More stories

  • in

    Human-associated microbiota suppress invading bacteria even under disruption by antibiotics

    1.
    Blair JMA, Webber MA, Baylay AJ, Ogbolu DO, Piddock LJV. Molecular mechanisms of antibiotic resistance. Nat Rev Microbiol. 2015;13:42–51.
    CAS  Article  Google Scholar 
    2.
    Palmer AC, Kishony R. Understanding, predicting and manipulating the genotypic evolution of antibiotic resistance. Nat Rev Genet. 2013;14:243–8.
    CAS  Article  Google Scholar 

    3.
    Blaser MJ. Antibiotic use and its consequences for the normal microbiome. Science. 2016;352:544–5.
    CAS  Article  Google Scholar 

    4.
    Modi SR, Collins JJ, Relman DA. Antibiotics and the gut microbiota. J Clin Investig. 2014;124:4212–8.
    Article  Google Scholar 

    5.
    Lawley TD, Walker AW. Intestinal colonization resistance. Immunology. 2013;138:1–11.
    CAS  Article  Google Scholar 

    6.
    Libertucci J, Young VB. The role of the microbiota in infectious diseases. Nat Microbiol. 2019;4:35–45.
    CAS  Article  Google Scholar 

    7.
    Bhalodi AA, Engelen TSR, van, Virk HS, Wiersinga WJ. Impact of antimicrobial therapy on the gut microbiome. J Antimicrobial Chemother. 2019;74:i6–i15.
    CAS  Article  Google Scholar 

    8.
    Klümper U, Recker M, Zhang L, Yin X, Zhang T, Buckling A, et al. Selection for antimicrobial resistance is reduced when embedded in a natural microbial community. ISME J. 2019;13, 2927–37.
    Article  Google Scholar 

    9.
    Baumgartner M, Bayer F, Pfrunder-Cardozo KR, Buckling A, Hall AR. Resident microbial communities inhibit growth and antibiotic-resistance evolution of Escherichia coli in human gut microbiome samples. PLOS Biol. 2020;18:e3000465.
    CAS  Article  Google Scholar 

    10.
    Ianiro G, Tilg H, Gasbarrini A. Antibiotics as deep modulators of gut microbiota: between good and evil. Gut. 2016;65:1906–15.
    CAS  Article  Google Scholar 

    11.
    Spaulding CN, Klein RD, Schreiber HL, Janetka JW, Hultgren SJ. Precision antimicrobial therapeutics: The path of least resistance? npj Biofilms Microbiomes. 2018;4:1–7.
    Article  Google Scholar 

    12.
    Moya A, Ferrer M. Functional redundancy-induced stability of gut microbiota subjected to disturbance. Trends Microbiol. 2016;24:402–13.
    CAS  Article  Google Scholar 

    13.
    Jernberg C, Löfmark S, Edlund C, Jansson JK. Long-term impacts of antibiotic exposure on the human intestinal microbiota. Microbiology. 2010;156:3216–23.
    CAS  Article  Google Scholar 

    14.
    Stecher B, Chaffron S, Käppeli R, Hapfelmeier S, Freedrich S, Weber TC, et al. Like will to like: abundances of closely related species can predict susceptibility to intestinal colonization by pathogenic and commensal bacteria. PLoS Pathog. 2010;6:e1000711.
    Article  Google Scholar 

    15.
    Van Der Waaij D, Berghuis-de Vries JM, Lekkerkerk-Van Der Wees JEC. Colonization resistance of the digestive tract in conventional and antibiotic-treated mice. J Hyg. 1971;69:405–11.
    Article  Google Scholar 

    16.
    Kim S, Covington A, Pamer EG. The intestinal microbiota: antibiotics, colonization resistance, and enteric pathogens. Immunological Rev. 2017;279:90–105.
    CAS  Article  Google Scholar 

    17.
    Mathieu A, Fleurier S, Frénoy A, Dairou J, Bredeche MF, Sanchez-Vizuete P, et al. Discovery and function of a general core hormetic stress response in E. coli induced by sublethal concentrations of antibiotics. Cell Rep. 2016;17:46–57.
    CAS  Article  Google Scholar 

    18.
    Pamer EG. Resurrecting the intestinal microbiota to combat antibiotic-resistant pathogens. Science. 2016;352:535–8.
    CAS  Article  Google Scholar 

    19.
    Louca S, Polz MF, Mazel F, Albright MBN, Huber JA, O’Connor MI, et al. Function and functional redundancy in microbial systems. Nat Ecol Evolution. 2018;2:936–43.
    Article  Google Scholar  More

  • in

    Erosion reduces soil microbial diversity, network complexity and multifunctionality

    1.
    FAO, ITPS. Status of the World’s Soil Resources – Main report. Food and Agriculture Organization of the United Nations and Intergovernmental Technical Panel on Soils, 2015. Rome, Italy. https://www.fao.org/3/a-i5199e.pdf.
    2.
    UN (United Nations). Sustainable Development Goals [online]. 2015. https://www.un.org/sustainabledevelopment/sustainabledevelopment-goals/.

    3.
    FAO. 2019. Soil erosion: the greatest challenge for sustainable soil management. Rome: Food and Agriculture Organization of the United Nations; 2019. p. 104.
    Google Scholar 

    4.
    Pimentel D, Harvey C, Resosudarmo P, Sinclair K, Kurz D, McNair M, et al. Environmental and economic costs of soil erosion and conservation benefits. Science. 1995;267:1117–23.
    CAS  PubMed  Article  Google Scholar 

    5.
    Borrelli P, Robinson DA, Fleischer LR, Lugato E, Ballabio C, Alewell C, et al. An assessment of the global impact of 21st century land use change on soil erosion. Nat Commun. 2017;8:1–13.
    CAS  Article  Google Scholar 

    6.
    UN (United Nations). World Soil Day [online]. 2019. https://www.un.org/en/observances/world-soil-day.

    7.
    Van Oost K, Bakker MM. Soil productivity and erosion. In: Wall DH, Bardgett RD, Behan-Pelletier V, Herrick JE, Jones H, Ritz K, et al. (eds.). Soil ecology and ecosystem services. Oxford, UK: Oxford University Press; 2012. 301–14.

    8.
    Gregorich EG, Greer KJ, Anderson DW, Liang BC. Carbon distribution and losses: erosion and deposition effects. Soil Res. 1998;47:291–302.
    Google Scholar 

    9.
    Lal R, Pimentel D. Soil erosion: a carbon sink or source? Science. 2008;319:1040–2.
    CAS  PubMed  Article  Google Scholar 

    10.
    Mendonça R, Müller RA, Clow D, Verpoorter C, Raymond P, Tranvik LJ, et al. Organic carbon burial in global lakes and reservoirs. Nat Commun. 2017;8:1694.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    11.
    Quinton JN, Govers G, Van Oost K, Bardgett RD. The impact of agricultural soil erosion on biogeochemical cycling. Nat Geosci. 2010;3:311–4.
    CAS  Article  Google Scholar 

    12.
    Smith RW, Bianchi TS, Allison M, Savage C, Galy V. High rates of organic carbon burial in fjord sediments globally. Nat Geosci. 2015;8:450–U46.
    CAS  Article  Google Scholar 

    13.
    Van Oost K, Quine TA, Govers G, De Gryze S, Six J, Harden JW, et al. The impact of agricultural soil erosion on the global carbon cycle. Science. 2007;318:626–9.
    PubMed  Article  CAS  Google Scholar 

    14.
    Maestre FT, Quero JL, Gotelli NJ, Escudero A, Ochoa V, Delgado-Baquerizo M, et al. Plant species richness and ecosystem multifunctionality in global drylands. Science. 2012;335:214–8.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    15.
    Delgado-Baquerizo M, Maestre FT, Reich PB, Jeffries TC, Gaitan JJ, Encinar D, et al. Microbial diversity drives multifunctionality in terrestrial ecosystems. Nat Commun. 2016;7:10541.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    16.
    Fanin N, Gundale MJ, Farrell M, Ciobanu M, Baldock JA, Nilsson MC, et al. Consistent effects of biodiversity loss on multifunctionality across contrasting ecosystems. Nat Ecol Evol. 2018;2:269–78.
    PubMed  Article  Google Scholar 

    17.
    Garland G, Banerjee S, Edlinger A, Oliveira EM, Herzog C, Wittwer R, et al. A closer look at the functions behind ecosystem multifunctionality: a review. J Ecol. 2020, https://doi.org/10.1111/1365-2745.13511.

    18.
    Bardgett RD, Van Der Putten WH. Belowground biodiversity and ecosystem functioning. Nature. 2014;515:505–11.
    CAS  PubMed  Article  Google Scholar 

    19.
    Fierer N. Embracing the unknown: disentangling the complexities of the soil microbiome. Nat Rev Microbiol. 2017;15:579–90.
    CAS  PubMed  Article  Google Scholar 

    20.
    Wall H, Nielsen UN, Six J. Soil biodiversity and human health. Nature. 2015;528:69–76.
    CAS  PubMed  Article  Google Scholar 

    21.
    Saleem M, Hu J, Jousset A. More than the sum of its parts: microbiome biodiversity as a driver of plant growth and soil health. Annu Rev Ecol Evol Syst. 2019;50:145–68.
    Article  Google Scholar 

    22.
    Crowther TW, Van Den Hoogen J, Wan J, Mayes MA, Keiser AD, Mo L, et al. The global soil community and its influence on biogeochemistry. Science. 2019;365:eaav0550.
    CAS  PubMed  Article  Google Scholar 

    23.
    de Vries FT, Griffiths RI, Bailey M, Craig H, Girlanda M, Gweon HS, et al. Soil bacterial networks are less stable under drought than fungal networks. Nat Commun. 2018;9:3033.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    24.
    Zhou J, Deng Y, Luo F, He Z, Yang Y. Phylogenetic molecular ecological network of soil microbial communities in response to elevated CO2. MBio. 2011;2:e00122–11.
    PubMed  PubMed Central  Article  Google Scholar 

    25.
    Bragazza L, Parisod J, Buttler A, Bardgett RD. Biogeochemical plant–soil microbe feedback in response to climate warming in peatlands. Nat Clim Change. 2013;3:273–7.
    CAS  Article  Google Scholar 

    26.
    Crowther TW, Thomas SM, Maynard DS, Baldrian P, Covey K, Frey SD, et al. Biotic interactions mediate soil microbial feedbacks to climate change. Proc Natl Acad Sci USA. 2015;112:7033–8.
    CAS  PubMed  Article  Google Scholar 

    27.
    Maestre FT, Delgado-Baquerizo M, Jeffries TC, Eldridge DJ, Ochoa V, Gozalo B, et al. Increasing aridity reduces soil microbial diversity and abundance in global drylands. Proc Natl Acad Sci USA. 2015;112:15684–9.
    CAS  PubMed  Article  Google Scholar 

    28.
    Guo X, Feng J, Shi Z, Zhou X, Yuan M, Tao X, et al. Climate warming leads to divergent succession of grassland microbial communities. Nat Clim Change. 2018;8:813–8.
    Article  Google Scholar 

    29.
    Li Z, Tian D, Wang B, Wang J, Wang S, Chen H, et al. Microbes drive global soil nitrogen mineralization and availability. Glob Change Biol. 2019;25:1078–88.
    Article  Google Scholar 

    30.
    Wieder WR, Bonan GB, Allison SD. Global soil carbon projections are improved by modelling microbial processes. Nat Clim Change. 2013;3:909–12.
    CAS  Article  Google Scholar 

    31.
    Chen Q, Dong J, Zhu D, Hu H, Delgado-Baquerizo M, Ma Y, et al. Rare microbial taxa as the major drivers of ecosystem multifunctionality in long-term fertilized soils. Soil Biol Biochem. 2020;141:107686.
    CAS  Article  Google Scholar 

    32.
    Delgado-Baquerizo M, Reich PB, Trivedi C, Eldridge DJ, Abades S, Alfaro FD, et al. Multiple elements of soil biodiversity drive ecosystem functions across biomes. Nat Ecol Evol. 2020;4:210–20.
    PubMed  Article  Google Scholar 

    33.
    Wagg C, Schlaeppi K, Banerjee S, Kuramae EE, Van Der Heijden MGA. Fungal-bacterial diversity and microbiome complexity predict ecosystem functioning. Nat Commun. 2019;10:4841.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    34.
    Van der Heijden MGA, Bardgett RD, Van Straalen NM. The unseen majority: Soil microbes as drivers of plant diversity and productivity in terrestrial ecosystems. Ecol Lett. 2008;11:296–310.
    Article  Google Scholar 

    35.
    Barberán A, Bates ST, Casamayor EO, Fierer N. Using network analysis to explore co-occurrence patterns in soil microbial communities. ISME J. 2012;6:343–51.
    Article  CAS  Google Scholar 

    36.
    Banerjee S, Walder F, Büchi L, Meyer M, Held AY, Gattinger A, et al. Agricultural intensification reduces microbial network complexity and the abundance of keystone taxa in roots. ISME J. 2019;13:1722–36.
    PubMed  PubMed Central  Article  Google Scholar 

    37.
    Freilich MA, Wieters E, Broitman BR, Marquet PA, Navarrete SA. Species co-occurrence networks: Can they reveal trophic and non-trophic interactions in ecological communities? Ecology. 2018;99:690–9.
    PubMed  Article  Google Scholar 

    38.
    Fuhrman JA. Microbial community structure and its functional implications. Nature. 2009;459:193–9.
    CAS  PubMed  Article  Google Scholar 

    39.
    Banerjee S, Schlaeppi K, Van Der Heijden MGA. Keystone taxa as drivers of microbiome structure and functioning. Nat Rev Microbiol. 2018;16:567–76.
    CAS  PubMed  Article  Google Scholar 

    40.
    Herren CM, McMahon KD. Keystone taxa predict compositional change in microbial communities. Environ Microbiol. 2018;20:2207–17.
    PubMed  Article  Google Scholar 

    41.
    Ochoa-Hueso R, Collins SL, Delgado-Baquerizo M, Hamonts K, Pockman WT, Sinsabaugh RL, et al. Drought consistently alters the composition of soil fungal and bacterial communities in grasslands from two continents. Glob Change Biol. 2018;24:2818–27.
    Article  Google Scholar 

    42.
    Mabuhay JA, Nakagoshi N, Isagi Y. Influence of erosion on soil microbial biomass, abundance and community diversity. Land Degrad Dev. 2004;15:183–95.
    Article  Google Scholar 

    43.
    Li Z, Xiao H, Tang Z, Huang J, Nie X, Huang B, et al. Microbial responses to erosion-induced soil physico-chemical property changes in the hilly red soil region of southern China. Eur J Soil Biol. 2015;71:37–44.
    CAS  Article  Google Scholar 

    44.
    Hou S, Xin M, Wang LL, Jiang H, Li N, Wang Z. The effects of erosion on the microbial populations and enzyme activity in black soil of northeastern China. Acta Ecologica Sin. 2014;34:295–301.
    Article  Google Scholar 

    45.
    Zhang Y, Wu Y, Liu B, Zheng Q, Yin J. Characteristics and factors controlling the development of ephemeral gullies in cultivated catchments of black soil region, Northeast China. Soil Res. 2007;96:28–41.
    Google Scholar 

    46.
    Li H, Zhu H, Qiu L, Wei X, Liu B, Shao M. Response of soil OC, N and P to land-use change and erosion in the black soil region of the Northeast China. Agr Ecosyst Environ. 2020;302:107081.
    CAS  Article  Google Scholar 

    47.
    Zheng F. Effect of vegetation changes on soil erosion on the Loess Plateau. Pedosphere 2006;16:420–7.
    Article  Google Scholar 

    48.
    Page A, Miller R, Keeney D. Methods of Soil Analysis, Part 2. Chemical and Microbiological Properties. Madison, Wisconsin, American Society of Agronomy, Inc., Soil Science Society of America, Inc, 1982.

    49.
    Brookes P, Landman A, Pruden G, Jenkinson D. Chloroform fumigation and the release of soil nitrogen: a rapid direct extraction method to measure microbial biomass nitrogen in soil. Soil Biol Biochem. 1985;17:837–42.
    CAS  Article  Google Scholar 

    50.
    Lefcheck JS, Byrnes JEK, Isbell F, Gamfeldt L, Griffin JN, Eisenhauer N, et al. Biodiversity enhances ecosystem multifunctionality across trophic levels and habitats. Nat Commun. 2015;6:6936.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    51.
    Maestre FT, Castillo-Monroy AP, Bowker MA, Ochoa-Hueso R. Species richness effects on ecosystem multifunctionality depend on evenness, composition and spatial pattern. J Ecol. 2012;100:317–30.
    CAS  Article  Google Scholar 

    52.
    Wang Z, Zhang Q, Staley C, Gao H, Ishii S, Wei X, et al. Impact of long-term grazing exclusion on soil microbial community composition and nutrient availability. Biol Fertil Soils. 2019;55:121–34.
    CAS  Article  Google Scholar 

    53.
    Caporaso JG, Kuczynski J, Stombaugh J, Bittinger K, Bushman FD, Costello EK, et al. QIIME allows analysis of high-throughput community sequencing data. Nat Methods. 2010;7:335–6.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    54.
    Mueller RC, Paula FS, Mirza BS, Rodrigues JLM, Nuesslein K, Bohannan BJM. Links between plant and fungal communities across a deforestation chronosequence in the Amazon rainforest. ISME J. 2014;8:1548–50.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    55.
    Al-Ghalith GA, Hillmann B, Ang K, Shields-Cutler R, Knights D. SHI7 is a self-learning pipeline for multipurpose short-read DNA quality control. mSystems. 2018;3:e00202–17.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    56.
    Al-Ghalith GA, Montassier E, Ward HN, Knights D. NINJA-OPS: fast accurate marker gene alignment using concatenated ribosomes. PLoS Comput Biol. 2016;12:e1004658.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    57.
    Langmead B, Salzberg SL. Fast gapped-read alignment with Bowtie 2. Nat Methods. 2012;9:357–9.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    58.
    McDonald D, Price MN, Goodrich J, Nawrocki EP, DeSantis TZ, Probst A, et al. An improved Greengenes taxonomy with explicit ranks for ecological and evolutionary analyses of bacteria and archaea. ISME J. 2012;6:610–8.
    CAS  PubMed  Article  Google Scholar 

    59.
    Bray JR, Curtis JT. An ordination of the upland forest communities of southern Wisconsin. Ecol Monogr. 1957;27:326–49.
    Article  Google Scholar 

    60.
    Anderson MJ, Willis TJ. Canonical analysis of principal coordinates: a useful method of constrained ordination for ecology. Ecology. 2003;84:511–25.
    Article  Google Scholar 

    61.
    Oksanen J, Blanchet FG, Kindt R, Legendre P, Minchin PR, O’Hara RB, et al. Vegan: community ecology package. R package version 2.3-1. 2015, http://CRAN.R-project.org/package=vegan.

    62.
    Langfelder P, Horvath S. WGCNA: an R package for weighted correlation network analysis. BMC Bioinforma. 2008;9:559.
    Article  CAS  Google Scholar 

    63.
    Luo F, Zhong J, Yang Y, Scheuermann RH, Zhou J. Application of random matrix theory to biological networks. Phys Lett A. 2006;357:420–3.
    CAS  Article  Google Scholar 

    64.
    Benjamini Y, Krieger AM, Yekutieli D. Adaptive linear step-up procedures that control the false discovery rate. Biometrika. 2006;93:491–507.
    Article  Google Scholar 

    65.
    Csardi G, Nepusz T. The igraph software package for complex network research. InterJournal. Complex Syst. 2006;1695:1–9.
    Google Scholar 

    66.
    Ma B, Wang HZ, Dsouza M, Lou J, He Y, Dai ZM, et al. Geographic patterns of co-occurrence network topological features for soil microbiota at continental scale in eastern China. ISME J. 2016;10:1891–901.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    67.
    Berry D, Widder S. Deciphering microbial interactions and detecting keystone species with co-occurrence networks. Front Microbiol. 2014;5:219.
    PubMed  PubMed Central  Article  Google Scholar 

    68.
    Bastian M, Heymann S, Jacomy M. Gephi: an open source software for exploring and manipulating networks. ICWSM Conf. 2009;8:361–2.
    Google Scholar 

    69.
    Louca S, Parfrey LW, Doebeli M. Decoupling function and taxonomy in the global ocean microbiome. Science. 2016;353:1272–7.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    70.
    Qiu L, Zhu H, Liu J, Yao Y, Wang X, Rong G, et al. Soil erosion significantly reduces organic carbon and nitrogen mineralization in a simulated experiment. Agr Ecosyst Environ. 2021;307:107232.
    CAS  Article  Google Scholar 

    71.
    Crits-Christoph A, Robinson CK, Barnum T, Fricke WF, Davila AF, Jedynak B, et al. Colonization patterns of soil microbial communities in the Atacama Desert. Microbiome. 2013;1:28.
    PubMed  PubMed Central  Article  Google Scholar 

    72.
    Tiemann LK, Billings SA. Changes in variability of soil moisture alter microbial community C and N resource use. Soil Biol Biochem. 2011;43:1837–47.
    CAS  Article  Google Scholar 

    73.
    Banerjee S, Misra A, Sar A, Pal S, Chaudhury S, Dam B. Poor nutrient availability in opencast coalmine influences microbial community composition and diversity in exposed and underground soil profiles. Appl Soil Ecol. 2020;152:103544.
    Article  Google Scholar 

    74.
    Abu-Hamdeh NH, Reeder RC. Soil thermal conductivity: effects of density, moisture, salt concentration, and organic matter. Soil Sci Soc Am J. 2000;64:1285–90.
    CAS  Article  Google Scholar 

    75.
    Bajracharya RM, Lal R, Kimble JM. Diurnal and seasonal CO2-C flux from soil as related to erosion phases in central Ohio. Soil Sci Soc Am J. 2000;64:286–93.
    CAS  Article  Google Scholar 

    76.
    Liang Y, Lal R, Guo S, Liu R, Hu Y. Impacts of simulated erosion and soil amendments on greenhouse gas fluxes and maize yield in Miamian soil of central Ohio. Sci Rep. 2018;8:520.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    77.
    Van Der Voort M, Kempenaar M, Van Driel M, Raaijmakers JM, Mendes R. Impact of soil heat on reassembly of bacterial communities in the rhizosphere microbiome and plant disease suppression. Ecol Lett. 2016;19:375–82.
    PubMed  Article  PubMed Central  Google Scholar 

    78.
    García-Palacios P, Vandegehuchte ML, Shaw EA, Dam M, Post KH, Ramirez KS, et al. Are there links between responses of soil microbes and ecosystem functioning to elevated CO2, N deposition and warming? A global perspective. Glob Change Biol. 2015;21:1590–1600.
    Article  Google Scholar 

    79.
    Karimi B, Terrat S, Dequiedt S, Saby NPA, Horriguel W, Lelievre M, et al. Biogeography of soil bacteria and archaea across France. Sci Adv. 2018;4:eaat1808.
    PubMed  PubMed Central  Article  Google Scholar 

    80.
    Janssen PH. Identifying the dominant soil bacterial taxa in libraries of 16S rRNA and 16S rRNA genes. Appl Environ Microb. 2006;72:1719–28.
    CAS  Article  Google Scholar 

    81.
    Spain AM, Krumholz LR, Elshahed MS. Abundance, composition, diversity and novelty of soil Proteobacteria. ISME J. 2009;3:992–1000.
    CAS  PubMed  Article  Google Scholar 

    82.
    Zhang C, Liu G, Xue S, Wang G. Soil bacterial community dynamics reflect changes in plant community and soil properties during the secondary succession of abandoned farmland in the Loess Plateau. Soil Biol Biochem. 2016;97:40–49.
    CAS  Article  Google Scholar 

    83.
    Wolińska A, Kuzniar A, Zielenkiewicz U, Izak D, Szafranek-Nakonieczna A, Banach A, et al. Bacteroidetes as a sensitive biological indicator of agricultural soil usage revealed by a culture-independent approach. Appl Soil Ecol. 2017;119:128–37.
    Article  Google Scholar 

    84.
    DeBruyn LM, Nixon LT, Fawaz MN, Johnson AM, Radosevich M. Global biogeography and quantitative seasonal dynamics of gemmatimonadetes in soil. Appl Environ Microb. 2011;77:6295–6300.
    CAS  Article  Google Scholar 

    85.
    Bouskill NJ, Lim HC, Borglin S, Salve R, Wood TE, Silver WL, et al. Pre-exposure to drought increases the resistance of tropical forest soil bacterial communities to extended drought. ISME J. 2013;7:384–94.
    CAS  PubMed  Article  Google Scholar 

    86.
    Naylor D, DeGraaf S, Purdom E, Coleman-Derr D. Drought and host selection influence bacterial community dynamics in the grass root microbiome. ISME J. 2017;11:2691–704.
    PubMed  PubMed Central  Article  Google Scholar 

    87.
    Santos-Medellin C, Edwards J, Liechty Z, Nguyen B, Sundaresan V. Drought stress results in a compartment-specific restructuring of the rice root-associated microbiomes. mBio. 2017;8:e00764–17.
    PubMed  PubMed Central  Article  Google Scholar 

    88.
    Chowdhury TR, Lee JY, Bottos EM, Brislawn CJ, White RA, Bramer LM, et al. Metaphenomic responses of a native prairie soil microbiome to moisture perturbations. mSystems. 2019;4:e00061–19.
    Google Scholar 

    89.
    Mickan BS, Abbott LK, Solaiman ZM, Mathes F, Siddique KHM, Jenkins SN. Soil disturbance and water stress interact to influence arbuscular mycorrhizal fungi, rhizosphere bacteria and potential for N and C cycling in an agricultural soil. Biol Fert Soils. 2019;55:53–66.
    CAS  Article  Google Scholar 

    90.
    Van Horn DJ, Okie JG, Buelow HN, Gooseff MN, Barrett JE, Takacs-Vesbach CD. Soil microbial responses to increased moisture and organic resources along a salinity gradient in a polar desert. Appl Environ Microb. 2014;80:3034–43.
    Article  CAS  Google Scholar 

    91.
    Kielak A, Pijl AS, Van Veen JA, Kowalchuk GA. Phylogenetic diversity of Acidobacteria in a former agricultural soil. ISME J. 2009;3:378–82.
    CAS  PubMed  Article  Google Scholar 

    92.
    Fierer N, Lauber CL, Ramirez KS, Zaneveld J, Bradford MA, Knight R. Comparative metagenomic, phylogenetic and physiological analyses of soil microbial communities across nitrogen gradients. ISME J. 2012;6:1007–17.
    CAS  PubMed  Article  Google Scholar 

    93.
    Wolińska A, Kuzniar A, Zielenkiewicz U, Banach A, Blaszczyk M. Indicators of arable soils fatigue Bacterial – families and genera: a metagenomic approach. Ecol Indic. 2018;93:490–500.
    Article  Google Scholar 

    94.
    Yang F, Niu KC, Collins CG, Yan XB, Ji YG, Ling N. Grazing practices affect the soil microbial community composition in a Tibetan alpine meadow. Land Degrad Dev. 2019;30:49–59.
    Article  Google Scholar 

    95.
    Vitousek PM, Menge DNL, Reed SC, Cleveland CC. Biological nitrogen fixation: rates, patterns and ecological controls in terrestrial ecosystems. Philos Trans R Soc B Biol Sci. 2013;368:1621.
    Article  CAS  Google Scholar 

    96.
    Fan KK, Delgado-Baquerizo M, Guo XS, Wang DZ, Wu YY, Zhu M, et al. Suppressed N fixation and diazotrophs after four decades of fertilization. Microbiome. 2019;7:143.
    PubMed  PubMed Central  Article  Google Scholar 

    97.
    Ryu MH, Zhang J, Toth T, Khokhani D, Geddes BA, Mus F, et al. Control of nitrogen fixation in bacteria that associate with cereals. Nat Microbiol. 2020;5:314–30.
    CAS  PubMed  Article  Google Scholar 

    98.
    Banerjee S, Kirkby CA, Schmutter D, Bissett A, Kirkegaard JA, Richardson AE. Network analysis reveals functional redundancy and keystone taxa amongst bacterial and fungal communities during organic matter decomposition in an arable soil. Soil Biol Biochem. 2016;97:188–98.
    CAS  Article  Google Scholar 

    99.
    Guo J, Ling N, Chen Z, Xue C, Li L, Liu L, et al. Soil fungal assemblage complexity is dependent on soil fertility and dominated by deterministic processes. N Phytol. 2020;226:232–43.
    Article  Google Scholar 

    100.
    Qi G, Ma G, Chen S, Lin C, Zhao X. Microbial network and soil properties are changed in bacterial wilt-susceptible soil. Appl Environ Microb. 2019;85:e00162–19.
    CAS  Google Scholar 

    101.
    Xue L, Ren H, Brodribb TJ, Wang J, Yao X, Li S. Long term effects of management practice intensification on soil microbial community structure and co-occurrence network in a non-timber plantation. For Ecol Manag. 2020;459:117805.
    Article  Google Scholar 

    102.
    Ling N, Zhu C, Xue C, Chen H, Duan YH, Peng C. Insight into how organic amendments can shape the soil microbiome in long-term field experiments as revealed by network analysis. Soil Biol Biochem 2016;99:137–49.
    CAS  Article  Google Scholar 

    103.
    Deng Y, Jiang YH, Yang YF, He ZL, Luo F, Zhou JZ. Molecular ecological network analyses. BMC Bioinforma. 2012;13:113.
    Article  Google Scholar 

    104.
    Bai YX, She WW, Miao L, Qin SG, Zhang YQ. Soil microbial interactions modulate the effect of Artemisia ordosica on herbaceous species in a desert ecosystem, northern China. Soil Biol Biochem. 2020;150:108013.
    CAS  Article  Google Scholar 

    105.
    Szoboszlay M, Dohrmann AB, Poeplau C, Don A, Tebbe CC. Impact of land-use change and soil organic carbon quality on microbial diversity in soils across Europe. FEMS Microbiol Ecol. 2017;93:fix146.
    Article  CAS  Google Scholar 

    106.
    Marcos MS, Bertiller MB, Olivera NL. Microbial community composition and network analyses in arid soils of the Patagonian Monte under grazing disturbance reveal an important response of the community to soil particle size. Appl Soil Ecol. 2019;138:223–32.
    Article  Google Scholar 

    107.
    Hamamura N, Olson SH, Ward DM, Inskeep WP. Microbial population dynamics associated with crude-oil biodegradation in diverse soils. Appl Environ Microb. 2006;72:6316–24.
    CAS  Article  Google Scholar 

    108.
    Acosta‐Martínez V, Cotton J, Gardner T, Moore‐Kucera J, Zak J, Wester D, et al. Predominant bacterial and fungal assemblages in agricultural soils during a record drought/heat wave and linkages to enzyme activities of biogeochemical cycling. Appl Soil Ecol. 2014;84:69–82.
    Article  Google Scholar 

    109.
    Peng M, Jia H, Wang Q. The effect of land use on bacterial communities in Saline-Alkali soil. Curr Microbiol. 2017;74:325–33.
    CAS  PubMed  Article  Google Scholar 

    110.
    Navarrete AA, Tsai SM, Mendes LW, Faust K, de Hollander M, Cassman NA, et al. Soil microbiome responses to the short-term effects of Amazonian deforestation. Mol Ecol. 2015;24:2433–48.
    CAS  PubMed  Article  Google Scholar 

    111.
    Byers AK, Condron L, Donavan T, O’Callaghan M, Patuawa T, Waipara N, et al. Soil microbial diversity in adjacent forest systems—contrasting native, old growth kauri (Agathis australis) forest with exotic pine (Pinus radiata) plantation forest. FEMS Microbiol Ecol. 2020;96:fiaa047.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    112.
    Schmidt MWI, Torn MS, Abiven S, Dittmar T, Guggenberger G, Janssens IA, et al. Trumbore, Persistence of soil organic matter as an ecosystem property. Nature. 2011;478:49–56.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    113.
    Lehmann J, Kleber M. The contentious nature of soil organic matter. Nature. 2015;528:60–68.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    114.
    Buzzard V, Michaletz ST, Deng Y, He Z, Ning D, Shen L, et al. Continental scale structuring of forest and soil diversity via functional traits. Nat Ecol Evol. 2019;3:1298–308.
    PubMed  Article  PubMed Central  Google Scholar 

    115.
    Wei X, Shao M, Gale W, Li L. Global pattern of soil carbon losses due to the conversion of forests to agricultural land. Sci Rep. 2014;4:4062.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    116.
    Delgado‐Baquerizo M, Reith F, Dennis PG, Hamonts K, Powell JR, Young A, et al. Ecological drivers of soil microbial diversity and soil biological networks in the Southern Hemisphere. Ecology. 2018;99:583–96.
    PubMed  Article  PubMed Central  Google Scholar 

    117.
    Delgado-Baquerizo M, Oliverio AM, Brewer TE, Benavent-González A, Eldridge DJ, Bardgett RD, et al. A global atlas of the dominant bacteria found in soil. Science. 2018;359:320–5.
    CAS  PubMed  Article  Google Scholar 

    118.
    Nottingham AT, Fierer N, Turner BL, Whitaker J, Ostle NJ, McNamara NP, et al. Microbes follow Humboldt: temperature drives plant and soil microbial diversity patterns from the Amazon to the Andes. Ecology. 2018;99:2455–66.
    PubMed  PubMed Central  Article  Google Scholar 

    119.
    Zhou J, Deng Y, Shen L, Wen C, Yan Q, Ning D, et al. Temperature mediates continental-scale diversity of microbes in forest soils. Nat Commun. 2016;7:12083.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    120.
    Allison SD, Martiny JBH, et al. Resistance resilience, and redundancy in microbial communities. Proc Natl Acad Sci USA. 2008;105:11512–9.
    CAS  PubMed  Article  Google Scholar 

    121.
    Hartmann M, Niklaus PA, Zimmermann S, Schmutz S, Kremer J, Abarenkov K, et al. Resistance and resilience of the forest soil microbiome to logging-associated compaction. ISME J. 2014;8:226–44.
    CAS  PubMed  Article  Google Scholar  More

  • in

    Easing COVID-19 lockdown measures while protecting the older restricts the deaths to the level of the full lockdown

    Overview
    In earlier work34, epidemiological models are broadly divided into two large categories, called forecasting and mechanistic. The former models fit a specific curve to the data and then attempt to predict the dynamics of the quantity under consideration. The most well known mechanistic models are the SIR-type models. As noted by Holmadahl and Buckee34, the mechanistic models involve substantially more complicated mathematical machinary than the forecasting models, but they have the advantage that they can make predictions even when the relevant circumstances change. In our case, since our goal is to make predictions after the situation changes due to the lifting of the lockdown measures, we need to consider a mechanistic model. However, it is widely known that the main limitation of mechanistic models is the difficulty of determining the parameters specifying such models. In this direction, a methodological advance was presented by the authors35, filling an important gap in the relevant literature: it was shown in35 that from the knowledge of the most reliable data of the epidemic in a given country, namely the cumulative number of deaths, it is possible to determine suitable combinations of the constant parameters (of the original model) which specify the differential equation characterizing the death dynamics. Furthermore, a robust numerical algorithm was presented for obtaining these parameters. One of these constants, denoted by c, is particularly important for the analysis of the effect of easing the lockdown conditions, because it is proportional to the number of contacts between asymptomatic individuals that are infected by SARS-CoV-2 and susceptible ones. Specifically, as the equations presented below will indicate, this coefficient is measured in units of inverse population (where the population represents the number of individuals to which we assign no units) times inverse days. This constant reflects the probability of infection given a contact which is proportional to the viral load (i.e., the viral concentration in the respiratory-tract fluid) of expelled respiratory droplets36. Easing the lockdown will lead to an increase of the value of this constant. Thus, in order to quantify this effect we assumed that the post-lockdown situation could be described by the same model but with c multiplied by an integer number (zeta), such as (zeta =2), or 3, etc. Assuming a fixed viral load emission (i.e., no face mask or similar protective measures), this would be tantamount to doubling or tripling the number of contacts per day. To put things into perspective, it is relevant to mention here that in the relevant literature a ballpark estimate for daily contacts of an individual is about 13.437.
    We first applied the above algorithm to the case of the COVID-19 epidemic in Greece. However, the novelty and the main interest of the present work consists of the extension and application of the above methodology to two subpopulations. This situation is significantly more complicated than that of2 and is described by 12 ODEs involving 18 parameters (details are discussed in the “Methods” section). Using this extended formulation, we analysed the effect of easing the lockdown measures under two distinct possible scenarios: in the first, we examined what would happen if the interactions between older persons, namely persons above 40 years of age, as well as between older and younger persons, namely those below 40, continue to be dictated by the same restrictions as those of the lockdown period. However, we assumed that the interaction among the young was progressively more free. In the second case, we analysed the effect of easing the lockdown measures in the entire population without distinguishing the older from the young. In principle, the effect on deaths in the above two scenarios could be analyzed by the extension of the rigorous results of35. However, due to the sparsity of the deaths data (especially for the younger population), this approach is practically not possible at present. Thus, we supplemented the data for deaths for the two subpopulations with data for the cumulative numbers of reported infected.
    Using four sets of data, namely the number of deaths and the number of reported infected for the older and the younger population we found that the above two alternatives would result in very different outcomes: in the first case, the total number of deaths of the two sub-populations and the number of total infections would be relatively small. In the second case, these numbers would be prohibitively high. Specifically, in the case of Greece, if the lockdown was to be continued indefinitely, our analysis suggests that the total numbers of deaths and infections would finally be around 165 and 2550, respectively. These numbers would remain essentially the same even if the lockdown measures for the interaction between the young people were eased substantially, provided that the interactions of older-older and older-young would remain the same as during the lockdown period. For example, even if the parameter measuring the effect of the lockdown restrictions on the young-young interactions were increased fourfold, the number of deaths and infections would be (according to the model extrapolation) 184 and 3585, respectively. On the other hand, even if the parameters characterizing all three interactions were increased only threefold, the relevant numbers would be 48144 and 1283462. It is clear that the latter numbers are prohibitive, suggesting that a generic release of the lockdown may be catastrophic.
    In our view, the explanations provided in the “Methods” section for the assumptions of our model, which show that these assumptions are typical in the standard epidemiological models, substantiate the qualitative conclusions (and notes of caution) regarding the impact of the above two different types of exit policies. This may provide a sense of how a partial restoration of regular life activities can be achieved without catastrophic consequences, while the race for pharmacological or vaccine-based interventions that will lead to an end of the current pandemic is still ongoing. Importantly, we also offer some caveats emphasizing the qualitative nature of our conclusions and possible factors that may substantially affect the actual outcome of the lifting of lockdown measures.
    Model setup: single population versus two age groups
    We divide the population in two subpopulations, the young (y) and the older (o). In order to explain the basic assumptions of our model we first consider a single population, and then discuss the needed modifications in our case which involves two subpopulations. Let E(t) denote the exposed (but not infectious) population. An individual in this population, after a median 4-day period (required for incubation — see e.g.38) will either become sick or will be asymptomatic; an interval of 3-10 days captures 98% of the cases. The sick (infected) and asymptomatic populations will be denoted, respectively, by I(t) and A(t). The rate at which an exposed person becomes asymptomatic is denoted by a; this means that each day aE(t) persons leave the exposed population and enter the asymptomatic population. Similarly, each day sE(t) leave the exposed population and enter the sick population. These processes, as well as the subsequent movements are depicted in the flowchart of Fig. 1.
    Figure 1

    Flowchart of the populations considered in the model and the rates of transformation between them. The corresponding dynamical equations are Eqs. (1)–(6).

    Full size image

    The asymptomatic individuals recover with a rate (r_1), i.e., each day (r_1A(t)) leave the asymptomatic population and enter the recovered population, which is denoted by R(t). The sick individuals either recover with a rate (r_2) or they become hospitalized, H(t), with a rate h. In turn, the hospitalized patients also have two possible destinations; either they recover with a rate (r_3), or they become deceased, D(t), with a rate d.
    It is straightforward to write the above statements in the language of mathematics; this gives rise to the equations (1)–(5) below:

    $$begin{aligned} frac{dA}{dt}= a E – r_1 A end{aligned}$$
    (1)

    $$begin{aligned} frac{dI}{dt}= s E – (h + r_2) I end{aligned}$$
    (2)

    $$begin{aligned} frac{dH}{dt}= h I – (r_3+d) H end{aligned}$$
    (3)

    $$begin{aligned} frac{dR}{dt}= r_1 A + r_2 I + r_3 H end{aligned}$$
    (4)

    $$begin{aligned} frac{dD}{dt}= d H end{aligned}$$
    (5)

    $$begin{aligned} frac{dE}{dt}= c left[ T – (E+I+A+H+R+D)right] left( A + b Iright) – (a+s) E end{aligned}$$
    (6)

    It is noted that our model is inspired by various expanded versions of the classic SIR model adapted to the particularities of COVID-19 (such as the key role of the asymptomatically infected). It is, in particular, inspired by, yet not identical with that of14. In order to complete the system of equations (1)–(6), it is necessary to describe the mechanism via which a person can become infected. For this purpose we adopt the standard assumptions made in the typical epidemiological models, such as the SIR (susceptible, infected, recovered) model: let T denote the total population and let c characterize the number of contacts per day made by an individual with the capacity to infect (c is thought of as being normalized by T). Such a person belongs to I, A or H. However, for simplicity we assume that the hospitalized population cannot infect; this assumption is based on two considerations: first, the strict protective measures taken at the hospital, and second, the fact that hospitalized patients are infectious only for part of their stay in the hospital. The latter fact is a consequence of the relevant time scales of virus shedding in comparison to the time to hospitalization and the duration of hospital stay. The asymptomatic individuals are (more) free to interact with others, whereas the (self-isolating) sick persons are not. Thus, we use c to characterize the contacts of the asymptomatic persons and b to indicate the different infectiousness (due to reduced contacts/self-isolation) of the sick in comparison to the asymptomatic individuals.
    The number of people available to be infected (i.e., the susceptible population) is (T-(E+I+A+H+R+D)). Indeed, the susceptible individuals consist of the total population minus all the individuals that are going or have gone through the course of some phase of infection, namely they either bear the infection at present ((E+A+I+H)) or have died from COVID-19 (D) or are assumed to have developed immunity to COVID-19 due to recovery (R). Hence, if we call the total initial individuals T, this susceptible population is given by the expression written earlier. The rate by which each day individuals enter E is given by the product of the above expression with (c(A+bI)). At the same time, as discussed earlier, every day ((a+s)E) persons leave the exposed population. It is relevant to note here that within this simpler model, it is possible to calculate the basic reproduction number (R_0), which is a quantity of substantial value in epidemiological studies32,33. In this model, this can be found to be33:

    $$begin{aligned} R_0=frac{c T}{a+s}left[ frac{a}{r_1} + frac{b s}{r_2+h} right] . end{aligned}$$
    (7)

    This will be useful below for the purposes of finding the change in c (under lockdown) needed in order for transmission to cross the threshold of (R_0=1) and thus to lead to growth of the epidemic. In the particular case of the data shown in Table 1, (R_0=0.4084), in accordance with the lockdown situation associated with a controlled epidemic.
    It is straightforward to modify the above model so that it can describe the dynamics of the older and younger subpopulations. Each subpopulation satisfies the same set of equations as those described above, except for the last equation which is modified as follows: the people available to be infected in each subpopulation are described by the expression given above where T, E, I, A, H, R, D have the superscripts (^o) or (^y), denoting older and young, respectively; (A+bI) is replaced in both cases by (A^o+A^y+b(I^o+I^y)) where for simplicity we have assumed that the infectiousness of the older and the young is the same. We have already considered the implications of the generalisation of the above model by allowing different parameters to describe the interaction of the older and young populations; this will be discussed in the “Methods” section. In what follows, we will discuss the results of this simpler “isotropic” interaction model.
    Quantitative model findings
    The parameters of the model are given in the flowchart of Fig. 1. Naturally, for the two-age model considered below, there is one set of such parameters associated with the younger population and one associated with the older one. The optimization routine used for the identification of these parameters is explained in detail in the “Methods” section. The parameters resulting from this optimization for the single population model are shown in Table 1, whereas for each of the two populations are given in Table 2. Clearly, many of these parameters are larger for the older population in comparison to the young, leading to a larger number of both infections and deaths in the older than in the young population.
    Table 1 Optimized model parameters for the single population model, and the variation interval of each parameter within the optimization process (for further details, see “Methods” section).
    Full size table

    Table 2 Optimized (isotropic) model parameters for the young and older populations, and the variation interval of each parameter within the optimization process (for further details, see “Methods” section).
    Full size table

    Support for the validity of our model is presented in Fig. 2, which depicts its comparison (using the above optimized parameters) with the available data. The situation corresponding to keeping the lockdown conditions indefinitely, is the one illustrated in Fig. 2. In this case, the number of deaths and cumulative infections rapidly reaches a plateau, indicating the elimination of the infection. Here, we have optimized the model on the basis of data used from Greece39 between April 3rd and May 4th. It is noted that daily updates occurred at 3pm for the country of Greece, hence it is not clear up to what time the data are collected that are included in the daily report. We have assumed that the data reflect the infections and deaths present on that particular day. This possibly shifts the starting point of our count by a few hours, but should not change the overall result trends.
    We next explain the implications of the model when different scenarios of ‘exit’ from the lockdown state are implemented. The relevant results are illustrated in Figs. 3, 4 and the essential conclusions are summarized in Table 3 for the numbers of deaths and cumulative infections, respectively. First, we need to explain the meaning of the parameter (zeta) appearing in the above tables: this parameter reflects the magnitude of the easing of the lockdown restrictions. Indeed, since the main effect of the lessening of these restrictions is that the number of contacts increases, we model the effect of easing the lockdown restrictions by multiplying the parameter c with a factor that we refer to as (zeta). The complete lockdown situation corresponds to (zeta)=1; the larger the value of (zeta), the lesser the restrictions imposed on the population. By employing the above quantitative measure of easing the lockdown restrictions, we consider in detail two distinct scenarios. In the first, which corresponds to the top rows of the Figures 3 and 4, we only allow the number of contacts of “young individuals with young individuals” (corresponding to the parameter (c^{yy}) mentioned in the “Methods” section) to be multiplied by the factor (zeta). This means that the lockdown measures are eased only with respect to the interaction of young individuals with other young individuals, while the interactions of the young individuals with the older ones, as well as the interactions among older individuals remain in the lockdown state. In the second scenario, corresponding to the bottom rows of the Figures 3 and 4, the restrictions of the lockdown are simultaneously eased in both the young and the older population; in this case all contacts are increased by the factor (zeta). It is noted that while we change c by this factor, we maintain the product cb at its previous value (i.e., we concurrently transform (crightarrow zeta c) and (brightarrow b/zeta)) considering that the sick still operate under self-isolation conditions and thus do not accordingly increase their number of contacts.
    Table 3 Deaths D(t) and cumulative infections C(t) in the case of increasing of the number of contacts by (zeta). The second and fourth columns refer to the case for which the lockdown measures are eased for the young population, whereas the third and fifth column refer to the one where this occurs for both the young and older populations.
    Full size table

    Figure 2

    Evolution of the current situation of deaths D(t) (left) and cumulative infections C(t) (right) in Greece, under the case of an indefinite continuation of the lockdown conditions. In this and all the figures that follow, the blue curve corresponds to the young population, while the red curve to the older population. The data for Greece from the 3rd of April to the 4th of May 2020 are depicted by dots. For the latter, alternate colors have been used (i.e., blue dots for the older population and red for the younger for clearer visualization).

    Full size image

    Fig. 3 corresponds to the case where the parameter (zeta) associated with the number of contacts between susceptible and asymptomatic individuals doubles. In this case, as also shown in Table 3, the situation does not worsen in a dramatic way. In particular, the number of deaths increases by 1, whereas the cumulative infections only increase by the small number of 58. In the second scenario where the number of contacts is doubled for both the young and the older populations, we find slightly larger (but not totally catastrophic) effects: the number of deceased individuals increases by 58 and the total number of infections grows by 1550.
    Figure 3

    Again the deaths D(t) and the cumulative infections C(t) are given for the case where the c factor (characterizing the number of contacts) amongst young individuals is doubled, but those of the older individuals (and of the young-older interaction) are kept fixed. This is shown in the top panels. In the bottom panels, the c’s of both young and old individuals are doubled.

    Full size image

    The situation becomes far more dire when the number of contacts is multiplied by a factor of 3 for both the young and older populations, meaning that the lockdown restrictions are eased significantly for the entire population. As shown in Table 3 and in Fig. 4, if the c’s of the young population only are multiplied by a factor of 3, then the deaths are increased by 3 and the infections by 198 (black line in the Figure and 3rd row of the Tables). This pales by comparison to the dramatic scenario when the c’s associated with both the young and older sub-populations are multiplied by 3; in this case, the number of deaths jumps dramatically to 48144, while the number of infections is a staggering 1283462, growing by about 500 times.
    Figure 4

    Same as reported in Fig. 3 but now where the contacts are multiplied by factors 3, 4 and 5. Full (dashed) lines hold for the young (older) population.

    Full size image

    An example corroborating the above qualitative trend can also be found in Fig. 4 and in the 4th and 5th rows of Table 3. Here, for e.g. (zeta =5), even the effect of releasing solely the young population leads to very substantial increases, namely to 6044 deaths and 306219 infections although of course it is nowhere near the scenarios of releasing both young and older populations. In the second scenario, the numbers are absolutely daunting: using the parameters of Table 2 we find that the number of deaths jumps to 83274 and the number of cumulative infections to 2221296.
    Figure 5

    Hospitalizations when only the young population (left) or both the young and older (right) population are released. Full (dashed) lines hold for the young (older) population.

    Full size image

    Finally, we show the prediction of the easing measures in the hospitalizations (i.e. daily occupied beds in hospitals). This is a crucial point to assess in order that the health system does not collapse because of COVID-19 patients. Figure 5 shows these trends for the above mentioned values of (zeta). In the case of releasing solely the young population (see left panel of the Figure), it is observed that the number of hospitalizations decreases monotonically except for (zeta =5), where the hospitalization peak is 523 for the young population and 1426 for the older one (values that are affordable by Greek health system); however, if both the young and older population are released (see right panel of the Figure), there is a monotonically decreasing behaviour only for (zeta =1) and 2. For higher (zeta) we observe that the height of the peak obviously increases with (zeta), while this peak also occurs earlier when the number of contacts is increased; for instance, for (zeta =3), the hospitalization peak number of the young population is 3844 whereas this value is 37030 for the older one, numbers that are, unfortunately, unaffordable for the Greek health system. These figures grow even further to 16869 and 163648 if (zeta =5).
    In light of the above results, the significance of preserving the lockdown restrictions of the sensitive groups of the older population is naturally emerging. It can be seen that in the case where the number of contacts is roughly doubled, the behavior of release of young or young and older individuals is not dramatic (although even in this case releasing only the young population is, of course, preferable). Nevertheless, a more substantial release of the young population is still not catastrophic. On the other hand, the higher rates of infection, hospitalization and proneness to death of senior individuals may bring about highly undesirable consequences, should both the young and older members of the population be allowed to significantly increase (by 3 times or more) their number of contacts. More

  • in

    Small phytoplankton contribute greatly to CO2-fixation after the diatom bloom in the Southern Ocean

    1.
    Longhurst A, Sathyendranath S, Platt T, Caverhill C. An estimate of global primary production in the ocean from satellite radiometer data. J Plankton Res. 1995;17:1245–71.
    Article  Google Scholar 
    2.
    Field CB, Behrenfeld MJ, Randerson JT, Falkowski P. Primary production of the biosphere: integrating terrestrial and oceanic components. Science. 1998;281:237–40.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    3.
    Falkowski PG, Raven JA. Aquatic photosynthesis. Princeton, NJ: Princeton University Press; 2013.

    4.
    Falkowski PG, Barber RT, Smetacek V. Biogeochemical controls and feedbacks on Ocean primary production. Science. 1998;281:200–6.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    5.
    Palmer JR, Totterdell IJ. Production and export in a global ocean ecosystem model. Deep Sea Res Part Oceanogr Res Pap. 2001;48:1169–98.
    CAS  Article  Google Scholar 

    6.
    Legendre L, Rivkin RB. Fluxes of carbon in the upper ocean: regulation by food-web control nodes. Mar Ecol Prog Ser. 2002;242:95–109.
    Article  Google Scholar 

    7.
    Guidi L, Stemmann L, Jackson GA, Ibanez F, Claustre H, Legendre L, et al. Effects of phytoplankton community on production, size, and export of large aggregates: a world-ocean analysis. Limnol Oceanogr. 2009;54:1951–63.
    Article  Google Scholar 

    8.
    Michaels AF, Silver MW. Primary production, sinking fluxes and the microbial food web. Deep Sea Res Part Oceanogr Res Pap. 1988;35:473–90.
    Article  Google Scholar 

    9.
    Calbet A, Landry MR. Phytoplankton growth, microzooplankton grazing, and carbon cycling in marine systems. Limnol Oceanogr. 2004;49:51–57.
    CAS  Article  Google Scholar 

    10.
    Jin X, Gruber N, Dunne JP, Sarmiento JL, Armstrong RA. Diagnosing the contribution of phytoplankton functional groups to the production and export of particulate organic carbon, CaCO3, and opal from global nutrient and alkalinity distributions. Glob Biogeochem Cycles. 2006;20:GB2015.
    Article  CAS  Google Scholar 

    11.
    Richardson TL, Jackson GA. Small phytoplankton and carbon export from the surface ocean. Science. 2007;315:838–40.
    CAS  PubMed  Article  Google Scholar 

    12.
    Uitz J, Claustre H, Griffiths FB, Ras J, Garcia N, Sandroni V. A phytoplankton class-specific primary production model applied to the Kerguelen Islands region (Southern Ocean). Deep Sea Res Part Oceanogr Res Pap. 2009;56:541–60.
    CAS  Article  Google Scholar 

    13.
    Uitz J, Claustre H, Gentili B, Stramski D. Phytoplankton class-specific primary production in the world’s oceans: seasonal and interannual variability from satellite observations. Glob Biogeochem Cycles. 2010;24:GB3016.

    14.
    Poulin FJ, Franks PJS. Size-structured planktonic ecosystems: constraints, controls and assembly instructions. J Plankton Res. 2010;32:1121–30.
    PubMed  PubMed Central  Article  Google Scholar 

    15.
    Ward BA, Dutkiewicz S, Jahn O, Follows MJ. A size-structured food-web model for the global ocean. Limnol Oceanogr. 2012;57:1877–91.
    Article  Google Scholar 

    16.
    Stoecker DK, Hansen PJ, Caron DA, Mitra A. Mixotrophy in the marine plankton. Annu Rev Mar Sci. 2017;9:311–35.
    Article  Google Scholar 

    17.
    Tréguer P, Bowler C, Moriceau B, Dutkiewicz S, Gehlen M, Aumont O, et al. Influence of diatom diversity on the ocean biological carbon pump. Nat Geosci. 2018;11:27.
    Article  CAS  Google Scholar 

    18.
    Lancelot C, Hannon E, Becquevort S, Veth C, De, Baar HJW. Modeling phytoplankton blooms and carbon export production in the Southern Ocean: dominant controls by light and iron in the Atlantic sector in Austral spring 1992. Deep Sea Res Part Oceanogr Res Pap. 2000;47:1621–62.
    CAS  Article  Google Scholar 

    19.
    Wang S, Moore JK. Incorporating Phaeocystis into a Southern Ocean ecosystem model. J Geophys Res Oceans 2011;116:C01019.

    20.
    Worthen DL, Arrigo KR. A coupled ocean-ecosystem model of the Ross Sea. Part 1: Interannual variability of primary production and phytoplankton community structure. In: DiTullio GR, Dunbar RB, editors. Biogeochemistry of the Ross Sea. Antarct Res Ser. 2003;78:93–105.

    21.
    Li WKW. Primary production of prochlorophytes, cyanobacteria, and eucaryotic ultraphytoplankton: measurements from flow cytometric sorting. Limnol Oceanogr. 1994;39:169–75.
    CAS  Article  Google Scholar 

    22.
    Jardillier L, Zubkov MV, Pearman J, Scanlan DJ. Significant CO2 fixation by small prymnesiophytes in the subtropical and tropical northeast Atlantic Ocean. ISME J. 2010;4:1180–92.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    23.
    Rii YM, Duhamel S, Bidigare RR, Karl DM, Repeta DJ, Church MJ. Diversity and productivity of photosynthetic picoeukaryotes in biogeochemically distinct regions of the South East Pacific Ocean: Picophytoplankton diversity and productivity in the S. Pacific. Limnol Oceanogr. 2016;61:806–24.
    Article  Google Scholar 

    24.
    Musat N, Halm H, Winterholler B, Hoppe P, Peduzzi S, Hillion F, et al. A single-cell view on the ecophysiology of anaerobic phototrophic bacteria. Proc Natl Acad Sci. 2008;105:17861–6.
    CAS  PubMed  Article  Google Scholar 

    25.
    Foster RA, Kuypers MMM, Vagner T, Paerl RW, Musat N, Zehr JP. Nitrogen fixation and transfer in open ocean diatom–cyanobacterial symbioses. ISME J. 2011;5:1484.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    26.
    Olofsson M, Robertson EK, Edler L, Arneborg L, Whitehouse MJ, Ploug H. Nitrate and ammonium fluxes to diatoms and dinoflagellates at a single cell level in mixed field communities in the sea. Sci Rep. 2019;9:1424.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    27.
    Berthelot H, Duhamel S, L’Helguen S, Maguer J-F, Wang S, Cetinić I, et al. NanoSIMS single cell analyses reveal the contrasting nitrogen sources for small phytoplankton. ISME J. 2019;13:651–62.
    CAS  PubMed  Article  Google Scholar 

    28.
    Ploug H, Musat N, Adam B, Moraru CL, Lavik G, Vagner T, et al. Carbon and nitrogen fluxes associated with the cyanobacterium Aphanizomenon sp. in the Baltic Sea. ISME J. 2010;4:1215–23.
    CAS  PubMed  Article  Google Scholar 

    29.
    Olofsson M, Kourtchenko O, Zetsche E-M, Marchant HK, Whitehouse MJ, Godhe A, et al. High single-cell diversity in carbon and nitrogen assimilations by a chain-forming diatom across a century. Environ Microbiol. 2019;21:142–51.
    CAS  PubMed  Article  Google Scholar 

    30.
    Zaoli S, Giometto A, Marañón E, Escrig S, Meibom A, Ahluwalia A, et al. Generalized size scaling of metabolic rates based on single-cell measurements with freshwater phytoplankton. Proc Natl Acad Sci. 2019;116:17323–9.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    31.
    Frölicher TL, Sarmiento JL, Paynter DJ, Dunne JP, Krasting JP, Winton M. Dominance of the Southern Ocean in anthropogenic carbon and heat uptake in CMIP5 models. J Clim. 2014;28:862–86.
    Article  Google Scholar 

    32.
    Landschützer P, Gruber N, Haumann FA, Rödenbeck C, Bakker DCE, van Heuven S, et al. The reinvigoration of the Southern Ocean carbon sink. Science. 2015;349:1221–4.
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    33.
    Martin JH. Glacial-interglacial CO2 change: the iron hypothesis. Paleoceanography. 1990;5:1–13.
    Article  Google Scholar 

    34.
    de Baar HJW, Jong JTM, de, Bakker DCE, Löscher BM, Veth C, Bathmann U, et al. Importance of iron for plankton blooms and carbon dioxide drawdown in the Southern Ocean. Nature. 1995;373:412.
    Article  Google Scholar 

    35.
    de Baar HJW, Boyd PW, Coale KH, Landry MR, Tsuda A, Assmy P, et al. Synthesis of iron fertilization experiments: from the iron age in the age of enlightenment. J Geophys Res Oceans. 2005;110:C09S16.

    36.
    Weber LH, El-Sayed SZ. Contributions of the net, nano- and picoplankton to the phytoplankton standing crop and primary productivity in the Southern Ocean. J Plankton Res. 1987;9:973–94.
    Article  Google Scholar 

    37.
    Froneman PW, Laubscher RK, Mcquaid CD. Size-fractionated primary production in the South Atlantic and Atlantic Sectors of the Southern Ocean. J Plankton Res. 2001;23:611–22.
    CAS  Article  Google Scholar 

    38.
    Blain S, Quéguiner B, Armand L, Belviso S, Bombled B, Bopp L, et al. Effect of natural iron fertilization on carbon sequestration in the Southern Ocean. Nature. 2007;446:1070–4.
    CAS  PubMed  Article  Google Scholar 

    39.
    Pollard R, Sanders R, Lucas M, Statham P. The Crozet Natural Iron Bloom and Export Experiment (CROZEX). Deep Sea Res Part II Top Stud Oceanogr. 2007;54:1905–14.
    CAS  Article  Google Scholar 

    40.
    Korb RE, Whitehouse MJ, Atkinson A, Thorpe SE. Magnitude and maintenance of the phytoplankton bloom at South Georgia: a naturally iron-replete environment. Mar Ecol Prog Ser. 2008;368:75–91.
    CAS  Article  Google Scholar 

    41.
    Kopczyńska EE, Fiala M, Jeandel C. Annual and interannual variability in phytoplankton at a permanent station off Kerguelen Islands, Southern Ocean. Polar Biol. 1998;20:342–51.
    Article  Google Scholar 

    42.
    Mosseri J, Quéguiner B, Armand L, Cornet-Barthaux V. Impact of iron on silicon utilization by diatoms in the Southern Ocean: a case study of Si/N cycle decoupling in a naturally iron-enriched area. Deep Sea Res Part II Top Stud Oceanogr. 2008;55:801–19.
    Article  Google Scholar 

    43.
    Park Y-H, Roquet F, Durand I, Fuda J-L. Large-scale circulation over and around the Northern Kerguelen Plateau. Deep Sea Res Part II Top Stud Oceanogr. 2008;55:566–81.
    Article  Google Scholar 

    44.
    Holmes RM, Aminot A, Kérouel R, Hooker BA, Peterson BJ. A simple and precise method for measuring ammonium in marine and freshwater ecosystems. Can J Fish Aquat Sci. 1999;56:1801–8.
    CAS  Article  Google Scholar 

    45.
    Aminot A, Kérouel R. Dosage automatique des nutriments dans les eaux marines: méthodes en flux continu. Editions Versailles, France: Quae; 2007.

    46.
    Ras J, Claustre H, Uitz J. Spatial variability of phytoplankton pigment distributions in the Subtropical South Pacific Ocean: comparison between in situ and predicted data. Biogeosciences. 2008;5:353–69.
    CAS  Article  Google Scholar 

    47.
    Mackey MD, Mackey DJ, Higgins HW, Wright SW. CHEMTAX—a program for estimating class abundances from chemical markers: application to HPLC measurements of phytoplankton. Mar Ecol Prog Ser. 1996;144:265–83.
    CAS  Article  Google Scholar 

    48.
    Wright SW, van den Enden RL, Pearce I, Davidson AT, Scott FJ, Westwood KJ. Phytoplankton community structure and stocks in the Southern Ocean (30–80 E) determined by CHEMTAX analysis of HPLC pigment signatures. Deep Sea Res Part II Top Stud Oceanogr. 2010;57:758–78.
    CAS  Article  Google Scholar 

    49.
    van Leeuwe MA, Visser RJW, Stefels J. The pigment composition of Phaeocystis antarctica (Haptophyceae) under various conditions of light, temperature, salinity, and iron. J Phycol. 2014;50:1070–80.
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    50.
    Szymczak-Żyla M, Kowalewska G, Louda JW. The influence of microorganisms on chlorophyll a degradation in the marine environment. Limnol Oceanogr. 2008;53:851–62.
    Article  Google Scholar 

    51.
    Strom SL. Production of pheopigments by marine protozoa: results of laboratory experiments analysed by HPLC. Deep Sea Res Part Oceanogr Res Pap. 1993;40:57–80.
    CAS  Article  Google Scholar 

    52.
    Roca-Martí M, Puigcorbé V, Iversen MH, van der Loeff MR, Klaas C, Cheah W, et al. High particulate organic carbon export during the decline of a vast diatom bloom in the Atlantic sector of the Southern Ocean. Deep Sea Res Part II Top Stud Oceanogr. 2017;138:102–15.
    Article  CAS  Google Scholar 

    53.
    Bower SM, Carnegie RB, Goh B, Jones SR, Lowe GJ, Mak MW. Preferential PCR amplification of parasitic protistan small subunit rDNA from metazoan tissues. J Eukaryot Microbiol. 2004;51:325–32.
    CAS  PubMed  Article  Google Scholar 

    54.
    Guillou L, Bachar D, Audic S, Bass D, Berney C, Bittner L, et al. The Protist Ribosomal Reference database (PR2): a catalog of unicellular eukaryote Small Sub-Unit rRNA sequences with curated taxonomy. Nucleic Acids Res. 2013;41:D597–D604.
    CAS  PubMed  Article  Google Scholar 

    55.
    Callahan BJ, McMurdie PJ, Rosen MJ, Han AW, Johnson AJA, Holmes SP. DADA2: High-resolution sample inference from Illumina amplicon data. Nat Methods. 2016;13:581–3.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    56.
    Andersen KS, Kirkegaard RH, Karst SM, Albertsen M. ampvis2: an R package to analyse and visualise 16S rRNA amplicon data. https://doi.org/https://www.biorxiv.org/content/10.1101/299537v1.article-info. 2018;299537.

    57.
    Not F, Simon N, Biegala IC, Vaulot D. Application of fluorescent in situ hybridization coupled with tyramide signal amplification (FISH TSA) to assess eukaryotic picoplankton composition. Aquat Micro Ecol. 2002;28:157–66.
    Article  Google Scholar 

    58.
    Sun J, Liu D. Geometric models for calculating cell biovolume and surface area for phytoplankton. J Plankton Res. 2003;25:1331–46.
    Article  Google Scholar 

    59.
    Verity PG, Robertson CY, Tronzo CR, Andrews MG, Nelson JR, Sieracki ME. Relationships between cell volume and the carbon and nitrogen content of marine photosynthetic nanoplankton. Limnol Oceanogr. 1992;37:1434–46.
    CAS  Article  Google Scholar 

    60.
    R Core Team. R: a language and environment for statistical computing. Vienna, Austria: R Foundation for Statistical Computing; 2018.

    61.
    Lafond A, Leblanc K, Legras J, Cornet V, Quéguiner B. The structure of diatom communities constrains biogeochemical properties in surface waters of the Southern Ocean (Kerguelen Plateau). J Mar Syst. 2020;212:103458.
    Article  Google Scholar 

    62.
    Chisholm SW. Phytoplankton size. In: Falkowski PG, Woodhead AD, Vivirito K, editors. Primary productivity and biogeochemical cycles in the sea. US, Boston, MA: Springer; 1992. p. 213–37.

    63.
    Marchetti A, Cassar N. Diatom elemental and morphological changes in response to iron limitation: a brief review with potential paleoceanographic applications. Geobiology. 2009;7:419–31.
    CAS  PubMed  Article  Google Scholar 

    64.
    Alderkamp A-C, Kulk G, Buma AGJ, Visser RJW, Van Dijken GL, Mills MM, et al. The effect of iron limitation on the photophysiology of Phaeocystis antarctica (prymnesiophyceae) and Fragilariopsis cylindrus (bacillariophyceae) under dynamic irradiance. J Phycol. 2012;48:45–59.
    CAS  PubMed  Article  Google Scholar 

    65.
    Marañón E. Inter-specific scaling of phytoplankton production and cell size in the field. J Plankton Res. 2008;30:157–63.
    Article  CAS  Google Scholar 

    66.
    Huete-Ortega M, Cermeño P, Calvo-Díaz A, Marañón E. Isometric size-scaling of metabolic rate and the size abundance distribution of phytoplankton. Proc R Soc B Biol Sci. 2012;279:1815–23.
    Article  Google Scholar 

    67.
    Hogle SL, Dupont CL, Hopkinson BM, King AL, Buck KN, Roe KL, et al. Pervasive iron limitation at subsurface chlorophyll maxima of the California Current. Proc Natl Acad Sci. 2018;115:13300–5.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    68.
    Marchetti A, Parker MS, Moccia LP, Lin EO, Arrieta AL, Ribalet F, et al. Ferritin is used for iron storage in bloom-forming marine pennate diatoms. Nature. 2009;457:467–70.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    69.
    Lampe RH, Mann EL, Cohen NR, Till CP, Thamatrakoln K, Brzezinski MA, et al. Different iron storage strategies among bloom-forming diatoms. Proc Natl Acad Sci. 2018;115:E12275–E12284.
    CAS  PubMed  Article  Google Scholar 

    70.
    Litchman E, Klausmeier CA. Trait-based community ecology of phytoplankton. Annu Rev Ecol Evol Syst. 2008;39:615–39.
    Article  Google Scholar 

    71.
    Luxem KE, Ellwood MJ, Strzepek RF. Intraspecific variability in Phaeocystis antarctica’s response to iron and light stress. PLOS One. 2017;12:e0179751.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    72.
    Closset I, Lasbleiz M, Leblanc K, Quéguiner B, Cavagna A-J, Elskens M, et al. Seasonal evolution of net and regenerated silica production around a natural Fe-fertilized area in the Southern Ocean estimated with Si isotopic approaches. Biogeosciences. 2014;11:5827–46.
    Article  Google Scholar 

    73.
    Egge J, Aksnes D. Silicate as regulating nutrient in phytoplankton competition. Mar Ecol Prog Ser. 1992;83:281–9.
    CAS  Article  Google Scholar 

    74.
    Lam PJ, Bishop JKB. High biomass, low export regimes in the Southern Ocean. Deep Sea Res Part II Top Stud Oceanogr. 2007;54:601–38.
    Article  Google Scholar 

    75.
    Maiti K, Charette MA, Buesseler KO, Kahru M. An inverse relationship between production and export efficiency in the Southern Ocean. Geophys Res Lett. 2013;40:1557–61.
    Article  Google Scholar 

    76.
    Le Moigne FAC, Henson SA, Cavan E, Georges C, Pabortsava K, Achterberg EP, et al. What causes the inverse relationship between primary production and export efficiency in the Southern Ocean? Geophys Res Lett. 2016;43:4457–66.
    Article  CAS  Google Scholar 

    77.
    Christaki U, Lefèvre D, Georges C, Colombet J, Catala P, Courties C, et al. Microbial food web dynamics during spring phytoplankton blooms in the naturally iron-fertilized Kerguelen area (Southern Ocean). Biogeosciences. 2014;11:6739–53.
    Article  Google Scholar 

    78.
    Christaki U, Gueneugues A, Liu Y, Blain S, Catala P, Colombet J, et al. Seasonal microbial food web dynamics in contrasting Southern Ocean productivity regimes. Limnol Oceanogr. 2021;66:108–22.

    79.
    Planchon F, Ballas D, Cavagna A-J, Bowie AR, Davies D, Trull T, et al. Carbon export in the naturally iron-fertilized Kerguelen area of the Southern Ocean based on the 234Th approach. Biogeosciences. 2015;12:3831–48.
    Article  Google Scholar 

    80.
    Cassar N, Wright SW, Thomson PG, Trull TW, Westwood KJ, Salas Mde, et al. The relation of mixed-layer net community production to phytoplankton community composition in the Southern Ocean. Glob Biogeochem Cycles. 2015;29:446–62.
    CAS  Article  Google Scholar 

    81.
    Sassenhagen I, Irion S, Jardillier L, Moreira D, Christaki U. Protist interactions and community structure during early autumn in the Kerguelen Region (Southern Ocean). Protist. 2020;171:125709.
    CAS  PubMed  Article  Google Scholar 

    82.
    Henschke N, Blain S, Cherel Y, Cotte C, Espinasse B, Hunt BPV, et al. Population demographics and growth rate of Salpa thompsoni on the Kerguelen Plateau. J Marine Systems. 2021;214:103489.
    Article  Google Scholar 

    83.
    Moline MA, Claustre H, Frazer TK, Schofield O, Vernet M. Alteration of the food web along the Antarctic Peninsula in response to a regional warming trend. Glob Change Biol. 2004;10:1973–80.
    Article  Google Scholar 

    84.
    Iversen MH, Pakhomov EA, Hunt BPV, van der Jagt H, Wolf-Gladrow D, Klaas C. Sinkers or floaters? Contribution from salp pellets to the export flux during a large bloom event in the Southern Ocean. Deep Sea Res Part II Top Stud Oceanogr. 2017;138:116–25.
    CAS  Article  Google Scholar 

    85.
    Irion S, Jardillier L, Sassenhagen I, Christaki U. Marked spatiotemporal variations in small phytoplankton structure in contrasted waters of the Southern Ocean (Kerguelen area). Limnol Oceanogr. 2020;65:2835–52.

    86.
    Le Moigne FAC, Poulton AJ, Henson SA, Daniels CJ, Fragoso GM, Mitchell E, et al. Carbon export efficiency and phytoplankton community composition in the Atlantic sector of the Arctic Ocean. J Geophys Res Oceans. 2015;120:3896–912.
    Article  Google Scholar 

    87.
    DiTullio GR, Grebmeier JM, Arrigo KR, Lizotte MP, Robinson DH, Leventer A, et al. Rapid and early export of Phaeocystis antarctica blooms in the Ross Sea, Antarctica. Nature. 2000;404:595–8.
    CAS  PubMed  Article  Google Scholar 

    88.
    Pauthenet E, Roquet F, Madec G, Guinet C, Hindell M, McMahon CR, et al. Seasonal meandering of the Polar Front upstream of the Kerguelen Plateau. Geophys Res Lett. 2018;45:9774–81.
    Article  Google Scholar  More

  • in

    First description of a widespread Mytilus trossulus-derived bivalve transmissible cancer lineage in M. trossulus itself

    1.
    Murchison, E. P. Clonally transmissible cancers in dogs and Tasmanian devils. Oncogene 27, S19–S30 (2008).
    CAS  PubMed  Article  PubMed Central  Google Scholar 
    2.
    Metzger, M. J. et al. Widespread transmission of independent cancer lineages within multiple bivalve species. Nature 534, 705–709 (2016).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    3.
    Storfer, A. et al. The devil is in the details: Genomics of transmissible cancers in Tasmanian devils. PLoS Pathog. 14, 1–7 (2018).
    Article  CAS  Google Scholar 

    4.
    Nowinsky, M. A. Zur Frage über die Impfung der krebsigen Geschwülste. Zentralbl Med Wissensch. 14, 790–791 (1876).
    Google Scholar 

    5.
    Murgia, C., Pritchard, J. K., Kim, S. Y., Fassati, A. & Weiss, R. A. Clonal origin and evolution of a transmissible cancer. Cell 126, 477–487 (2006).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    6.
    Pearse, A. M. & Swift, K. Allograft theory: transmission of devil facial-tumour disease. Nature 439, 549 (2006).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    7.
    Siddle, H. V. et al. Transmission of a fatal clonal tumor by biting occurs due to depleted MHC diversity in a threatened carnivorous marsupial. Proc. Natl. Acad. Sci. USA 104, 16221–16226 (2007).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    8.
    Metzger, M. J. J., Reinisch, C., Sherry, J. & Goff, S. P. P. Horizontal transmission of clonal cancer cells causes leukemia in soft-shell clams. Cell 161, 255–263 (2015).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    9.
    Riquet, F., Simon, A. & Bierne, N. Weird genotypes? Don’t discard them, transmissible cancer could be an explanation. Evol. Appl. 10, 140–145 (2017).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    10.
    Yonemitsu, M. A. et al. A single clonal lineage of transmissible cancer identified in two marine mussel species in South America and Europe. Elife 8, 1–21 (2019).
    Article  Google Scholar 

    11.
    Woods, G. M., Bruce Lyons, A. & Bettiol, S. S. A devil of a transmissible cancer. Trop. Med. Infect. Dis. 5, 1–10 (2020).
    Google Scholar 

    12.
    Ganguly, B., Das, U. & Das, A. K. Canine transmissible venereal tumour: a review. Vet. Comp. Oncol. 14, 1–12 (2016).
    CAS  PubMed  Article  Google Scholar 

    13.
    Paynter, A. N., Metzger, M. J., Sessa, J. A. & Siddall, M. E. Evidence of horizontal transmission of the cancer-associated Steamer retrotransposon among ecological cohort bivalve species. Dis. Aquat. Organ. https://doi.org/10.3354/dao03113 (2017).
    Article  PubMed  Google Scholar 

    14.
    Barber, B. J. Neoplastic diseases of commercially important marine bivalves. Aquat. Living Resour. 17, 449–466 (2004).
    Article  Google Scholar 

    15.
    Farley, C. A. Sarcomatoid proliferative disease in a wild population of blue mussels (Mytilus edulis). J. Natl. Cancer Inst. 43, 509–516 (1969).
    CAS  PubMed  PubMed Central  Google Scholar 

    16.
    Sparks, A. K. Synopsis of Invertebrate Pathology Exclusive of Insects (Elsevier, 1985).
    Google Scholar 

    17.
    Muttray, A. F. & Vassilenko, K. Mollusca: disseminated neoplasia in bivalves and the p53 protein family. Adv. Comp. Immunol. https://doi.org/10.1007/978-3-319-76768-0 (1969).
    Article  Google Scholar 

    18.
    Bihari, N., Mičić, M., Batel, R. & Zahn, R. K. Flow cytometric detection of DNA cell cycle alterations in hemocytes of mussels (Mytilus galloprovincialis) off the Adriatic coast Croatia. Aquat. Toxicol. 64, 121–129 (2003).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    19.
    Vassilenko, E. & Baldwin, S. A. Using flow cytometry to detect haemic neoplasia in mussels (Mytilus trossulus) from the Pacific Coast of Southern British Columbia Canada. J. Invertebr. Pathol. 117, 68–72 (2014).
    PubMed  Article  PubMed Central  Google Scholar 

    20.
    Moore, J. D., Elston, R. A., Drum, A. S. & Wilkinson, M. T. Alternate pathogenesis of systemic neoplasia in the bivalve mollusc Mytilus. J. Invertebr. Pathol. 58, 231–243 (1991).
    CAS  PubMed  Article  Google Scholar 

    21.
    Benabdelmouna, A., Saunier, A., Ledu, C., Travers, M. A. & Morga, B. Genomic abnormalities affecting mussels (Mytilus edulis-galloprovincialis) in France are related to ongoing neoplastic processes, evidenced by dual flow cytometry and cell monolayer analyses. J. Invertebr. Pathol. 157, 45–52 (2018).
    Article  CAS  Google Scholar 

    22.
    Burioli, E. A. V. et al. Implementation of various approaches to study the prevalence, incidence and progression of disseminated neoplasia in mussel stocks. J. Invertebr. Pathol. 168, 107271 (2019).
    CAS  PubMed  Article  Google Scholar 

    23.
    Moore, C. A., Beckmann, N. & Patricia Morse, M. Cytoskeletal structure of diseased and normal hemocytes of Mya arenaria. J. Invertebr. Pathol. 60, 141–147 (1992).
    Article  Google Scholar 

    24.
    Carballal, M. J., Barber, B. J., Iglesias, D. & Villalba, A. Neoplastic diseases of marine bivalves. J. Invertebr. Pathol. https://doi.org/10.1016/j.jip.2015.06.004 (2015).
    Article  PubMed  Google Scholar 

    25.
    Odintsova, N. A. Leukemia-like cancer in bivalves. Russ. J. Mar. Biol. 46, 59–67 (2020).
    Article  Google Scholar 

    26.
    Smolarz, K., Renault, T. & Wołowicz, M. Ultrastructural study of neoplastic cells in Macoma balthica (Bivalvia) from the Gulf of Gdansk (Poland). J. Invertebr. Pathol. 92, 79–84 (2006).
    PubMed  Article  Google Scholar 

    27.
    Carella, F. et al. Nuclear morphometry and ploidy of normal and neoplastic haemocytes in mussels. PLoS ONE 12, e0173219 (2017).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    28.
    Smolowitz, R. M. & Reinisch, C. L. Indirect peroxidase staining using monoclonal antibodies specific for Mya arenaria neoplastic cells. J. Invertebr. Pathol. 48, 139–145 (1986).
    CAS  PubMed  Article  Google Scholar 

    29.
    Carella, F., Figueras, A., Novoa, B. & De Vico, G. Cytomorphology and PCNA expression pattern in bivalves Mytilus galloprovincialis and Cerastoderma edule with haemic neoplasia. Dis. Aquat. Organ. 105, 81–87 (2013).
    PubMed  Article  Google Scholar 

    30.
    McDonald, J. H., Seed, R. & Koehn, R. K. Allozymes and morphometric characters of three species of Mytilus in the Northern and Southern Hemispheres. Mar. Biol. 111, 323–333 (1991).
    Article  Google Scholar 

    31.
    Larraín, M. A., Zbawicka, M., Araneda, C., Gardner, J. P. A. & Wenne, R. Native and invasive taxa on the Pacific coast of South America: impacts on aquaculture, traceability and biodiversity of blue mussels (Mytilus spp.). Evol. Appl. 11, 298–311 (2018).
    Article  CAS  Google Scholar 

    32.
    Pye, R. J. et al. A second transmissible cancer in Tasmanian devils. Proc. Natl. Acad. Sci. USA 113, 374–379 (2016).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    33.
    Strakova, A. et al. Mitochondrial genetic diversity, selection and recombination in a canine transmissible cancer. Elife 5, 1–25 (2016).
    Article  Google Scholar 

    34.
    Rawson, P. D. & Harper, F. M. Colonization of the northwest Atlantic by the blue mussel, Mytilus trossulus postdates the last glacial maximum. Mar. Biol. 156, 1857–1868 (2009).
    Article  Google Scholar 

    35.
    Wenne, R., Bach, L., Zbawicka, M., Strand, J. & McDonald, J. H. A first report on coexistence and hybridization of Mytilus trossulus and M. edulis mussels in Greenland. Polar Biol. 39, 343–355 (2016).
    Article  Google Scholar 

    36.
    Laakkonen, H. M., Hardman, M., Strelkov, P. & Väinölä, R. Cycles of trans-Arctic dispersal and vicariance, and diversification of the amphi-boreal marine fauna. J. Evol. Biol. https://doi.org/10.1111/jeb.13674 (2020).
    Article  PubMed  PubMed Central  Google Scholar 

    37.
    Zbawicka, M., Wenne, R. & Burzyński, A. Mitogenomics of recombinant mitochondrial genomes of Baltic Sea Mytilus mussels. Mol. Genet. Genom. 289, 1275–1287 (2014).
    CAS  Article  Google Scholar 

    38.
    Śmietanka, B. & Burzyński, A. Disruption of doubly uniparental inheritance of mitochondrial DNA associated with hybridization area of European Mytilus edulis and Mytilus trossulus in Norway. Mar. Biol. https://doi.org/10.1007/s00227-017-3235-5 (2017).
    Article  PubMed  PubMed Central  Google Scholar 

    39.
    Sunila, I. Histopathology of mussels (Mytilus edulis L.) from the Tvarminne area, the Gulf of Finland (Baltic Sea). Ann. Zool. Fennici 24, 55–69 (1987).
    Google Scholar 

    40.
    Usheva, L. N. & Frolova, L. T. Neoplasia in the connective tissue tumor in the mussel Mytilus trossulus from a polluted region of Nakhodka Bay, the Sea of Japan. Ontogenez 31, 63–70 (2000).
    CAS  PubMed  PubMed Central  Google Scholar 

    41.
    Maiorova, M. A. & Odintsova, N. A. β integrin-like protein-mediated adhesion and its disturbances during cell cultivation of the mussel Mytilus trossulus. Cell Tissue Res. 361, 581–592 (2015).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    42.
    Marigomez, I., Lekube, X. & Cancio, I. Immunochemical localisation of proliferating cells in mussel digestive gland tissue. Histochem. J. 31, 781–788 (1999).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    43.
    Maiorova, M. A. & Odintsova, N. A. Proliferative potential of larval cells of the mussel Mytilus trossulus and their capacity to differentiate into myogenic cells in culture. Russ. J. Mar. Biol. 42, 281–285 (2016).
    Article  Google Scholar 

    44.
    Voronezhskaya, E. E., Nezlin, L. P., Odintsova, N. A., Plummer, J. T. & Croll, R. P. Neuronal development in larval mussel Mytilus trossulus (Mollusca: Bivalvia). Zoomorphology 127, 97–110 (2008).
    Article  Google Scholar 

    45.
    Presa, P., Pérez, M., Diz, A. P., Perez, M. & Diz, A. P. Polymorphic microsatellite markers for blue mussels (Mytilus spp.). Conserv. Genet. 3, 441–443 (2002).
    CAS  Article  Google Scholar 

    46.
    Zouros, E. Biparental inheritance through uniparental transmission: the doubly uniparental inheritance (DUI) of mitochondrial DNA. Evol. Biol. 40, 1–31 (2013).
    Article  Google Scholar 

    47.
    Kumar, S., Stecher, G., Li, M., Knyaz, C. & Tamura, K. MEGA X: molecular evolutionary genetics analysis across computing platforms. Mol. Biol. Evol. 35, 1547–1549 (2018).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    48.
    Paabo, S., Irwin, M. & Wilson, C. DNA damage promotes jumping between templates during enzymatic amplification. J. Biol. Chem. 265, 4718–4721 (1990).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    49.
    Bradley, R. D. & Hillis, D. A. Recombinant DNA sequences generated by PCR amplification. Mol. Biol. Evol. 14, 592–593 (1997).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    50.
    Letunic, I. & Bork, P. Interactive Tree of Life (iTOL) v4: recent updates and new developments. Nucl. Acids Res. 47, 256–259 (2019).
    Article  CAS  Google Scholar 

    51.
    Martin, D. P., Murrell, B., Golden, M., Khoosal, A. & Muhire, B. RDP4: detection and analysis of recombination patterns in virus genomes. Virus Evol. 1, 1–5 (2015).
    Article  Google Scholar 

    52.
    Breton, S., Burger, G., Stewart, D. T. & Blier, P. U. Comparative analysis of gender-associated complete mitochondrial genomes in marine mussels (Mytilus spp.). Genetics 172, 1107–1119 (2006).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    53.
    Marko, P. B. et al. The ‘Expansion-contraction’ model of Pleistocene biogeography: rocky shores suffer a sea change?. Mol. Ecol. 19, 146–169 (2010).
    CAS  PubMed  Article  Google Scholar 

    54.
    Crego-Prieto, V., Juanes, F. & Garcia-Vazquez, E. Aquaculture and the spread of introduced mussel genes in British Columbia. Biol. Invasions https://doi.org/10.1007/s10530-015-0853-z (2015).
    Article  Google Scholar 

    55.
    Chung, J. M. et al. Molecular phylogenetic study of bivalvia from four countries (China, Japan, Russia and Myanmar) using 3 types of primers. Korean J. Malacol. 35, 137–148 (2019).
    Google Scholar 

    56.
    Layton, K. K. S., Martel, A. L., Hebert, P. D. N. & Layton, K. K. S. Patterns of DNA barcode variation in Canadian Marine molluscs. PLoS ONE 9, 1–9 (2014).
    Article  Google Scholar 

    57.
    deWaard, J. R. et al. A reference library for Canadian invertebrates with 1.5 million barcodes, voucher specimens, and DNA samples. Sci. Data 6, 1–12 (2019).
    Article  CAS  Google Scholar 

    58.
    Clement, M., Snell, Q., Walker, P., Posada, D. & Crandall, K. TCS: Estimating Gene Genealogies. Parallel Distrib. Process. Symp. Int. 2, 184 (2002).

    59.
    Leigh, J. W. & Bryant, D. POPART: full-feature software for haplotype network construction. Methods Ecol. Evol. 6, 1110–1116 (2015).
    Article  Google Scholar 

    60.
    Kartavtsev, Y. P., Masalkova, N. A. & Katolikova, M. V. Genetic and morphometric variability in settlements of two mussel species (Mytilus ex. Gr. Edulis), Mytilus trossulus and Mytilus galloprovincialis, in the Northwestern Sea of Japan. J. Shellfish Res. 37, 103–119 (2018).
    Article  Google Scholar 

    61.
    Larraín, M. A., González, P., Pérez, C. & Araneda, C. Comparison between single and multi-locus approaches for specimen identification in Mytilus mussels. Sci. Rep. 9, 1–13 (2019).
    Article  CAS  Google Scholar 

    62.
    Inoue, K., Waite, J. H., Matsuoka, M., Odo, S. & Harayama, S. Interspecific variations in adhesive protein sequences of Mytilus edulis, M. galloprovincialis, and M. trossulus. Biol. Bull. 189, 370–375 (1995).
    CAS  PubMed  Article  Google Scholar 

    63.
    Heath, D. D., Rawson, P. D. & Hilbish, T. J. PCR-based nuclear markers identify alien blue mussel Mytilus spp. genotypes on the west coast of Canada. Can. J. Fish. Aquat. Sci. 52, 2621–2627 (1995).
    CAS  Article  Google Scholar 

    64.
    Rawson, P. D., Secor, C. L. & Hilbish, T. J. The effects of natural hybridization on the regulation of doubly uniparental mtDNA inheritance in blue mussels (Mytilus spp.). Genetics 144, 241–248 (1996).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    65.
    Rawson, P. D. & Hilbish, T. J. Distribution of male and female mtDNA lineages in populations of blue mussels, Mytilus trossulus and M. galloprovincialis, along the Pacific coast of North America. Mar. Biol. 124, 245–250 (1995).
    Article  Google Scholar 

    66.
    Cremonte, F., Vázquez, N. & Silva, M. R. Gonad atrophy caused by disseminated neoplasia in Mytilus chilensis cultured in the Beagle channel, Tierra Del Fuego Province Argentina. J. Shellfish Res. 30, 845–849 (2011).
    Article  Google Scholar 

    67.
    Ciocan, C. & Sunila, I. Disseminated neoplasia in blue mussels, Mytilus galloprovincialis, from the Black Sea Romania. Mar. Pollut. Bull. 50, 1335–1339 (2005).
    CAS  PubMed  Article  Google Scholar 

    68.
    Gombač, M., Sitar, R., Pogačnik, M., Arzul, I. & Jenčič, V. Haemocytic neoplasia in Mediterranean mussels (Mytilus galloprovincialis). Mar. Freshw. Behav. Physiol. 46, 135–143 (2013).
    Article  Google Scholar 

    69.
    Villalba, A., Mourelle, S. G., Carballal, M. J. & López, C. Symbionts and diseases of farmed mussels Mytilus galloprovincialis throughout the culture process in the Rias of Galicia (NW Spain). Dis. Aquat. Organ. 31, 127–139 (1997).
    Article  Google Scholar 

    70.
    Cremonte, F., Puebla, C., Tillería, J. & Videla, V. Estudio histopatológico del chorito Mytilus chilensis (Mytilidae) y del culengue Gari solida (Psammobiidae) en el sur de Chile. Lat. Am. J. Aquat. Res. 43, 248–254 (2015).
    Article  Google Scholar 

    71.
    Vassilenko, E. I., Muttray, A. F., Schulte, P. M. & Baldwin, S. A. Variations in p53-like cDNA sequence are correlated with mussel haemic neoplasia: a potential molecular-level tool for biomonitoring. Mutat. Res. Genet. Toxicol. Environ. Mutagen. https://doi.org/10.1016/j.mrgentox.2010.06.001 (2010).
    Article  Google Scholar 

    72.
    Caza, F., Bernet, E., Veyrier, F. J., Betoulle, S. & St-Pierre, Y. Hemocytes released in seawater act as Trojan horses for spreading of bacterial infections in mussels. Sci. Rep. 10, 1–12 (2020).
    Article  CAS  Google Scholar 

    73.
    Rebbeck, C. A., Leroi, A. M. & Burt, A. Mitochondrial capture by a transmissible cancer. Science (80-) 331, 303 (2011).
    ADS  CAS  Article  Google Scholar 

    74.
    Rawson, P. D. & Hilbish, T. J. Asymmetric introgression of mitochondrial DNA among European populations of Blue Mussels (Mytilus spp.). Evolution (N. Y.) 52, 100–108 (1998).
    Google Scholar 

    75.
    Kijewski, A., Zbawicka, M., Väinölä, R. & Wenne, T. K. Introgression and mitochondrial DNA heteroplasmy in the Baltic populations of mussels Mytilus trossulus and M. edulis. Mar. Biol. https://doi.org/10.1007/s00227-006-0316-2 (2006).
    Article  Google Scholar 

    76.
    Burzyński, A., Zbawicka, M., Skibinski, D. O. F. & Wenne, R. Evidence for recombination of mtDNA in the marine mussel Mytilus trossulus from the Baltic. Mol. Biol. Evol. 20, 388–392 (2003).
    PubMed  Article  PubMed Central  Google Scholar 

    77.
    Burzynski, A., Zbawicka, M., Skibinski, D. O. F. & Wenne, R. Doubly uniparental inheritance is associated with high polymorphism for rearranged and recombinant control region haplotypes in Baltic Mytilus trossulus. Genetics 174, 1081–1094 (2006).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    78.
    Stuckas, H., Stoof, K., Quesada, H. & Tiedemann, R. Evolutionary implications of discordant clines across the Baltic Mytilus hybrid zone (Mytilus edulis and Mytilus trossulus). Heredity (Edinb). 103, 146–156 (2009).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    79.
    Cao, L., Kenchington, E. & Zouros, E. Differential segregation patterns of sperm mitochondria in embryos of the Blue Mussel (Mytilus edulis). Genetics https://doi.org/10.1534/genetics.166.2.883 (2004).
    Article  PubMed  PubMed Central  Google Scholar  More

  • in

    Key soil parameters affecting the survival of Panax notoginseng under continuous cropping

    1.
    Zhao, Y. Panax notoginseng (Burk.) F.H. Chen (Sanqi, Notoginseng) (eds. Liu, Y., Wang, Z. & Zhang, J.) 185–193 (Springer, 2015).
    2.
    Liao, P. et al. Stereoscopic cultivation of Panax notoginseng: A new approach to overcome the continuous cropping obstacle. Ind. Crop Prod. 126, 38–47 (2018).
    CAS  Article  Google Scholar 

    3.
    Yang, M. et al. Autotoxic ginsenosides in the rhizosphere contribute to the replant failure of Panax notoginseng. PLoS ONE 10, e0118555 (2015).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    4.
    Mazzola, M. & Manici, L. M. Apple replant disease: Role of microbial ecology in cause and control. Annu. Rev. Phytopathol. 50, 45–65 (2012).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    5.
    Lovaisa, N. C. et al. Strawberry monocropping: Impacts on fruit yield and soil microorganisms. J. Soil Sci. Plant Nutr. 17, 868–883 (2017).
    CAS  Article  Google Scholar 

    6.
    Petermann, J. S., Fergus, A. J. F., Turnbull, L. A. & Schmid, B. Janzen–Connell effects are widespread and strong enough to maintain diversity in grasslands. Ecology 89, 2399–2406 (2008).
    PubMed  Article  PubMed Central  Google Scholar 

    7.
    Xia, P. G. et al. Optimal fertilizer application for Panax notoginseng and effect of soil water on root rot disease and saponin contents. J. Ginseng Res. 40, 38–46 (2016).
    PubMed  Article  PubMed Central  Google Scholar 

    8.
    Dong, L., Xu, J., Feng, G., Li, X. & Chen, S. Soil bacterial and fungal community dynamics in relation to Panax notoginseng death rate in a continuous cropping system. Sci. Rep. 6, 31802 (2016).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    9.
    Zhang, Y., Zheng, Y., Xia, P., Xun, L. & Liang, Z. Impact of continuous Panax notoginseng plantation on soil microbial and biochemical properties. Sci. Rep. 9, 13205 (2019).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    10.
    Liu, D. H. et al. Study on dynamic change law of N, P and K in Panax notoginseng plant soils with different interval year. China J. Chin. Mater. Med. 39, 572–579 (2014).
    ADS  Google Scholar 

    11.
    Vukicevich, E., Lowery, T., Bowen, P., Úrbez-Torres, J. R. & Hart, M. Cover crops to increase soil microbial diversity and mitigate decline in perennial agriculture. A review. Agron. Sustain. Dev. 36, 48 (2016).
    Article  CAS  Google Scholar 

    12.
    Manici, L. M., Caputo, F. & Saccà, M. L. Secondary metabolites released into the rhizosphere by Fusarium oxysporum and Fusarium spp. as underestimated component of nonspecific replant disease. Plant Soil 415, 85–98 (2016).
    Article  CAS  Google Scholar 

    13.
    Tan, Y. et al. Rhizospheric soil and root endogenous fungal diversity and composition in response to continuous Panax notoginseng cropping practices. Microbiol. Res. 194, 10–19 (2016).
    PubMed  Article  PubMed Central  Google Scholar 

    14.
    Dumbrell, A. J., Nelson, M., Helgason, T., Dytham, C. & Fitter, A. H. Relative roles of niche and neutral processes in structuring a soil microbial community. ISME J. 4, 337–345 (2010).
    PubMed  Article  PubMed Central  Google Scholar 

    15.
    Jeanbille, M. et al. Soil parameters drive the structure, diversity and metabolic potentials of the bacterial communities across temperate beech forest soil sequences. Microb. Ecol. 71, 482–493 (2016).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    16.
    Wu, F. Y., Cui, X. M., Yang, Y. & Guan, H. L. Effects of soil pH values regulated by different fertilization on the disease incidence and growth of Panax notoginseng. J. Yunnan Univ. 39, 908–914 (2017).
    Google Scholar 

    17.
    Wei, W., Yang, M., Liu, Y., Huang, H. & Zhu, S. Fertilizer N application rate impacts plant–soil feedback in a Sanqi production system. Sci. Total Environ. 633, 796–807 (2018).
    ADS  CAS  PubMed  Article  Google Scholar 

    18.
    Killham, K. & Staddon, W. J. Bioindicators and sensors of soil health and the application of geostatistics (eds. Burns, R. G. & Dick, R. P.) 391–405 (Marcel Dekker, 2002).

    19.
    Kotroczó, Z. et al. Soil enzyme activity in response to long-term organic matter manipulation. Soil Biol. Biochem. 70, 237–243 (2014).
    Article  CAS  Google Scholar 

    20.
    Sinsabaugh, R. L., Antibus, R. K. & Linkins, A. E. An enzymic approach to the analysis of microbial activity during plant litter decomposition. Agric. Ecosyst. Environ. 34, 43–54 (1991).
    CAS  Article  Google Scholar 

    21.
    Li, W. H., Liu, Q. Z. & Chen, P. Effect of long-term continuous cropping of strawberry on soil bacterial community structure and diversity. J. Integr. Agric. 17, 2570–2582 (2018).
    Article  Google Scholar 

    22.
    Sun, X. T. et al. Properties of soil physical–chemistry and activities of soil enzymes in context of continuous cropping obstacles for Panax notoginseng. Ecol. Environ. Sci. 24, 409–417 (2015).
    Google Scholar 

    23.
    Sun, X. T. et al. The progress and prospect on consecutive monoculture problems of Panax notoginseng. Chin. J. Ecol. 34, 885–893 (2015).
    Google Scholar 

    24.
    Tyler, G. & Falkengren-Grerup, U. Soil chemistry and plant performance-ecological consideration. Progr. Bot. 59, 634–658 (1998).
    CAS  Article  Google Scholar 

    25.
    Utkhede, R. S. Soil sickness, replant problem or replant disease and its integrated control. Allelopathy J. 18, 23–38 (2006).
    Google Scholar 

    26.
    Hijmans, R. J., Cameron, S. E., Parra, J. L., Jones, P. G. & Jarvis, A. Very high resolution interpolated climate surfaces for global land areas. Int. J. Climatol. 25, 1965–1978 (2005).
    Article  Google Scholar 

    27.
    Lu, R. K. Analytical Method of Soil and Agricultural Chemistry 260–344 (China Agricultural Science and Technology Press, 2000).
    Google Scholar 

    28.
    Guan, S. Soil Enzyme and Its Research Methods 22–36 (Agricultural Publisher, 1986).
    Google Scholar 

    29.
    Zhang, Q. et al. Microcalorimetric study of the effects of long-term fertilization on soil microbial activity in a wheat field on the loess plateau. Ecotoxicology 23, 2035–2040 (2014).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    30.
    Fellows, I. Deducer: A data analysis GUI for R. J. Stat. Softw. 49, 1–15 (2012).
    Article  Google Scholar 

    31.
    Yang, M. et al. Steaming combined with biochar application eliminates negative plant-soil feedback for Sanqi cultivation. Soil Till. Res. 189, 189–198 (2019).
    Article  Google Scholar 

    32.
    Tan, Y. et al. Diversity and composition of rhizospheric soil and root endogenous bacteria in Panax notoginseng during continuous cropping practices. J. Basic Microbiol. 57, 337–344 (2017).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    33.
    Zhao, Y. M. et al. Role of phenolic acids from the rhizosphere soils of Panax notoginseng as a double-edge sword in the occurrence of root-rot disease. Molecules 23, 819 (2018).
    PubMed Central  Article  CAS  Google Scholar 

    34.
    Wu, L. J., Guan, Y. M. & Liu, J. Y. Pollution-Free Cultivation of Panax notoginseng 69–79 (Chemical Industry Press, 2004).
    Google Scholar 

    35.
    Du, C. Y., Zhang, N. M., Jiang, R., Wang, T. & Liu, Y. Evaluation of main soil nutrients characteristics for Panax notoginseng planting area of Yunnan. Southwest China J. Agric. Sci. 29, 599–605 (2016).
    Google Scholar 

    36.
    Zhou, X. et al. Changes in rhizosphere soil microbial communities in a continuously monocropped cucumber (Cucumis sativus L.) system. Eur. J. Soil Biol. 60, 1–8 (2014).
    CAS  Article  Google Scholar 

    37.
    Qin, S. H. et al. Breaking continuous potato cropping with legumes improves soil microbial communities, enzyme activities and tuber yield. PLoS ONE 12, e0175934 (2017).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    38.
    Dou, F., Wright, A. L., Mylavarapu, R. S., Jiang, X. J. & Matocha, J. E. Soil enzyme activities and organic matter composition affected by 26 years of continuous cropping. Pedosphere 26, 618–625 (2016).
    CAS  Article  Google Scholar 

    39.
    Wang, Y. et al. Effects of biochar on the growth of apple seedlings, soil enzyme activities and fungal communities in replant disease soil. Sci. Hortic. 256, 108641 (2019).
    CAS  Article  Google Scholar 

    40.
    Adetunji, A. T., Lewu, F. B., Mulidzi, R. & Ncube, B. The biological activities of β-glucosidase, phosphatase and urease as soil quality indicators: A review. J. Soil Sci. Plant Nutr. 17, 794–807 (2017).
    CAS  Article  Google Scholar 

    41.
    Strickland, M. S. & Rousk, J. Considering fungal:bacterial dominance in soils—Methods, controls, and ecosystem implications. Soil Biol. Biochem. 42, 1385–1395 (2010).
    CAS  Article  Google Scholar 

    42.
    Malik, A. A. et al. Soil fungal:bacterial ratios are linked to altered carbon cycling. Front. Microbiol. 7, 1247 (2016).
    PubMed  PubMed Central  Article  Google Scholar 

    43.
    Liu, X. et al. Soil fumigation and bio-organic fertilizer application promotes potato growth and affects soil bio-chemical properties in a continuous cropping system. Acta Pratacult. Sin. 24, 122–133 (2015).
    Google Scholar 

    44.
    Liu, H. J. et al. Characteristics of soil microflora of Panax notoginseng in different continuous cropping years. Allelopathy J. 44, 145–158 (2018).
    ADS  Article  Google Scholar 

    45.
    Singh, B. P., Cowi, A. L. & Chan, K. Y. Soil Health and Climate Change 69–85 (Springer, 2011).
    Google Scholar 

    46.
    Turner, B. L. Variation in pH optima of hydrolytic enzyme activities in tropical rain forest soils. Appl. Environ. Microbiol. 76, 6485–6493 (2011).
    Article  CAS  Google Scholar 

    47.
    Burns, R. G. et al. Soil enzymes in a changing environment: Current knowledge and future directions. Soil. Biol. Biochem. 58, 216–234 (2013).
    CAS  Article  Google Scholar 

    48.
    Stark, S., Männistö, M. K. & Eskelinen, A. Nutrient availability and pH jointly constrain microbial extracellular enzyme activities in nutrient-poor tundra soils. Plant Soil 383, 373–385 (2014).
    CAS  Article  Google Scholar 

    49.
    Hermans, S. M. et al. Bacteria as emerging indicators of soil condition. Appl. Environ. Microbiol. 83, e02826-e2916 (2017).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    50.
    Dong, L. L. et al. Manipulation of microbial community in the rhizosphere alleviates the replanting issues in Panax ginseng. Soil. Biol. Biochem. 125, 64–74 (2018).
    CAS  Article  Google Scholar 

    51.
    Li, X. et al. Effects of long-term continuous cropping on soil nematode community and soil condition associated with replant problem in strawberry habitat. Sci. Rep. 6, 30466 (2016).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar  More

  • in

    Increases in Great Lake winds and extreme events facilitate interbasin coupling and reduce water quality in Lake Erie

    Climate change has increased water temperature and altered wind-driven water movements in aquatic systems1,2. This applies not only to the mean conditions3,4, but also to the frequency of extreme events (i.e., near the upper ends of the range of observed values5,  > 80th percentile). For example, high air temperature or powerful winds5,6,7,8,9 has affected the behaviour of surface gravity waves10. Understanding the changes in wind and wave climate provides insight into the prediction and management of climate change impacts related to coastal dynamics, such as coastal erosion and sediment budgets, water motions, and biological responses6,11,12. Several studies on the impacts of climate change on oceanic waves12,13,14,15 have been undertaken, including a recent study16 that shows a 0.41% annual increase in global wave power (WP; the transport of energy by waves, which represents the temporal variations of energy transferred from the atmosphere to the ocean surface motion over cumulative periods of time16,17 (Eq. 2) due to stronger winds caused by increases in sea surface temperature. The oceanic wave climate also responds to global atmospheric phenomena (e.g., El Niño Southern Oscillation and the Atlantic Multidecadal Oscillation), in which sea surface temperature modifies wind patterns and storm cyclogenesis12,18,19,20. A systematic long-term assessment of climate warming impacts on waves in lakes remains to be undertaken, but should include winds, which are one of the principal sources of mechanical energy for lake circulation and interbasin coupling (e.g., exchange)21,22,23,24.
    The Laurentian Great Lakes, which consist of lakes Superior, Michigan, Huron, Erie, and Ontario (Fig. 1a), are the largest group of freshwater lakes on Earth; they contain 21% of the world’s volume of fresh surface water. These lakes have been affected by climate change in several ways including increased surface water temperature, longer summer stratification related hypoxia (i.e., dissolved oxygen [DO] concentrations  0.05); all the black bars are significant (i.e., p  8 m s−1) from the south and southwest that are the common wind directions over the Great Lakes23, tilt the thermocline upward in the western and northern part of the central basin due to Ekman transport of surface water southward22,38,42,43,44,45. As this hypolimnetic water upwells into shallower depths it can be transported counter clockwise by the alongshore surface currents moving to the west32. If there is a calm period following the high winds, the upwelled water in the northwestern part of the central basin will flow southward because of the pressure gradient and also in a clockwise direction (to the west) because of the Coriolis effect, and so will intrude into the western basin (i.e., a geostrophic flow) opposite to the hydraulic flow from the Detroit River (Fig. 1c)22,32,46. This causes the rapid (on the order of hours) formation of a thermocline within the northeastern portion of the western basin (Pelee Passage) due to the intrusion of low temperature bottom water22,42, which can also be hypoxic22 or anoxic (i.e., DO (approx) 0) at the sediment surface22 and contain high soluble reactive phosphorus concentrations (SRP; 0.02–0.05 mg L−1)47,48,49. Low values of sediment oxygen uptake are observed during these events in the western basin due to stratification and weak bottom shear and turbulence, which results in thicker diffusive sublayer22.
    Interbasin exchange has been observed in lakes with multiple basins elsewhere (e.g., Lake Geneva50, Nechako Reservoir51) as well as in the Great Lakes region (e.g., Muskegon Bay52, Green Bay53, Kempenfelt Bay54, Pere Marquette River55). In Lake Michigan, for example, high winds can lead to coastal upwelling into Muskegon Lake causing episodic hypoxia52. In case of Lake Erie, interbasin exchange was identified as the dominant cause (63%) of hypoxia in the northeastern portion of the western basin during biweekly fishing trawls in August over the past 30 years22. However, there are no long-term continuous water quality observations to assess the occurrence and historic trends in these hypoxic events. Extreme winds prevailing from upwelling favourable directions (i.e., from the south and southwest) can generate strong surface waves and water currents through momentum flux at the air–water interface. Therefore, WP can be used as an indicator or proxy (but not the cause) of interbasin exchange. Here, we examine the historical trends in water temperature, winds and resultant waves in the context of climate change in the summer in the Great Lakes (Fig. 1a) with an emphasis on the western basin of Lake Erie (Fig. 1b). We examine data for August, which is the month when hypoxia is most likely to occur in dimictic north temperate lakes before the fall turnover, and when large HAB have been observed in the western basin of Lake Erie. August is also the time when the spatial extent of hypoxia in the central basin is the largest and when the aforementioned upwelling into the western basin is likely to occur22,40,56. The data examined are from buoys with the longest historical records (Fig. 1a and Table S1). We examine winds from the south and southwest directions, which are the common wind directions over the Great Lakes during August, and which are favourable for upwelling into the western basin of Lake Erie. The results show that the WP in Great Lakes has increased in the past 40 years. A pattern in WP (a proxy for hypoxic upwelling events into the western basin of Lake Erie) has also increased in frequency over this time, which has implications for the water quality (e.g., dissolved oxygen and total phosphorus) of the lake. The increased frequency of interbasin upwelling was confirmed using historical records of lake bottom water temperature (LBT), as well as dissolved oxygen and total phosphorus concentrations. This is the first time that WP has been identified as an indicator of climate change-driven biogeochemical responses in lakes.
    Long-term trends in WP and LST in the Great Lakes
    First, we investigate the historical trends in average lake surface temperature (LST), wind, and waves in the Great Lakes during August. Results show that LST and LSTw (hereinafter subscript ”w” is used to denote the variables measured during upwelling favourable winds from 180° to 270°, clockwise from north) have both increased significantly (p  0.2% year−1) since 1980, although lower trends were observed in lake Erie and Michigan (Figs. S1–S5 ((a) and (b)) and Table S1). These changes in the LST correspond to a warming trend in air temperature (Tair); the average Tair over the Great Lakes increased significantly by ~ 0.4 (pm) 0.2 ((pm) standard error) oC decade−1 since 1980 (Fig. S6a,b). There was an associated significant increase in wind speed (W) over the Great Lakes during August (Ww) of ~ 0.4 (pm) 0.1 m s−1 decade−1 for winds from the south and southwest (Figs. S1–S5 ((c) and (d)) and Table S1). Consequently, the wind stress associated with wind from the south and southwest over the water surface of the Great Lakes (({tau }_{w}=0.0012{rho }_{air}{W}_{w}^{2}), where ({rho }_{air})=1.22 kg m−3 is the density of air57, and the wind speed is measured 10 m above the water) increased significantly by 0.006 (pm) 0.002 Pa decade−1 during August (3.0 (pm) 0.9% year−1; Figs. S1–S5 ((e) and (f)) and Table S1).
    The effects of increased wind stress can also be seen in wave power, which is a function of the square of significant wave height (the mean value of the largest third of the wave heights during typically 1 h, SWH) and the wave period (({T}_{p}); i.e., (WP propto {{T}_{p} times SWH}^{2})); and changes in wind are reflected in wave power ((WP propto {W}^{2.4}) and (propto {W}^{5}) for developing and fully developed waves, respectively; see “Materials and methods”). The average SWH and SWHw in the Great Lakes during August have increased significantly by 0.03 (pm) 0.02 and 0.04 (pm) 0.03 m decade−1, respectively (i.e., ~ 1.0 (pm) 0.8% and ~ 1.7 (pm) 1.5% year−1, respectively), and this is largely driven by the increase in the frequency of extreme surface winds58 (Figs. S1–S5g and h; WP responds to changes in mean values, but it is more sensitive to extreme events because WP (propto { SWH}^{2})16). Consequently, the average WP and WPw in the Great Lakes during August have increased by ~ 0.04 (pm) 0.02 and ~ 0.06 (pm) 0.03 kW m−1 decade−1, respectively (i.e., ~ 1.0 (pm) 0.6% and ~ 2.0 (pm) 0.9% year−1, respectively; Fig. 2). In Lake Erie, WPw during August increased significantly by 0.02 (pm) 0.01 kW m−1 decade−1 (1.4 (pm) 0.2% year−1; Fig. 2 and Table S1; the increasing trend in WP = 0.02 (pm) 0.02 or 0.5 (pm) 0.1% was not statistically significant). It is relevant to note that these results are based on observations from a single buoy per lake; the one with the longest available data records (Fig. 1a and Table S2). However, the wind records and historical wave trends between buoys Sta. NDBC 45005 and Port Stanley in Lake Erie (Fig. 1a), which are ~ 130 km apart, are consistent based on the available records. Specifically, wind speed and direction in 2018 have Pearson correlation coefficients, r  > 0.6 (Fig. S7a,b, respectively); Ww and WPw are also correlated with r = 0.51 and 0.67, respectively, during August of 1990–2018 and the buoys show similar temporal increases in WPw (~ 0.025 (pm) 0.02 and 0.02 ± 0.01 kW m−1 decade−1 in Port Stanley and Sta. NDBC 45005, respectively). The trends in historical LSTw and WPw are related statistically (i.e., higher mutual information; Fig. S8) similar to the relationship described for global sea surface temperature and oceanic WP used as an indicator of climate change16.
    Figure 2

    Historical patterns in wave power in Great Lakes. 10 year moving average of wave power (WP) during the August (a) and during August with the wind from south and southwest and (WPw; b). The dashed lines show the linear regression (statistical results provided in Table S1).

    Full size image

    The long-term variations in WP and LST may be related to the global atmospheric phenomena. The LSTw anomaly in all the lakes show an increasing trend beginning in 1995 (Fig. S9a), which corresponds to the switch from the negative mode of the Atlantic Multidecadal Oscillation (AMO) to the positive mode (associated with increased tropical cyclone activity and stronger westerly winds) between the 1980s and the early 2000s (Fig. S9b)16. Both the WPw and LSTw anomaly are positively correlated with the AMO (r ~ 0.50 and ~ 0.55, respectively, since 1990). Similar to global oceanic wave power16, peaks in WPw in the Great Lakes are associated with strong El Niño years (i.e., Multivariate El Niño/Southern Oscillation (MEI) greater than 1.5; Fig. S9c,d), which can contribute to the enhanced wind energy due to increased cyclonic events16. MEI and WPw in Great Lakes are generally correlated by r  > 0.45 since 1990, however, the impacts of global atmospheric events on temperature and water dynamics of Great Lakes requires further study.
    Episodic hypoxic upwelling events in the western basin of Lake Erie
    We used historical records (Table S2) of long-term near-bottom water temperature (1998–2018) and dissolved oxygen (2007–2018) in the northeastern portion of the western basin of Lake Erie as well as wave observations in the western portion of the central basin (1980–2018 in Sta. NDBC 45005, Fig. 1) in August to determine the frequency of hypoxic upwelling events and the impacts of these events on the total phosphorus concentration in the northeast portion of the western basin. These analyses do not include the local hypoxia due to periods of calm and warm atmospheric conditions that may occur annually31 and, which are different than episodic upwelling events. Intrusion of cold hypoxic hypolimnetic water from the central basin into the western basin, following high winds from upwelling favourable directions, can cause a sudden drop (on the order of hours) in LBT and dissolved oxygen (DO) when the hypolimnetic water in the central basin is hypoxic22. The LBT time series in the western basin from 2017 to 2018 show that LBT decreased more than 3 °C in less than 12 h during upwelling events; e.g., 9–16, 18–22 and 26–31 August 2018 at Sta E (Fig. 3b) and 24–29 August 2017 at Leamington and Sta E (Fig. S10b). The records of LBT measured by the Ontario Ministry of Natural Resources and Forestry (MNRF) in August in Leamington Ontario between 1998 and 2018 detected 23 events of intrusion of cold water, which are consistent with upwelling (the blue symbols in Fig. 4a).
    Figure 3

    Wave power and bottom water temperature during August 2018 in the western basin of Lake Erie. (a) Time series of wave power (WP; black line), wave period (Tp; magenta), and significant wave height (SWH; blue) recorded at Sta. NDBC 45005. (b) Time series of dissolved oxygen (DO; red) and water temperature (LBT; blue dashed-line) in Sta. E at 1 m above the bed and bottom water temperature in Leamington (blue solid-line) in August 2018. The red triangles represent the observed hypoxic events in the western basin of Lake Erie. The wave power of the waves from south and southwest (i.e., favourable for upwelling) are positive preceding upwelling.

    Full size image

    Figure 4

    Number of hypoxic upwelling events in the western basin. (a) The number of hypoxic upwelling events based on patterns in wave power at Sta. NDBC 45005 (dark grey: average WPw  > 0.44 kW m−1, light grey: 0.37  8 m s−1 from similar directions, which corresponds to the ~ 80th percentile of wind speeds and is greater than the sum of the average and standard deviation of the wind speed (~ 6 and 2 m s−1, respectively). This wind threshold is consistent with Rao et al.’s44 wind speed that led to upwelling, which resulted in a fish kill along the north shore of the central basin in 2012.
    We used a least-square method to find a wave pattern (i.e., wave direction, duration, and power) that could be applied to predict the number of upwelling events that could be hypoxic between 1998 and 2018 based on LBT observations. A rapid decrease in the LBT at both Sta E and Leamington (12 km vs. 20 km from the Pelee Passage, respectively) occurred during events in which the average WP was  > 0.44 kW m−1 (i.e., 22–24 August 2017; Fig. S10a,b). The model predicted 25 upwelling events at Leamington (dark bars in Fig. 4a) of which 23 were observed (as stated above; no data were available for 2012; blue circles in Fig. 4a) for waves from south and southwest that lasted for at least 15 h with an average wave power greater than 0.37 kW m−1. Of the 23 observed events, the model predicted 21 events providing a root mean square error [RMSE] of 0.20 events. We validated the model predictions using the biweekly DO measurements from MNRF cruises between 2007 and 2018, which happened to sample 17 of the 23 observed events of low LBT. We note, however, that two hypoxic upwelling events were also recorded outside the study period, i.e., early September; this supports the study’s focus on August. Hypoxic conditions (DO  1.6 events year−1 in 2018 based on a 10-year moving average. Specifically, 21 of 49 (~ 43%) upwelling events in the last four decades have occurred in the past 10 years. Thirty-two of these were strong events with WP  > 0.44 kW m−1, 15 of which (~ 47%) occurred after 2009. Interestingly, this pattern in wave power (i.e., waves from south and southwest that last for  > 15 h with an average WP  > 0.37 kW m−1 from the historical data) was also observed in August 1980 (Fig. 4a), when the LBT dropped following rapid formation of a thermocline, which at the time was attributed to the upwelling of hypolimnetic water from the central basin40,42. These results indicate that an increase in extreme winds from south and southwest during August, over the last four decades, has resulted in more frequent upwelling from the central basin into the western basin and consequently a greater number of episodic hypoxic events in that part of Lake Erie.
    The effect of upwelling on phosphorus concentrations was examined through an analysis of the water column-average total phosphorus (TP) observations from biweekly cruises conducted by the MNRF at station W5 (Fig. 1b). We examined the available data recorded between 15 July and 15 September from 2000 to 2018 (3–5 records year−1; 66 observations in total), which is a period in which linear patterns in TP vs. sampling date were not evident (p  >   > 0.05). The z-score (standard deviate) was determined for the data within a given year (({mathrm{Z}}_{mathrm{TP}}=left(mathrm{TP}-{mathrm{TP}}_{mathrm{mean}}right)/mathrm{SD}), where ({mathrm{TP}}_{mathrm{mean}}) is the annual average of TP and SD is the standard deviation). Positive ({mathrm{Z}}_{mathrm{TP}}) values (i.e., (mathrm{TP} >{mathrm{TP}}_{mathrm{mean}})) were observed in 11 cases in which the sampling occurred  1) observed during 5 August–8 September sampling (black solid circles in Fig. 4b). Statistical comparison revealed that the average ({mathrm{Z}}_{mathrm{TP}}) was significantly higher during upwelling vs. non-upwelling samples (i.e., 0.95 ± 0.18, n = 11 vs. − 0.26 ± 0.12, n = 25; ANOVA F1,34 = 29.64, p  More

  • in

    Biological and biochemical diversity in different biotypes of spotted stem borer, Chilo partellus (Swinhoe) in India

    1.
    Dujardin, J. P. Aporte de la genetica poblacional al control y vigilancia de vectores de la enfermedad de Chagas. In Curso Posgrado Genética Poblacional de Triatomineos Aplicada al Control Vectorial de la Enfermedad de Chagas (ed. Guhl, F.) 13–15 (Corcas Editores Ltda, 1997).
    Google Scholar 
    2.
    Pires, H. H. R., Barbosa, S. E., Margonari, C., Jurberg, J. & Diotaiuti, L. Variations of the external male genitalia in three populations of Triatoma infestans Klug, 1834. Minist. Saúde 93(4), 479–483 (1998).
    CAS  Google Scholar 

    3.
    Bambou, A. E. et al. Comparing genetic diversity of Sitophilus zeamais(Motchulsky) populations sampled in several agro-ecological areas between Central African Republic and Senegal. South Asian J. Exp. Biol. 4(4), 172–182 (2014).
    Google Scholar 

    4.
    Baldwin, J. D. & Dingle, H. Geographic variation in the effects of temperature on life-history traits in the large milkweed bug Oncopeltus fasciatus. Oecologia 69, 64–71 (1986).
    ADS  Article  Google Scholar 

    5.
    Blanckenhorn, W. U. Altitudinal life history variation in the dung flies Scathophaga stercoraria and Sepsis cynipsea. Oecologia 109, 342–352 (1997).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    6.
    Ikten, C., Skoda, S. R., Hunt, T. E., Molina-Ochoa, J. & Foster, J. E. Genetic variation and inheritance of diapause induction in two distinct voltine ecotypes of Ostrinia nubilalis (Lepidoptera: Crambidae). Ann. Entomol. Soc. Am. 104, 567–575 (2011).
    Article  Google Scholar 

    7.
    Dhillon, M. K., Hasan, F., Tanwar, A. K. & Bhadauriya, A. P. S. Effects of thermo-photoperiod on induction and termination of hibernation in Chilo partellus (Swinhoe). Bull. Entomol. Res. 107, 294–302 (2017).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    8.
    Sharma, H. C. Host plant resistance to insects in sorghum and its role in integrated pest management. Crop Prot. 12, 11–34 (1993).
    Article  Google Scholar 

    9.
    Dhillon, M. K., Hasan, F., Tanwar, A. K. & Bhadauriya, A. P. S. Factors responsible for aestivation in spotted stem borer, Chilo partellus (Swinhoe). J. Exp. Zool. A 331, 326–340 (2019).
    Article  Google Scholar 

    10.
    Dhillon, M. K. & Hasan, F. Consequences of diapause on post-diapause development, reproductive physiology and population growth of Chilo partellus (Swinhoe). Physiol. Entomol. 43, 196–206 (2018).
    CAS  Article  Google Scholar 

    11.
    Dhillon, M. K., Tanwar, A. K. & Hasan, F. Fitness consequences of delayed mating on reproductive performance of Chilo partellus (Swinhoe). J. Exp. Zool. A 331, 161–167 (2019).
    Article  Google Scholar 

    12.
    Dhillon, M. K. et al. Genetic regulation of diapause and associated traits in Chilo partellus (Swinhoe). Sci. Rep. 10, 1793 (2020).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    13.
    Chown, S. L. & Terblanche, J. S. Physiological diversity in insects: Ecological and evolutionary contexts. Adv. Insect Physiol. 33, 50–152 (2006).
    Article  Google Scholar 

    14.
    Sezonlin, M. et al. Phylogeography and population genetics of the maize stalk borer Busseola fusca (Lepidoptera, Noctuidae) in sub-Saharan Africa. Mol. Ecol. 15(2), 407–420 (2006).
    CAS  PubMed  Article  Google Scholar 

    15.
    Rowntree, J. K., Cameron, D. D. & Preziosi, R. F. Genetic variation changes the interactions between the parasitic plant-ecosystem engineer Rhinanthus and its hosts. Philos. Trans. R. Soc. B 366, 1380–1388 (2011).
    Article  Google Scholar 

    16.
    Giron, D. et al. Promises and challenges in insect–plant interactions. Entomol. Exp. Appl. 166(5), 319–343 (2018).
    Article  Google Scholar 

    17.
    Williams, R. S. & Howells, J. M. Effects of intraspecific genetic variation and prior herbivory in an old-field plant on the abundance of the specialist aphid Uroleucon nigrotuberculatum (Hemiptera: Aphididae). Environ. Entomol. 47, 422–431 (2018).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    18.
    Matsubayashi, K. W., Ohshima, I. & Nosil, P. Ecological speciation in phytophagous insects. Entomol. Exp. Appl. 134, 1–27 (2010).
    Article  Google Scholar 

    19.
    Feder, J. L. et al. Allopatric genetic origins for sympatric host-plant shifts and race formation in Rhagoletis. Proc. Natl Acad. Sci USA 100, 10314–10319 (2003).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    20.
    Althoff, M. D. & Pellmyr, O. Examining genetic structure in bogus yucca moth: A sequential approach to phylogeography. Evolution 56, 1632–1643 (2002).
    PubMed  Article  Google Scholar 

    21.
    Knowles, L. L. & Maddison, W. P. Statistical phylogeography. Mol. Ecol. 11, 2623–2635 (2002).
    PubMed  Article  Google Scholar 

    22.
    Templeton, A. R. Statistical phylogeography: Methods of evaluating and minimizing inference errors. Mol. Ecol. 13, 789–809 (2004).
    PubMed  Article  Google Scholar 

    23.
    Thomas, Y., Bethenod, M. T., Pelozuelo, L., Frérot, B. & Bourguet, D. Genetic isolation between two sympatric host-plant races of the European corn borer, Ostrinia nubilalis Hübner. I. Sex pheromone, moth emergence timing, and parasitism. Evolution 57, 261–273 (2003).
    PubMed  Google Scholar 

    24.
    Sharma, H. C., Taneja, S. L., Kameswara Rao, N. & Prasada Rao, K. E. Evaluation of sorghum germplasm for resistance to insect pests. Inf. Bull. 63, 177 (2003).
    Google Scholar 

    25.
    Sharma, H. C., Dhillon, M. K., Pampapathy, G. & Reddy, B. V. S. Inheritance of resistance to spotted stem borer, Chilo partellus in sorghum, Sorghum bicolor. Euphytica 156, 117–128 (2007).
    Article  Google Scholar 

    26.
    Kanta, U., Dhillon, B. S. & Sekhon, S. S. Evaluation and development of maize germplasm for resistance to spotted stem borer. In Insect Resistant Maize: Recent Advances and Utilization (ed. Mihm, J. A.) 246–254 (Proceedings of an International Symposium CIMMYT, 1997).
    Google Scholar 

    27.
    Rakshit, S. et al. Catalogue of Indian maize inbred lines. Tech. Bull. 3, 40 (2008).
    Google Scholar 

    28.
    Agrawal, A. A. Phenotypic plasticity in the interactions and evolution of species. Science 294, 321–326 (2001).
    ADS  CAS  PubMed  Article  Google Scholar 

    29.
    Stireman, J. O. III., Nason, J. D. & Heard, S. B. Host-associated genetic differentiation in phytophagous insects: general phenomenon or isolated exceptions? Evidence from a goldenrod-insect community. Evolution 59, 2573–2587 (2005).
    CAS  PubMed  Article  Google Scholar 

    30.
    Zytynska, S. E. & Preziosi, R. F. Genetic interactions influence host preference and performance in a plant-insect system. Evol. Ecol. 25, 1321–1333 (2011).
    Article  Google Scholar 

    31.
    Zytynska, S. E. & Preziosi, R. F. Host preference of plant genotypes is altered by intraspecific competition in a phytophagous insect. Arthropod-Plant Interact. 7, 349–357 (2013).
    Article  Google Scholar 

    32.
    Sharma, H. C. Biotechnological Approaches for Pest Management and Ecological Sustainability (CRC Press, 2009).
    Google Scholar 

    33.
    Sharma, H. C. & Dhillon, M. K. Climate change effects on arthropod diversity and its implications for pest management and sustainable crop production. In Agroclimatology: Linking Agriculture to Climate (eds Hatfield, J. L. et al.) 595–619 (Crop Science Society of America and Soil Science Society of America Inc, Madison, WI, 2020).
    Google Scholar 

    34.
    Smith, C. M. Plant Resistance to Arthropods: Molecular and Conventional Approaches (Springer, 2005).
    Google Scholar 

    35.
    Dhillon, M. K. & Sharma, H. C. Paradigm shifts in research on host plant resistance to insect pests. Indian J. Plant Protect. 40(1), 1–11 (2012).
    Google Scholar 

    36.
    Funk, D. J. Isolating a role for natural selection in speciation: Host adaptation and sexual isolation in Neochlamisus bebbianae leaf beetles. Evolution 52, 1744–1759 (1998).
    PubMed  Article  PubMed Central  Google Scholar 

    37.
    Dres, M. & Mallet, J. Host races in plant-feeding insects and their importance in sympatric speciation. Philos. Trans. R. Soc. B 357, 471–492 (2002).
    Article  Google Scholar 

    38.
    Rundle, H. D. & Nosil, P. Ecological speciation. Ecol. Lett. 8, 336–352 (2005).
    Article  Google Scholar 

    39.
    Schluter, D. Evidence for ecological speciation and its alternative. Science 323, 737–741 (2009).
    ADS  CAS  PubMed  Article  Google Scholar 

    40.
    Ishiguro, N. & Tsuchida, K. Polymorphic microsatellite loci for the rice stem borer, Chilo suppressalis (Walker) (Lepidoptera: Crambidae). Appl. Entomol. Zool. 41, 565–568 (2006).
    CAS  Article  Google Scholar 

    41.
    Mukhopadhyay, J., Ghosh, K., Rangel, E. F. & Munstermann, L. E. Genetic variability in biochemical characters of Brazilian field populations of the Leishmania vector, Lutzomyia longipalpis (Diptera: Psychodidae). Am. J. Trop. Med. Hyg. 59(6), 893–901 (1998).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    42.
    Vijaya Lakshmi, P., Amudhan, S., Bindu, K. H., Cheralu, C. & Bentur, J. S. A new biotype of the Asian rice gall midge Orseolia oryzae (Diptera: Cecidomyiidae) characterized from the Warangal population in Andhra Pradesh, India. Int. J. Trop. Insect Sci. 26, 207–211 (2006).
    Google Scholar 

    43.
    Himabindu, K., Suneetha, K., Sama, V. S. A. K. & Bentur, J. S. A new rice gall midge resistance gene in the breeding line CR57-MR1523, mapping with flanking markers and development of NILs. Euphytica 174, 179–187 (2010).
    CAS  Article  Google Scholar 

    44.
    Ratcliffe, R. H. et al. Biotype composition of Hessian fly (Diptera: Cecidomyiidae) populations from the Southeastern, Midwestern, and Northwestern United States and virulence to resistance genes in wheat. J. Econ. Entomol. 93(4), 1319–1328 (2000).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    45.
    Zhou, H. et al. Genetic analysis and fine mapping of the gall midge resistance gene Gm5 in rice (Oryza sativa L.). Theor. Appl. Genet. 133, 2021–2033 (2020).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    46.
    Dhillon, M. K. & Kumar, S. Amino acid profiling of Sorghum bicolor vis-à-vis Chilo partellus (Swinhoe) for biochemical interactions and plant resistance. Arthropod-Plant Interact. 11, 537–550 (2017).
    Article  Google Scholar 

    47.
    Dhillon, M. K. & Kumar, S. Lipophilic profiling of Sorghum bicolor (L.) seedlings vis-à-vis Chilo partellus (Swinhoe) larvae reveals involvement of biomarkers in sorghum-stem borer interactions. Indian J. Exp. Biol. 58, 95–108 (2020).
    CAS  Google Scholar 

    48.
    Atray, I., Bentur, J. S. & Nair, S. The Asian rice gall midge (Orseolia oryzae) mitogenome has evolved novel gene boundaries and tandem repeats that distinguish its biotypes. PLoS ONE 10(7), e0134625 (2015).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    49.
    Fujita, D., Kohli, A. & Horgan, F. G. Rice resistance to planthoppers and leafhoppers. Crit. Rev. Plant Sci. 32, 162–191 (2013).
    CAS  Article  Google Scholar 

    50.
    Diehl, R. S. & Bush, G. L. An evolutionary and applied perspective of insect biotypes. Annu. Rev. Entomol. 29, 471–504 (1984).
    Article  Google Scholar 

    51.
    Claridge, M. F. & Den Hollander, J. A biotype concept and its application to insect pests of agriculture. Crop Prot. 2(1), 85–95 (1983).
    Article  Google Scholar 

    52.
    Downie, D. A. Baubles, bangles, and biotypes: A critical review of the use and abuse of the biotype concept. J. Insect Sci. 10, 176 (2010).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    53.
    Perring, T. M. The Bemisia tabaci species complex. Crop Prot. 20, 725–737 (2001).
    Article  Google Scholar 

    54.
    Wenger, J. A. & Michel, A. P. Implementing an evolutionary framework for understanding genetic relationships of phenotypically defined insect biotypes in the invasive soybean aphid (Aphis glycines). Evol. Appl. 6(7), 1041–1053 (2013).
    PubMed  PubMed Central  Article  Google Scholar 

    55.
    Sharma, H. C., Taneja, S. L., Leuschner, K. & Nwanze, K. F. Techniques to screen sorghum for resistance to insect pests. Inf. Bull. 32, 48 (1992).
    Google Scholar 

    56.
    Kumar, S. & Dhillon, M. K. Lipophilic metabolite profiling of maize and sorghum seeds and seedlings, and their pest spotted stem borer larvae: A standardized GC-MS based approach. Indian J. Exp. Biol. 53, 170–176 (2015).
    PubMed  Google Scholar 

    57.
    Dhillon, M. K., Kumar, S. & Gujar, G. T. A common HPLC-PDA method for amino acid analysis in insects and plants. Indian J. Exp. Biol. 52, 73–79 (2014).
    CAS  PubMed  Google Scholar  More