google Scholar 

    29.
    Ingrisch, J. & Bahn, M. Towards a comparable quantification of resilience. Trends Ecol. Evol. 33, 251–259 (2018).
    PubMed  Google Scholar 

    30.
    Morton, J. T. et al. Establishing microbial composition measurement standards with reference frames. Nat. Commun. 10, 2719 (2019).
    PubMed  PubMed Central  Google Scholar 

    31.
    Trindade, S., Sousa, A. & Gordo, I. Antibiotic resistance and stress in the light of Fisher’s model. Evolution 66, 3815–3824 (2012).
    PubMed  Google Scholar 

    32.
    Jasinska, W. et al. Chromosomal barcoding of E. coli populations reveals lineage diversity dynamics at high resolution. Nat. Ecol. Evol. 4, 437–452 (2020).
    PubMed  Google Scholar 

    33.
    Springer, B. et al. Mechanisms of streptomycin resistance: selection of mutations in the 16S rRNA gene conferring resistance. Antimicrob. Agents Chemother. 45, 2877–2884 (2001).
    CAS  PubMed  PubMed Central  Google Scholar 

    34.
    Nishimura, K., Hosaka, T., Tokuyama, S., Okamoto, S. & Ochi, K. Mutations in rsmG, encoding a 16S rRNA methyltransferase, result in low-level streptomycin resistance and antibiotic overproduction in Streptomyces coelicolor A3(2). J. Bacteriol. 189, 3876–3883 (2007).
    CAS  PubMed  PubMed Central  Google Scholar 

    35.
    Macfarlane, E. L. A., Kwasnicka, A. & Hancock, R. E. W. Role of Pseudomonas aeruginosa PhoP-PhoQ in resistance to antimicrobial cationic peptides and aminoglycosides. Microbiology 146, 2543–2554 (2000).
    CAS  PubMed  Google Scholar 

    36.
    Girgis, H. S., Hottes, A. K. & Tavazoie, S. Genetic architecture of intrinsic antibiotic susceptibility. PLoS ONE 4, e5629 (2009).
    PubMed  PubMed Central  Google Scholar 

    37.
    Liu, Y., Chen, X., Pan, L. & Mao, Z. Differential protein expression of a streptomycin-resistant Streptomyces albulus mutant in high yield production of ε-poly-l-lysine: a proteomics study. RSC Adv. 9, 24092–24104 (2019).
    CAS  Google Scholar 

    38.
    Beardmore, R. E. et al. Drug-mediated metabolic tipping between antibiotic resistant states in a mixed-species community. Nat. Ecol. Evol. 2, 1312–1320 (2018).
    PubMed  Google Scholar 

    39.
    Perron, G. G., Gonzalez, A. & Buckling, A. The rate of environmental change drives adaptation to an antibiotic sink. J. Evol. Biol. 21, 1724–1731 (2008).
    CAS  PubMed  Google Scholar 

    40.
    Retel, C. et al. The feedback between selection and demography shapes genomic diversity during coevolution. Sci. Adv. 5, eaax0530 (2019).
    PubMed  PubMed Central  Google Scholar 

    41.
    Day, T., Huijben, V. & Read, A. F. Is selection relevant in the evolutionary emergence of drug resistance? Trends Microbiol. 23, 126–133 (2015).
    CAS  PubMed  PubMed Central  Google Scholar 

    42.
    Melnyk, A. H., Wong, A. & Kassen, R. The fitness costs of antibiotic resistance mutations. Evol. Appl. 8, 273–283 (2015).
    PubMed  Google Scholar 

    43.
    Wein, T. et al. Carrying capacity and colonization dynamics of Curvibacter in the Hydra host habitat. Front. Microbiol. 9, 443 (2018).
    PubMed  PubMed Central  Google Scholar 

    44.
    Zhou, J. et al. Stochastic assembly leads to alternative communities with distinct functions in a bioreactor microbial community. mBio 4, e00584-12 (2013).
    PubMed  PubMed Central  Google Scholar 

    45.
    Lozupone, C. A., Stombaugh, J. I., Gordon, J. I., Jansson, J. K. & Knight, R. Diversity, stability and resilience of the human gut microbiota. Nature 489, 220–230 (2012).
    CAS  PubMed  PubMed Central  Google Scholar 

    46.
    Shade, A. et al. Fundamentals of microbial community resistance and resilience. Front. Microbiol. 3, 417 (2012).
    PubMed  PubMed Central  Google Scholar 

    47.
    Waples, R. S., Beechie, T. & Pess, G. R. Evolutionary history, habitat disturbance regimes, and anthropogenic changes: what do these mean for resilience of Pacific salmon populations? Ecol. Soc. 14, 3 (2009).
    Google Scholar 

    48.
    Johnstone, J. F. et al. Changing disturbance regimes, ecological memory, and forest resilience. Front. Ecol. Environ. 14, 369–378 (2016).
    Google Scholar 

    49.
    Cairns, J. et al. Construction and characterization of synthetic bacterial community for experimental ecology and evolution. Front. Genet. 9, 312 (2018).
    PubMed  PubMed Central  Google Scholar 

    50.
    Datta, N., Hedges, R. W., Shaw, E. J., Sykes, R. B. & Richmond, M. H. Properties of an R factor from Pseudomonas aeruginosa. J. Bacteriol. 108, 1244–1249 (1971).
    CAS  PubMed  PubMed Central  Google Scholar 

    51.
    Cairns, J. et al. Ecology determines how low antibiotic concentration impacts community composition and horizontal transfer of resistance genes. Commun. Biol. 1, 35 (2018).
    PubMed  PubMed Central  Google Scholar 

    52.
    Zhang, J., Kobert, K., Flouri, T. & Stamatakis, A. PEAR: a fast and accurate Illumina paired-end read merger. Bioinformatics 30, 614–620 (2014).
    CAS  PubMed  Google Scholar 

    53.
    Martin, M. & Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet J. 17, 10–12 (2011).
    Google Scholar 

    54.
    Ewels, P., Magnusson, M., Lundin, S. & Kaller, M. MultiQC: summarize analysis results for multiple tools and samples in a single report. Bioinformatics 32, 3047–3048 (2016).
    CAS  PubMed  PubMed Central  Google Scholar 

    55.
    Edgar, R. C. UPARSE: highly accurate OTU sequences from microbial amplicon reads. Nat. Methods 10, 996–998 (2013).
    CAS  PubMed  Google Scholar 

    56.
    Rognes, T., Flouri, T., Nichols, B., Quince, C. & Mahe, F. VSEARCH: a versatile open source tool for metagenomics. PeerJ 4, e2584 (2016).
    PubMed  PubMed Central  Google Scholar 

    57.
    Edgar, R. C. Accuracy of microbial community diversity estimated by closed- and open-reference OTUs. PeerJ 5, e3889 (2017).
    PubMed  PubMed Central  Google Scholar 

    58.
    Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359 (2012).
    CAS  PubMed  PubMed Central  Google Scholar 

    59.
    Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).
    PubMed  PubMed Central  Google Scholar 

    60.
    Quinlan, A. R. & Hall, I. M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010).
    CAS  PubMed  PubMed Central  Google Scholar 

    61.
    Garrison, E. & Gabor, M. Haplotype-based variant detection from short-read sequencing. Preprint at arXiv https://arxiv.org/abs/1207.3907 (2012).

    62.
    Cingolani, P. et al. A program for annotating and predicting the effects of single nucleotide polymorphisms, SnpEff: SNPs in the genome of Drosophila melanogaster strain w(1118); iso-2; iso-3. Fly 6, 80–92 (2012).
    CAS  PubMed  PubMed Central  Google Scholar 

    63.
    R Core Team R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, 2019); http://www.R-project.org/

    64.
    Krijthe, J. Rtsne: T-distributed stochastic neighbor embedding using a Barnes-Hut implementation. R package version 0.15 https://github.com/jkrijthe/Rtsne (2015).

    65.
    Liaw, A. & Wiener, M. Classification and regression by randomForest. R News 2, 18–22 (2002).
    Google Scholar 

    66.
    Kuhn, M. et al. caret: Classification and regression training. R package version 6.0-84 https://CRAN.R-project.org/package=caret (2019).

    67.
    Pinheiro, J., Bates, D., DebRoy, S., Sarkar, D. & R Core Team nlme: Linear and nonlinear mixed effects models. R package version 3.1-142 https://CRAN.R-project.org/package=nlme (2019).

    68.
    Venables, W. N. & Ripley, B. D. Modern Applied Statistics With S 4th edn (Springer, 2002)

    69.
    Oksanen, J. et al. vegan: Community ecology package. R package version 2.5-6 https://CRAN.R-project.org/package=vegan (2019).

    70.
    Zapala, M. A. & Schork, N. J. Multivariate regression analysis of distance matrices for testing associations between gene expression patterns and related variables. Proc. Natl Acad. Sci. USA 103, 19430–19435 (2006).
    CAS  PubMed  Google Scholar 

    71.
    Excoffier, L., Smouse, P. E. & Quattro, J. M. Analysis of molecular variance inferred from metric distances among DNA haplotypes: application to human mitochondrial DNA restriction data. Genetics 131, 479–491 (1992).
    CAS  PubMed  PubMed Central  Google Scholar 

    72.
    Anderson, M. J. Distance-based tests for homogeneity of multivariate dispersions. Biometrics 62, 245–253 (2006).
    PubMed  Google Scholar 

    73.
    Anderson, M. J. A new method for non-parametric multivariate analysis of variance. Austral Ecol. 26, 32–46 (2001).
    Google Scholar  More

  • in

    Dynamics of entomopathogenic nematode foraging and infectivity in microgravity

    Rearing EPNs
    Steinernema feltiae (SN strain) IJs were obtained from the International Entomopathogenic Nematode Collection (USDA-ARS, Byron Georgia USA). The nematodes were cultured in the laboratory using the White trap method41. Galleria mellonella (Vanderhorst Wholesale Inc. St. Marys, OH) were exposed to 100 IJs per larva. Infected G. mellonella larvae were incubated for 4 days at RT (20 ± 1 °C) and insect cadavers were transferred to White traps for IJ collection.
    Why Steinernema feltiae was chosen
    S. feltiae is a widely studied nematode and commercially used as a biocontrol agent in diverse systems. We selected S. feltiae for two reasons. (1) The dispersal behavior at the expected ISS temperature was the main factor in our decision. S. feltiae disperses without a quiescent period at temperatures from 15 °C to 30 °C. Some other EPNs such as S. carpocapsae IJs disperse normally at or above 25 °C, but at 20 °C and below, the nematodes have a quiescent period where IJs stay stand still for period of 40 min to 24h14. ISS ambient temperature is between 21 and 23 °C. We did not want to take chance with potential failure due to a temperature fluctuation so we chose S. feltiae, which is known to be active at the temperatures42 of the ISS. (2) S. feltiae is an intermediate forager thus incorporating behaviors of both foraging extremes43.
    NASA safety certification, flight configuration, and safety experiments
    The NASA safety certification was conducted by an implementation partner, NanoRacks, LLC (Houston, TX). The safety certification included filling out safety and experimental plan documents, submitting a bill of materials, and providing material safety sheets for the materials that would be flown to the ISS, their MSDS, and conducting safety and flight configuration experiments. To determine the risk of freezing tubes with moist sand on the ISS cold stowage, we sent three 15 ml conical tubes filled with 10% moist sand to NanoRacks for NASA safety testing. Briefly, play sand (Quikrete, Atlanta, GA) was washed three times in tap water and rinsed three times with MILLIQ water (MILLIPORE, Burlington, MA) using a gold pan (Garrett’s gravity trap, Garland, TX) then dried at 60 °C in the oven for 3 days or until all the moisture was evaporated. In a separate dish, dry sand was moistened with 10% MILLIQ water (w/v). Twelve ml moist sand was placed in 15 ml conical tubes that were rinsed three times with MILLIQ water and dried prior to use. We sent three tubes with moist sand to NanoRacks to be tested at NASA for safety. Next, we moved forward with flight configuration experiments for Specimens 2–6, prepared as shown in Fig. 5. Specimens 2–6 each with two replications were conducted once at 23 ± 2 °C and once at 15 ± 1 °C to determine whether nematodes can emerge and go through in the sand, infect the bait insects and reproduce within a month in a horizontal position. At the same time, we examined whether the tubes for Specimens 2–6 developed any discoloration or warp due to the byproducts produced during infection and decomposition of the cadavers for a month. Later, the flight configuration experiment was extended to six months in case return to Earth would be delayed due to lack of space in the resupply capsule. The lower temperature (15 °C) was tested as a precaution to determine whether nematodes still infect if the temperature fluctuated unexpectedly at ISS. We did one last safety test for Specimen 6 (insect only) to determine whether gasses released by decomposing insects could discolor or warp the tubes, causing a hazard on the ISS. For this experiment, two G. mellonella prepared in Fig. 5c were first placed in −80 °C overnight to kill insect larvae humanely and then the tubes incubated at 23 ± 2 °C. Two replications with a total four insect larvae were monitored monthly for over 6 months for discoloration and warping due to decomposing insect larvae. No warping or discoloration was observed, fulfilling safety requirements to move forward with the experiments.
    Fig. 5: Experimental design of specimens for microgravity and their Earth control.

    b Specimens 1–4. One end of the tube, designated as 1, had either two infected insect hosts for providing infective juvenile nematodes (Steinernema feltiae) or 2000 IJs in 0.5 ml of polyacrylamide gel. The other end of the tube, designated as 2, had either two healthy bait insects or infected insect. bS. feltiae IJs in gel. 4000 IJs/ml of polyacrylamide gel. c Two healthy Galleria mellonella larvae with wood shavings as shown.

    Full size image

    Preparing IJs for the microgravity experiments
    S. feltiae IJs were removed from the White trap 4 days after the beginning of emergence and rinsed three times in deionized water to remove residual cadaver-derived metabolites and pheromones (Supplementary Table 1) as described by Kaplan et al.14. Rinsed IJs were resuspended to infect the Galleria mellonella larvae for Specimens 1, 2, and 3 and used to prepare IJs in polyacrylamide gel (75 ml of water 1 g of a polyacrylamide gel (Soil Moist, JRM Chemical, Cleveland OH) with a density of 4000 IJs/ml for Specimens 4 and 5 as shown in Fig. 5a for both microgravity experiments and their corresponding Earth controls.
    Six concurrent experiments in microgravity for emergence, infection and reproduction and space flight
    We designed multiple concurrent experiments that were conducted at the International Space Station U.S. National Laboratory (ISS) with controls conducted on Earth. The experiments were designated as Specimens 1–6 (Fig. 5). The concurrent experiments were conducted to capture the maximum amount of information about the EPN infection process and life cycle in microgravity and make sure that usable data was returned even if microgravity interrupted part of the EPN life cycle. The space limitations of conducting an experiment on the ISS also had to be considered. Microgravity causes changes in many variables (not just gravity), which may impact a multistep infection requiring the cooperation of two organisms (nematode and its mutualistic bacteria). Furthermore, some of the facilities on the ISS, such as cold stowage (−80 °C), have limited capacity and high demand. A timeline of the experiments (Specimens) is presented in Fig. 2 and Supplementary Table 1.
    We used a modified sand column assay44,45 shown in Fig. 5. The experimental units consisted of 15 ml conical tubes (Fig. 5) filled with S. feltiae infected G. mellonella larvae or 2000 S. feltiae IJs in 0.5 ml of polyacrylamide gel were placed in one end of the tube and the healthy larvae were placed on the other end of the tube (CELLTREAT scientific products, San Diego, CA). Based on preliminary evidence, the set-up allows a 3-day time delay for the nematodes to move through the sand and infect the bait larvae. However, we did not know what delay there would be under microgravity when setting up the experiment.
    Specimen 1 was prepared as shown in Fig. 5a to determine whether nematodes reproduce, form the IJ stage, emerge, disperse, travel 10 cm in moist sandy soil, find a bait insect and infect/invade the host in microgravity (Fig. 1a–d). If the infection failed, this specimen was also designed to determine whether it was due to insect host-immune response. Specimen 1 was placed in cold stowage (−80 °C) 7 days after the launch at ISS, which was after the anticipated time of IJ invasion. A corresponding Earth control was prepared for both IJ invasion and insect host immunity response. Since we did not know how freezing or a month in −80 °C storage would affect hemocyte counts, we prepared additional Earth controls to be analyzed on the day the Earth Specimen 1 controls were placed in −80 °C (Supplementary Table 1). Three replicates were prepared for each treatment; microgravity, corresponding Earth control and additional Earth control for hemocyte counts. A total of 18 insects in nine individual tubes were analyzed for IJ invasion, to determine the IJ developmental stage and hemocyte counts for both Space and Earth specimens. The hemocyte counts were done for the additional Earth control (no freezing) on the day the other set was placed in the −80 °C freezer. Mean hemocyte count for the additional Earth control (no freezing) was 2.8 ± 1.7 from three replicates.
    Specimen 2 was prepared in the same manner as Specimen 1 except that the tubes were not frozen to allow IJs to develop and emerge from the bait insect. Unfortunately, they died upon return to Earth. Three replicates were prepared for each treatment; microgravity and Earth control with a total of six tubes analyzed.
    Specimen 3 was prepared the same as Specimen 1 except that it did not include a bait insect and the location of the infected insect was on the opposite side (Fig. 5). This specimen was designed to determine whether nematodes can continue their development and reproduction during the flight and emerge in microgravity. Since we did not know whether IJs could emerged from consumed cadavers under microgravity, we prepared Specimen 3. Sand was included to make sure that the emerged nematodes have a place to live until the experiment was returned to Earth and to make a good comparison to Specimen 1. Furthermore, Specimen 3 was kept at ambient temperature to test IJs dispersal after returning to Earth. Again, three replicate tubes for each treatment (microgravity and Earth control) were prepared with a total of six replicate tubes.
    Specimen 4 was prepared the same as specimen 1 except that it had 2000 IJs in 0.5 ml of polyacrylamide gel instead of an infected insect (Fig. 5a). This was a backup experiment just in case Specimens 1 and 2 failed to emerge in microgravity to infect bait insects. The column contained sand in the middle (causing a time delay) so the infection event would occur in space not on Earth. The experiment had to be delivered to NASA 36 h before the rocket launch. The sand provided enough time delay (2–3 days) that IJs would only find the bait insects when they reached the ISS or when the Dragon Capsule was in space. This experiment had an Earth control like the other experiments and one additional Earth control was included to determine whether the bait insects were infected before the rocket launch. Three replicates were prepared for each treatment; microgravity and two sets of Earth controls with a total of nine tubes analyzed.
    The rocket launch was delayed 1 day from December 4 to 5. The IJs had 2 days, 21 h and 20 min to go through 10 cm sand to reach the bait insect as opposed to 1 day 21 h and 20 min as planned launch day December 4. This was very concerning. At the earliest time (1 and ½ h) after the rocket launch, we analyzed the additional Earth control (three replicates) for insect invasion in two steps. First, we placed the IJs in a petri dish and added water to rinse the IJs off the surface of the insects. Each replicate was analyzed on a separate petri dish. The bait insects were alive and in all three replicates had IJs on them. Next, we washed all the insects to remove all the IJs from their surface and looked inside the insects to determine whether any of the IJs were inside the insects. No IJs were found inside the bait insects suggesting that none of the insects were infected while they were on Earth.
    Specimen 5 contained S. feltiae IJs in polyacrylamide gel with a density of 4000 IJ/ml (Fig. 5b). A total volume of 10 ml of IJ acrylamide suspension was placed in each of three 15 ml conical tubes. This specimen was prepared to determine how 33 days of space flight would affect IJ adjustment (movement) and infectivity after returning to Earth. Upon return to Earth at the earliest time (3 days after returning), IJ movement was compared to an Earth control and 10 days later, IJ infectivity was tested. This was also a backup experiment in case nematodes died due to the effect of sand because IJs were known to safely travel on Earth in polyacrylamide gel43. Like the other experiments, three replicate tubes (120,000 IJs) for each treatment (microgravity and Earth control) were prepared with a total of six replicate tubes (240,000 IJs).
    Specimen 6 had two G. mellonella last instar larvae in 15 ml conical tubes in wood shaving (Fig. 5c). This was to determine if the infection failed in Specimen 1 due negative effects on the insect host. At the end of the space flight, insects in both microgravity and Earth control were dead at different developmental stages. Several insect stages were observed in the various tubes (larvae, pupae and adult), all of which died. A total of 12 insect larvae in six replicate tubes were analyzed.
    All Specimens, except for Specimen 1, were kept at ambient temperature on the ISS, during the flight back to Earth, and for 3 days while being shipped to our laboratory for further analysis on nematode IJ fitness and infectivity.
    Handing over the experiments to NanoRacks for delivery to NASA for launch
    Specimens 1–4 and 6 in Fig. 5 were assembled at Kennedy Space Center (KSC) Space Station Processing Facility (SSPF) on December 2 (Supplementary Table 1). Specimen 5 in Fig. 5b was prepared on Nov 29 at the Shapiro lab in Byron, GA, and kept at 5 °C until December 2. After all of the specimens were prepared (Supplementary Fig. 1), Specimen 1 replications were placed in a sealed bag and Specimens 2–6 were placed in a separate sealed bag. They are weighed and placed in Nanotracks’ NanoLab in a Horizontal form (Supplementary Fig. 1). The Nanolab lid was sealed with Velcro. After that, the Nanolab was delivered to NASA and placed in the Dragon Capsule on SpaceX Falcon9 rocket for CRS-19 on the same day (Supplementary Table 1). The experiments in the Nanolab were placed in a horizontal position and kept 20 °C for 2 days, 21 h and 20 min in the Dragon capsule until the launch on December 5, 2019. Experiments were kept in a horizontal position during the launch, in microgravity in Space, and at ISS at ambient temperature until returned to Earth.
    Earth controls: Specimen 1 for Earth control was kept in a sealed plastic bag and Specimens 2–6 were placed in a separate sealed plastic bag at ambient temperature at the KSC SSPF until launch on December 5. Specimen 1 Earth controls were taken to the Shapiro Lab by car and kept at RT until December 11. Specimens 2–6 were shipped by FedEx at ambient temperature using cool packs for maintenance of the temperature to the Pheronym R&D laboratory, Davis, CA
    In orbit and return to Earth
    Specimen 1 (the three 15 ml conical tubes) was transferred to cold storage on December 12 at 346/14:30 (8:30am CST) by European Space Agency Astronaut Luca Parmitano (https://blogs.nasa.gov/stationreport/2019/12/12/). Specimens 2–6 were kept in the NanoLab on the Dragon Capsule which returned to Earth on January 7, 2020 and was retrieved from the Pacific Ocean by Space X. NanoRacks received the samples on Jan 9, 2020, in their facility in Houston, TX and shipped Specimen 1 (kept frozen) in dry ice overnight by FedEx to the Shapiro Lab in Byron, GA. The Shapiro Lab in Byron GA placed the samples in −80 °C until analysis. On the same day (Jan 9), NanoRacks shipped Specimens 2–6 at RT overnight by FedEx to Pheronym, Davis, CA. On January 10, 2020 at 10 AM, Specimens 2–6 were received by Pheronym Lab, and were inspected immediately for dead, alive sluggish or active nematodes, pictures were taken, and videos recorded before removing the Teflon seals. Pictures were taken using a Nikon D60 (Nikon, Tokyo, Japan). Videos were taken with a hand-held USB Celestron microscope (Celestron, Torrance, CA) and/or iPhone 6 S (Apple, Cupertino, CA).
    Immune response assessment of Specimen 1
    The primary immune defense in insects against multicellular parasites including EPNs is encapsulation. Immune cells (hemocytes) bind to the nematode and one another to form a multicellular envelope; we therefore assessed immune response based on hemocyte activity and encapsulation46. G. mellonella larvae were dissected longitudinally. The presence of hemocytes adhering to the nematode body was confirmed at 400–600X. The number of nematodes showing hemocyte activity was compared between treatments.
    Assessing the IJs after returning to Earth
    When Specimen 2, 3, 4, and 5 were received, S. feltiae IJs were observed visually to assess whether they were alive, sluggish, or dead before removing the seals. Microgravity and corresponding Earth controls were inspected at the same time. There were several IJs visible prior to opening the tubes. Subsequently, we unsealed the tubes for Specimen 2, 3, and 4, removed half of the sand to a 10 cm petri dish and harvested the IJs by washing the sand three times with water. The IJs were then rinsed with MILLIQ water to test for dispersal behavior at Pheronym and sent to the Shapiro-Ilan’s lab for assessing symbiotic bacteria load. In case we missed any live IJs in Specimens 2, 3, and 4, we baited the other half of the sand with two G. mellonella larvae and incubated them at 22 ± 1 °C for three days before inspecting for dead or live insects. After that the samples were shipped to the Shapiro-Ilan’s lab for a second inspection and analysis of the bait insects for IJ presence.
    Specimen 5 was in polyacrylamide gel which was transparent, making visually scoring easy dead or alive and taking videos (Supplementary Video 1 and 2). After inspection, three ml of IJs with a density of 4000 IJs/ml were shipped in gel at ambient temperature overnight by FedEx to the Shapiro-Ilan lab to conduct infectivity experiments.
    Dispersal assays of Specimens 2 and 3
    Dispersal assays were conducted as described by Kaplan et al.14,15 with a few differences. Petri dishes (6 cm) containing 0.9% agar with a gel strength: ≥900 g/cm2 (Caisson Agar, Type I, Smithfield, UT) were used for the assays. Briefly, IJs from Specimens 2 and 3 were tested for dispersal immediately by rinsing three times with MILLIQ water after the harvest from sand. The assays were conducted in the afternoon between 2 and 3 PM. Approximately ~26–55 IJs (due to a limited number of IJs) in 10 µL of water were placed in the center of an agar plate with 6 cm diameter Petri dishes and counted. The assays were run for 30 min during which IJs were free to move on agar plates. After 30 min, IJs inside and outside the 1.3 cm ring were counted. IJs remaining inside the 1.3 cm ring were considered non-dispersed, and those outside were considered to have dispersed. Percentage dispersal was calculated as the number dispersed relative to the total number of IJs on the plate. Three replications for Specimens 2 (proxy Earth control) and Specimen 3 (treatment, microgravity), a total of six plates were analyzed.
    Infectivity after 33-day space flight, Specimen 5
    Infectivity was assessed based on procedures describe by Mbata et al.46. IJs were extracted from gel by dilution and centrifuged at 582g to concentrate them. A total of 960 IJs per replication were pipetted onto filter paper (Whatman No. 1) in a 0.35 mm Petri dish. A single greater wax moth, Galleria mellonella (L.) larva was added to each Petri dish. Insects were incubated in the dark for at 25 °C for 72 h and then dissected under a stereomicroscope to determine the number of invading IJs. There were five insects per replicate of Specimen 5 (total six replicates microgravity and Earth controls) and insect only (no IJs) control. A total of 35 insects were analyzed.
    Comparison of symbiotic bacterial load per nematode in Specimens 2 and 3
    Methods to compare symbiotic nematode bacteria load among treatments were based on those described by Kaya and Stock1,5. IJs were surface sterilized with 0.5% NaClO and then washed three times with sterile distilled water (centrifuging at 582g for 2 min between each wash). The final pellet was suspended in 0.5 ml. Single IJs were homogenized for 60 s with a sterile motor-driven polypropylene pestle and then transferred onto 60 mm nutrient agar plates. The plates were incubated at 25 °C. The number of bacterial colonies per IJ was assessed at 3 and 7 days.
    Statistical analysis
    For comparisons of infectivity, dispersal behavior, symbiotic bacteria load, and immune response the treatments effects (from space) were compared to Earth controls using Student’s t tests (SAS version 9.4, SAS 2002). Based on the inspection of residual plots, numerical data were log transformed prior to analysis47 (SAS, 2002).
    Ethics statement
    We use the invertebrate model system Steinernema feltiae and Galleria mellonella for this study, in accordance, the study was exempt from ethics committee approval.
    Reporting summary
    Further information on research design is available in the Nature Research Reporting Summary linked to this article. More

  • in

    Bacterial mock communities as standards for reproducible cytometric microbiome analysis

    1.
    Müller, S. & Nebe-von-Caron, G. Functional single-cell analyses: flow cytometry and cell sorting of microbial populations and communities. FEMS Microbiol. Rev. 34, 554–587 (2010).
    Article  Google Scholar 
    2.
    Günther, S. et al. Species-sorting and mass-transfer paradigms control managed natural metacommunities. Environ. Microbiol. 18, 4862–4877 (2016).
    Article  Google Scholar 

    3.
    Props, R., Monsieurs, P., Mysara, M., Clement, L. & Boon, N. Measuring the biodiversity of microbial communities by flow cytometry. Methods Ecol. Evol. 7, 1376–1385 (2016).
    Article  Google Scholar 

    4.
    Liu, Z. et al. Ecological stability properties of microbial communities assessed by flow cytometry. mSphere 3, e00564–17 (2018).
    CAS  PubMed  PubMed Central  Google Scholar 

    5.
    Liu, Z. et al. Neutral mechanisms and niche differentiation in steady-state insular microbial communities revealed by single cell analysis. Environ. Microbiol. 21, 164–181 (2019).
    CAS  Article  Google Scholar 

    6.
    De Vrieze, J., Boon, N. & Verstrate, W. Taking the technical microbiome into the next decade. Environ. Microbiol. 20, 1991–2000 (2018).
    Article  Google Scholar 

    7.
    Koch, C. et al. Cytometric fingerprinting for analyzing microbial intracommunity structure variation and identifying subcommunity function. Nat. Protoc. 8, 190–202 (2013).
    CAS  Article  Google Scholar 

    8.
    Mage, L. M. et al. Shape-based separation of synthetic microparticles. Nat. Mater. 18, 82–89 (2019).
    CAS  Article  Google Scholar 

    9.
    Müller, S. Modes of cytometric bacterial DNA pattern: a tool for pursuing growth. Cell Prolif. 40, 621–639 (2007).
    Article  Google Scholar 

    10.
    Ludwig, J., Höner zu Siederdissen, C., Liu, Z., Stadler, P. F. & Müller, S. flowEMMi: an automated model-based clustering tool for microbial cytometric data. BMC Bioinforma. 20, 643 (2019).
    CAS  Article  Google Scholar 

    11.
    Koch, C., Fetzer, I., Harms, H. & Müller, S. CHIC-an automated approach for the detection of dynamic variations in complex microbial communities. Cytom. A 83, 561–567 (2013).
    Article  Google Scholar 

    12.
    Liu, Z. & Müller, S. Bacterial community diversity dynamics highlight degrees of nestedness and turnover patterns. Cytom. Part A https://onlinelibrary.wiley.com/doi/10.1002/cyto.a.23965 (2020)

    13.
    Aghaeepour, N. et al. Critical assessment of automated flow cytometry data analysis techniques. Nat. Methods 10, 228–238 (2013).
    CAS  Article  Google Scholar 

    14.
    Peters, J. M. & Ansari, M. Q. Multiparameter flow cytometry in the diagnosis and management of acute leukemia. Arch. Pathol. Lab. Med. 135, 44–54 (2011).
    PubMed  Google Scholar 

    15.
    Bendall, S. C. et al. Single-cell mass cytometry of differential immune and drug responses across a human hematopoietic continuum. Science 332, 687–696 (2011).
    CAS  Article  Google Scholar 

    16.
    Spitzer, H. M. & Nolan, G. P. Mass cytometry: single cells, many features. Cell 165, 780–791 (2016).
    CAS  Article  Google Scholar 

    17.
    Overmann, J., Abt, B. & Sikorski, J. Present and future of culturing bacteria. Annu. Rev. Microbiol. 8, 711–730 (2017).
    Article  Google Scholar 

    18.
    Nayfach, S., Shi, Z. J., Seshadri, R., Pollard, K. S. & Kyrpides, N. C. New insights from uncultivated genomes of the global human gut microbiome. Nature 568, 505–510 (2019).
    CAS  Article  Google Scholar 

    19.
    Roesch, L. F. et al. Pyrosequencing enumerates and contrasts soil microbial diversity. ISME J. 1, 283–290 (2007).
    CAS  Article  Google Scholar 

    20.
    Singer, E. et al. Next generation sequencing data of a defined microbial mock community. Sci. Data 3, 160081 (2016).
    Article  Google Scholar 

    21.
    Hallmaier-Wacker, L. K., Lueert, S., Roos, C. & Knauf, S. The impact of storage buffer, DNA extraction method, and polymerase on microbial analysis. Sci. Rep. 8, 6292 (2018).
    Article  Google Scholar 

    22.
    Hardwick, S. A. et al. Synthetic microbe communities provide internal reference standards for metagenome sequencing and analysis. Nat. Commun. 9, 3096 (2018).
    Article  Google Scholar 

    23.
    Hornung, B. V. H., Zwittink, R. D. & Kuijper, E. J. Issues and current standards of controls in microbiome research. FEMS Microbiol. Ecol. 95, fiz045 (2019).
    CAS  Article  Google Scholar 

    24.
    Sze, M. A. & Schloss, P. D. The impact of DNA polymerase and number of rounds of amplification in PCR on 16S rRNA gene sequence data. mSphere 4, e00163–19 (2019).
    CAS  Article  Google Scholar 

    25.
    Clingenpeel, S., Clum, A., Schwientel, P., Rinke, C. & Woyke, T. Reconstructing each cell’s genome within complex communities—dream or reality? Front. Microbiol. 8, 771 (2015).
    Google Scholar 

    26.
    Stepanauskas, R. et al. Improved genome recovery and intergrated cell-size analyses of individual uncultured microbial cells and viral particles. Nat. Commun. 8, 84 (2017).
    Article  Google Scholar 

    27.
    De Bruin, O. M. & Birnboim, H. C. A method for assessing efficiency of bacterial cell disruption and DNA release. BMC Microbiol. 16, 197 (2016).
    Article  Google Scholar 

    28.
    Mie, G. Beiträge zur optik trüber medien, speziell kolloidaler metallösungen. Ann. Phys. 25, 377–445 (1908).
    CAS  Article  Google Scholar 

    29.
    Woyke, T., Doud, D. F. R. & Schulz, F. The trajectory of microbial single-cell sequencing. Nat. Methods 14, 1045–1054 (2017).
    CAS  Article  Google Scholar 

    30.
    Jahn, M. et al. Subpopulation-proteomics in prokaryotic populations. Curr. Opin. Biotech. 24, 79–87 (2013).
    CAS  Article  Google Scholar 

    31.
    Schloss, P. D. et al. Introducing mothur: open-source, platform-independent, community-supported software for describing and comparing microbial communities. Appl. Environ. Microbiol. 75, 7537–7541 (2009).
    CAS  Article  Google Scholar 

    32.
    Edgar, R. C., Haas, B. J., Clemente, J. C., Quince, C. & Knight, R. UCHIME improves sensitivity and speed of chimera detection. Bioinformatics 27, 2194–2200 (2011).
    CAS  Article  Google Scholar 

    33.
    Wickham, H. ggplot2: Elegant Graphics for Data Analysis (Springer, 2016).

    34.
    Lane, D. J. in Nucleic Acid Techniques in Bacterial Systematics (eds. Stackebrandt, E. & Goodfellow, M.) 115–175 (Wiley, 1991).

    35.
    Lambrecht, J. et al. Flow cytometric quantification, sorting and sequencing of methanogenic archaea based on F420 autofluorescence. Microb. Cell Fact. 16, 180 (2017).
    Article  Google Scholar 

    36.
    Besmer, M. D. et al. The feasibility of automated online flow cytometry for in-situ monitoring of microbial dynamics in aquatic ecosystems. Front. Microbiol 5, 265 (2014).
    Article  Google Scholar 

    37.
    Takahashi, S., Tomita, J., Nishioka, K., Hisada, T. & Nishijima, M. Development of a prokaryotic universal primer for simultaneous analysis of bacteria and archaea using next-generation sequencing. PLoS ONE 9, e105592 (2014).
    Article  Google Scholar 

    38.
    Herlemann, D. P. et al. Transitions in bacterial communities along the 2000 km salinity gradient of the Baltic Sea. ISME J. 5, 1571–1579 (2011).
    CAS  Article  Google Scholar  More