More stories

  • in

    The potential for a CRISPR gene drive to eradicate or suppress globally invasive social wasps

    1.
    Teem, J. L. et al. Genetic biocontrol for invasive species. Front. Bioeng. Biotechnol. 8, 452. https://doi.org/10.3389/fbioe.2020.00452 (2020).
    Article  PubMed  PubMed Central  Google Scholar 
    2.
    McFarlane, G. R., Whitelaw, C. B. A. & Lillico, S. G. CRISPR-based gene drives for pest control. Trends Biotechnol. 36, 130–133. https://doi.org/10.1016/j.tibtech.2017.10.001 (2018).
    CAS  Article  PubMed  Google Scholar 

    3.
    Dearden, P. K. et al. The potential for the use of gene drives for pest control in New Zealand: a perspective. J. R. Soc. N. Z. 48, 225–244. https://doi.org/10.1080/03036758.2017.1385030 (2017).
    Article  Google Scholar 

    4.
    Esvelt, K. M., Smidler, A. L., Catteruccia, F. & Church, G. M. Concerning RNA-guided gene drives for the alteration of wild populations. eLife 3, e03401. https://doi.org/10.7554/eLife.03401 (2014).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    5.
    Barrangou, R. & Doudna, J. A. Applications of CRISPR technologies in research and beyond. Nat. Biotechnol. 34, 933–941. https://doi.org/10.1038/nbt.3659 (2016).
    CAS  Article  PubMed  Google Scholar 

    6.
    Kandul, N. P. et al. Transforming insect population control with precision guided sterile males with demonstration in flies. Nat. Commun. 10, 84. https://doi.org/10.1038/s41467-018-07964-7 (2019).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    7.
    Kyrou, K. et al. A CRISPR-Cas9 gene drive targeting doublesex causes complete population suppression in caged Anopheles gambiae mosquitoes. Nat. Biotechnol. 36, 1062–1066. https://doi.org/10.1038/nbt.4245 (2018).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    8.
    Drury, D. W., Dapper, A. L., Siniard, D. J., Zentner, G. E. & Wade, M. J. CRISPR/Cas9 gene drives in genetically variable and nonrandomly mating wild populations. Sci. Adv. 3, e1601910. https://doi.org/10.1126/sciadv.1601910 (2017).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    9.
    Hammond, A. M. et al. The creation and selection of mutations resistant to a gene drive over multiple generations in the malaria mosquito. PLoS Genet. 13, e1007039. https://doi.org/10.1371/journal.pgen.1007039 (2017).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    10.
    Webber, B. L., Raghu, S. & Edwards, O. R. Opinion: is CRISPR-based gene drive a biocontrol silver bullet or global conservation threat?. Proc. Natl. Acad. Sci. U.S.A. 112, 10565–10567. https://doi.org/10.1073/pnas.1514258112 (2015).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    11.
    Wilkins, K. E., Prowse, T. A. A., Cassey, P., Thomas, P. Q. & Ross, J. V. Pest demography critically determines the viability of synthetic gene drives for population control. Math. Biosci. 305, 160–169. https://doi.org/10.1016/j.mbs.2018.09.005 (2018).
    MathSciNet  Article  PubMed  MATH  Google Scholar 

    12.
    de la Filia, A. G., Bain, S. A. & Ross, L. Haplodiploidy and the reproductive ecology of Arthropods. Curr. Opin. Insect Sci. 9, 36–43. https://doi.org/10.1016/j.cois.2015.04.018 (2015).
    Article  Google Scholar 

    13.
    Deredec, A., Burt, A. & Godfray, H. C. The population genetics of using homing endonuclease genes in vector and pest management. Genetics 179, 2013–2026. https://doi.org/10.1534/genetics.108.089037 (2008).
    Article  PubMed  PubMed Central  Google Scholar 

    14.
    Rode, N. O., Estoup, A., Bourguet, D., Courtier-Orgogozo, V. & Débarre, F. Population management using gene drive: molecular design, models of spread dynamics and assessment of ecological risks. Conserv. Genet. 20, 671–690. https://doi.org/10.1007/s10592-019-01165-5 (2019).
    CAS  Article  Google Scholar 

    15.
    Alphey, N. & Bonsall, M. B. Interplay of population genetics and dynamics in the genetic control of mosquitoes. J. R. Soc. Interface 11, 20131071. https://doi.org/10.1098/rsif.2013.1071 (2014).
    Article  PubMed  PubMed Central  Google Scholar 

    16.
    Prowse, T. A. A. et al. Dodging silver bullets: good CRISPR gene-drive design is critical for eradicating exotic vertebrates. Proc. R. Soc. B https://doi.org/10.1098/rspb.2017.0799 (2017).
    Article  PubMed  Google Scholar 

    17.
    Lowe, S., Browne, M., Boudjelas, S. & De Poorter, M. 100 of the World’s Worst Invasive Alien Species. A Selection from the Global Invasive Species Database Vol. 12 (The Invasive Species Specialist Group (ISSG) a specialist group of the Species Survival Commission (SSC) of the World Conservation Union (IUCN), Auckland, 2000).
    Google Scholar 

    18.
    Lester, P. J. & Beggs, J. R. Invasion success and management strategies for social Vespula wasps. Annu. Rev. Entomol. 64, 51–71. https://doi.org/10.1146/annurev-ento-011118-111812 (2019).
    CAS  Article  PubMed  Google Scholar 

    19.
    Lester, P. J. et al. Determining the origin of invasions and demonstrating a lack of enemy release from microsporidian pathogens in common wasps (Vespula vulgaris). Divers. Distrib. 20, 964–974. https://doi.org/10.1111/ddi.12223 (2014).
    Article  Google Scholar 

    20.
    Harris, R. J. Diet of the wasps Vespula vulgaris and V. germanica in honeydew beech forest of the South Island, New Zealand. N. Z. J. Zool. 18, 159–169 (1991).
    Article  Google Scholar 

    21.
    Grangier, J. & Lester, P. J. A novel interference behaviour: invasive wasps remove ants from resources and drop them from a height. Biol. Lett. 7, 664–667. https://doi.org/10.1098/rsbl.2011.0165 (2011).
    Article  PubMed  PubMed Central  Google Scholar 

    22.
    Wilson, P. R., Karl, B. J., Toft, R. J., Beggs, J. R. & Taylor, R. H. The role of introduced predators and competitors in the decline of kaka (Nestor meridionalis) populations in New Zealand. Biol. Conserv. 83, 175–185. https://doi.org/10.1016/S0006-3207(97)00055-4 (1998).
    Article  Google Scholar 

    23.
    Dobelmann, J. et al. Fitness in invasive social wasps: the role of variation in viral load, immune response and paternity in predicting nest size and reproductive output. Oikos 126, 1208–1218. https://doi.org/10.1111/oik.04117 (2017).
    CAS  Article  Google Scholar 

    24.
    Sekine, K., Furusawa, T. & Hatakeyama, M. The boule gene is essential for spermatogenesis of haploid insect male. Dev. Biol. 399, 154–163. https://doi.org/10.1016/j.ydbio.2014.12.027 (2015).
    CAS  Article  PubMed  Google Scholar 

    25.
    Ferree, P. M. et al. Identification of genes uniquely expressed in the germ-line tissues of the jewel wasp Nasonia vitripennis. G3-Genes Genom. Genet. 5, 2647–2653. https://doi.org/10.1534/g3.115.021386 (2015).
    CAS  Article  Google Scholar 

    26.
    Mikhaylova, L. M., Boutanaev, A. M. & Nurminsky, D. I. Transcriptional regulation by Modulo integrates meiosis and spermatid differentiation in male germ line. Proc. Natl. Acad. Sci. U.S.A. 103, 11975–11980. https://doi.org/10.1073/pnas.0605087103 (2006).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    27.
    Parsch, J., Meiklejohn, C. D., Hauschteck-Jungen, E., Hunziker, P. & Hartl, D. L. Molecular evolution of the ocnus and janus genes in the Drosophila melanogaster species subgroup. Mol. Biol. Evol. 18, 801–811. https://doi.org/10.1093/oxfordjournals.molbev.a003862 (2001).
    CAS  Article  PubMed  Google Scholar 

    28.
    Dang, Y. et al. Optimizing sgRNA structure to improve CRISPR-Cas9 knockout efficiency. Genome Biol. 16, 280. https://doi.org/10.1186/s13059-015-0846-3 (2015).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    29.
    Yarrington, R. M., Verma, S., Schwartz, S., Trautman, J. K. & Carroll, D. Nucleosomes inhibit target cleavage by CRISPR-Cas9 in vivo. Proc. Natl. Acad. Sci. U.S.A. 115, 9351–9358. https://doi.org/10.1073/pnas.1810062115 (2018).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    30.
    Chaverra-Rodriguez, D. et al. Targeted delivery of CRISPR-Cas9 ribonucleoprotein into arthropod ovaries for heritable germline gene editing. Nat. Commun. 9, 3008. https://doi.org/10.1038/s41467-018-05425-9 (2018).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    31.
    Noble, C. et al. Daisy-chain gene drives for the alteration of local populations. Proc. Natl. Acad. Sci. U.S.A. 116, 8275–8282. https://doi.org/10.1073/pnas.1716358116 (2019).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    32.
    KaramiNejadRanjbar, M. et al. Consequences of resistance evolution in a Cas9-based sex conversion-suppression gene drive for insect pest management. Proc. Natl. Acad. Sci. U.S.A. 115, 6189–6194. https://doi.org/10.1073/pnas.1713825115 (2018).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    33.
    Brenton-Rule, E. C. et al. The origins of global invasions of the German wasp (Vespula germanica) and its infection with four honey bee viruses. Biol. Invasions 20, 3445–3460. https://doi.org/10.1007/s10530-018-1786-0 (2018).
    Article  Google Scholar 

    34.
    Schmack, J. M. et al. Lack of genetic structuring, low effective population sizes and major bottlenecks characterise common and German wasps in New Zealand. Biol. Invasions 21, 3185–3201. https://doi.org/10.1007/s10530-019-02039-0 (2019).
    Article  Google Scholar 

    35.
    Tanaka, H., Stone, H. A. & Nelson, D. R. Spatial gene drives and pushed genetic waves. Proc. Natl. Acad. Sci. U.S.A. 114, 8452–8457. https://doi.org/10.1073/pnas.1705868114 (2017).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    36.
    Hammond, A. et al. A CRISPR-Cas9 gene drive system targeting female reproduction in the malaria mosquito vector Anopheles gambiae. Nat. Biotechnol. 34, 78–83. https://doi.org/10.1038/nbt.3439 (2016).
    CAS  Article  PubMed  Google Scholar 

    37.
    Marshall, J. M., Buchman, A., Sanchez, C. H. & Akbari, O. S. Overcoming evolved resistance to population-suppressing homing-based gene drives. Sci. Rep. 7, 3776. https://doi.org/10.1038/s41598-017-02744-7 (2017).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    38.
    Eckhoff, P. A., Wenger, E. A., Godfray, H. C. & Burt, A. Impact of mosquito gene drive on malaria elimination in a computational model with explicit spatial and temporal dynamics. Proc. Natl. Acad. Sci. U.S.A. 114, E255–E264. https://doi.org/10.1073/pnas.1611064114 (2017).
    CAS  Article  PubMed  Google Scholar 

    39.
    North, A., Burt, A. & Godfray, H. C. Modelling the spatial spread of a homing endonuclease gene in a mosquito population. J. Appl. Ecol. 50, 1216–1225. https://doi.org/10.1111/1365-2664.12133 (2013).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    40.
    Kirk, N., Kannemeyer, R., Greenaway, A., MacDonald, E. & Stronge, D. Understanding attitudes on new technologies to manage invasive species. Pac. Conserv. Biol. https://doi.org/10.1071/pc18080 (2019).
    Article  Google Scholar 

    41.
    Mercier, O. R., KingHunt, A. & Lester, P. J. Novel biotechnologies for eradicating wasps: seeking Māori studies students’ perspectives with Q method. Kōtuitui N. Z. J. Soc. Sci. 14, 136–156. https://doi.org/10.1080/1177083x.2019.1578245 (2019).
    Article  Google Scholar 

    42.
    Peters, R. S. et al. Evolutionary history of the Hymenoptera. Curr. Biol. 27, 1013–1018. https://doi.org/10.1016/j.cub.2017.01.027 (2017).
    CAS  Article  PubMed  Google Scholar 

    43.
    Stein, K. J. & Fell, R. D. Correlation of queen sperm content with colony size in yellowjackets (Hymenoptera: Vespidae). Environ. Entomol. 23, 1497–1500. https://doi.org/10.1093/ee/23.6.1497 (1994).
    Article  Google Scholar 

    44.
    Lester, P. J., Haywood, J., Archer, M. E. & Shortall, C. R. The long-term population dynamics of common wasps in their native and invaded range. J. Anim. Ecol. 86, 337–347. https://doi.org/10.1111/1365-2656.12622 (2017).
    Article  PubMed  Google Scholar 

    45.
    Burt, A. & Deredec, A. Self-limiting population genetic control with sex-linked genome editors. Proc. R. Soc. B https://doi.org/10.1098/rspb.2018.0776 (2018).
    Article  PubMed  Google Scholar 

    46.
    Prowse, T. A., Adikusuma, F., Cassey, P., Thomas, P. & Ross, J. V. A Y-chromosome shredding gene drive for controlling pest vertebrate populations. eLife 8, e41873. https://doi.org/10.7554/eLife.41873 (2019).
    Article  PubMed  PubMed Central  Google Scholar 

    47.
    Li, J. et al. Can CRISPR gene drive work in pest and beneficial haplodiploid species?. Evol. Appl. https://doi.org/10.1111/eva.13032 (2020).
    Article  PubMed  PubMed Central  Google Scholar 

    48.
    Esvelt, K. M. & Gemmell, N. J. Conservation demands safe gene drive. PLoS Biol. 15, e2003850. https://doi.org/10.1371/journal.pbio.2003850 (2017).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    49.
    Piaggio, A. J. et al. Is it time for synthetic biodiversity conservation?. Trends Ecol. Evol. 32, 97–107. https://doi.org/10.1016/j.tree.2016.10.016 (2017).
    Article  PubMed  Google Scholar 

    50.
    Edgington, M. P., Harvey-Samuel, T. & Alphey, L. Population-level multiplexing, a promising strategy to manage the evolution of resistance against gene drives targeting a neutral locus. Evol. Appl. https://doi.org/10.1111/eva.12945 (2020).
    Article  Google Scholar 

    51.
    Sumner, S., Law, G. & Cini, A. Why we love bees and hate wasps. Ecol. Entomol. 43, 836–845. https://doi.org/10.1111/een.12676 (2018).
    Article  Google Scholar 

    52.
    Southon, R. J., Fernandes, O. A., Nascimento, F. S. & Sumner, S. Social wasps are effective biocontrol agents of key lepidopteran crop pests. Proc. R. Soc. B https://doi.org/10.1098/rspb.2019.1676 (2019).
    Article  PubMed  Google Scholar 

    53.
    Harris, R. J., Thomas, C. D. & Moller, H. The influence of habitat use and foraging on the replacement of one introduced wasp species by another in New Zealand. Ecol. Entomol. 16, 441–448. https://doi.org/10.1111/j.1365-2311.1991.tb00237.x (1991).
    Article  Google Scholar 

    54.
    Lester, P. J. et al. Critical issues facing New Zealand entomology. N. Z. Entomol. 37, 1–13. https://doi.org/10.1080/00779962.2014.861789 (2014).
    Article  Google Scholar 

    55.
    Hare, K. M. et al. Intractable: species in New Zealand that continue to decline despite conservation efforts. J. R. Soc. N. Z. 49, 301–319. https://doi.org/10.1080/03036758.2019.1599967 (2019).
    Article  Google Scholar 

    56.
    Hu, X. F., Zhang, B., Liao, C. H. & Zeng, Z. J. High-Efficiency CRISPR/Cas9-mediated gene editing in honeybee (Apis mellifera) embryos. G3-Genes Genom. Genet. 9, 1759–1766. https://doi.org/10.1534/g3.119.400130 (2019).
    CAS  Article  Google Scholar 

    57.
    Yan, H. et al. An engineered orco mutation produces aberrant social behavior and defective neural development in ants. Cell 170, 736-747 e739. https://doi.org/10.1016/j.cell.2017.06.051 (2017).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    58.
    Oksanen, J. et al. vegan: community ecology package. (R package version 2.4-0. https://CRAN.R-project.org/package=vegan, 2016). More

  • in

    Increased rainfall stimulates permafrost thaw across a variety of Interior Alaskan boreal ecosystems

    1.
    Screen, J. A. & Simmonds, I. The central role of diminishing sea ice in recent Arctic temperature amplification. Nature464, 1334–1337 (2010).
    Google Scholar 
    2.
    Serreze, M. C. & Barry, R. G. Processes and impacts of Arctic amplification: a research synthesis. Glob. Plan. Change77.1, 85–96 (2011).
    Google Scholar 

    3.
    Collins, M. et al. in Climate Change 2013: The Physical Science Basis (eds Stocker, T. F. et al.) 1029–1136 (IPCC, Cambridge Univ. Press, 2013).

    4.
    Zhang, X. et al. Enhanced poleward moisture transport and amplified northern high-latitude wetting trend. Nat. Clim. Change3, 47–51 (2013).
    Google Scholar 

    5.
    Kattsov, V. M. et al. Simulation and projection of Arctic freshwater budget components by the IPCC AR4 global climate models. J. Hydrometeor.8, 571–589 (2007).
    Google Scholar 

    6.
    Bintanja, R. & Andry, O. Towards a rain-dominated Arctic. Nat. Clim. Change7, 263–267 (2017).
    Google Scholar 

    7.
    Vihma, T. Effects of Arctic sea ice decline on weather and climate: a review. Surv. Geophys.35, 1175–1214 (2014).
    Google Scholar 

    8.
    van der Kolk, H. J. et al. Potential Arctic tundra vegetation shifts in response to changing temperature, precipitation and permafrost thaw. Biogeosci13, 6229 (2016).
    Google Scholar 

    9.
    Fujinami, H., Yasunari, T. & Watanabe, T. Trend and interannual variation in summer precipitation in eastern Siberia in recent decades. Int. J. Climatol.36, 355–368 (2016).
    Google Scholar 

    10.
    Kopec, B. G., Feng, X., Michel, F. A. & Posmentier, E. S. Influence of sea ice on Arctic precipitation. Proc. Nat. Acad. Sci. USA113, 46–51 (2016).
    Google Scholar 

    11.
    Bintanja, R. & Selten, F. M. Future increases in Arctic precipitation linked to local evaporation and sea-ice retreat. Nature509, 479–482 (2014).
    Google Scholar 

    12.
    Park, T. et al. Changes in growing season duration and productivity of northern vegetation inferred from long-term remote sensing data. Env. Res. Lett.11, 084001 (2016).
    Google Scholar 

    13.
    Koenigk, T. et al. Arctic climate change in 21st century CMIP5 simulations with EC-Earth. Clim. Dyn. 111–12, 2719–2743 (2013).
    Google Scholar 

    14.
    Peterson, B. J. et al. Increasing river discharge to the Arctic Ocean. Science 13298, 2171–2173 (2002).
    Google Scholar 

    15.
    Euskirchen, E. S. et al. Importance of recent shifts in soil thermal dynamics on growing season length, productivity, and carbon sequestration in terrestrial high‐latitude ecosystems. Glob. Change Biol.12, 731–750 (2006).
    Google Scholar 

    16.
    Neumann, R. B. Warming effects of spring rainfall increase methane emissions from thawing permafrost. Geophys. Res. Lett.46, 1393–1401 (2019).
    Google Scholar 

    17.
    Vincent, W. F., Lemay, M. & Allard, M. Arctic permafrost landscapes in transition: towards an integrated Earth system approach. Arct. Sci.3, 39–64 (2017).
    Google Scholar 

    18.
    Payette, S., Delwaide, A., Caccianiga, M. & Beauchemin, M. Accelerated thawing of subarctic peatland permafrost over the last 50 years. Geophys. Res. Lett.31, 18 (2014).
    Google Scholar 

    19.
    Åkerman, H. J. & Johansson, M. Thawing permafrost and thicker active layers in sub‐arctic Sweden. Permafr. Periglac. Proc.19, 279–292 (2008).
    Google Scholar 

    20.
    Iijima, Y. et al. Abrupt increases in soil temperatures following increased precipitation in a permafrost region, central Lena River basin, Russia. Permafr. Periglac. Proc.21, 30–41 (2010).
    Google Scholar 

    21.
    Jorgenson, M. T., Shur, Y. L. & Pullman, E. R. Abrupt increase in permafrost degradation in Arctic Alaska. Geophys. Res. Lett.33, 2 (2006).
    Google Scholar 

    22.
    Kokfelt, U. et al. Ecosystem responses to increased precipitation and permafrost decay in subarctic Sweden inferred from peat and lake sediments. Glob. Change Biol.15, 1652–1663 (2009).
    Google Scholar 

    23.
    Loranty, M. M. et al. Changing ecosystem influences on soil thermal regimes in northern high-latitude permafrost regions. Biogeosci.15, 5287–5313 (2018).
    Google Scholar 

    24.
    Shiklomanov, N. I., Streletskiy, D. A., Little, J. D. & Nelson, F. E. Isotropic thaw subsidence in undisturbed permafrost landscapes. Geophys. Res. Lett.40, 6356–6361 (2013).
    Google Scholar 

    25.
    Shur, Y. L. & Jorgenson, M. T. Patterns of permafrost formation and degradation in relation to climate and ecosystems. Permafr. Periglac. Proc.18, 7–19 (2007).
    Google Scholar 

    26.
    Jorgenson, M. T. et al. Resilience and vulnerability of permafrost to climate change. Can. J. Res.40, 1219–1236 (2010).
    Google Scholar 

    27.
    Chapin, F. S. III, Van Cleve, K. & Chapin, M. C. Soil temperature and nutrient cycling in the tussock growth form of Eriophorum vaginatum. J. Ecol. Mar.1, 169–189 (1979).
    Google Scholar 

    28.
    Fisher, J. P. et al. The influence of vegetation and soil characteristics on active‐layer thickness of permafrost soils in boreal forest. Glob. Change Biol.22, 3127–3140 (2016).
    Google Scholar 

    29.
    Brown, D. et al. Interactive effects of wildfire and climate on permafrost degradation in Alaskan lowland forests. J. Geophys. Res. Biogeosci.120, 1619–1637 (2015).
    Google Scholar 

    30.
    Jafarov, E. E., Romanovsky, V. E., Genet, H., McGuire, A. D. & Marchenko, S. S. The effects of fire on the thermal stability of permafrost in lowland and upland black spruce forests of interior Alaska in a changing climate. Environ. Res. Lett.8, 035030 (2013).
    Google Scholar 

    31.
    Vogelmann, J. E. et al. Completion of the 1990s National Land Cover Data Set for the conterminous United States from Landsat Thematic Mapper data and ancillary data sources. Photogram. Engin. Rem. Sens.67, 6 (2001).
    Google Scholar 

    32.
    Van Tatenhove, F. G. & Olesen, O. B. Ground temperature and related permafrost characteristics in West Greenland. Permafr. Periglac. Proc.5, 199–215 (1994).
    Google Scholar 

    33.
    Hinkel, K. M., Paetzold, F., Nelson, F. E. & Bockheim, J. G. Patterns of soil temperature and moisture in the active layer and upper permafrost at Barrow, Alaska: 1993–1999. Glob. Planet. Change29, 293–309 (2001).
    Google Scholar 

    34.
    Harris, S. A. Causes and consequences of rapid thermokarst development in permafrost or glacial terrain. Permafr. Periglac. Proc.13, 237–242 (2002).
    Google Scholar 

    35.
    Ling, F. & Zhang, T. A numerical model for surface energy balance and thermal regime of the active layer and permafrost containing unfrozen water. Cold Reg. Sci. Technol.38, 1–5 (2004).
    Google Scholar 

    36.
    Price, A. G., Dunham, K., Carleton, T. & Band, L. Variability of water fluxes through the black spruce (Picea mariana) canopy and feather moss (Pleurozium schreberi) carpet in the boreal forest of Northern Manitoba. J. Hydrol.196, 310–323 (1997).
    Google Scholar 

    37.
    Hamada, S. et al. Hydrometeorological behaviour of pine and larch forests in eastern Siberia. Hydrol. Proc.18, 23–39 (2004).
    Google Scholar 

    38.
    Goetz, J. D. & Price, J. S. Role of morphological structure and layering of Sphagnum and Tomenthypnum mosses on moss productivity and evaporation rates. Can. J. Soil Sci.95, 109–124 (2015).
    Google Scholar 

    39.
    Liljedahl, A. K. et al. Pan-Arctic ice-wedge degradation in warming permafrost and its influence on tundra hydrology. Nat. Geosci.9, 312 (2016).
    Google Scholar 

    40.
    Gibson, C. M. et al. Wildfire as a major driver of recent permafrost thaw in boreal peatlands. Nat. Comm.9, 3041 (2018).
    Google Scholar 

    41.
    Lewkowicz, A. G. & Way, R. G. Extremes of summer climate trigger thousands of thermokarst landslides in a High Arctic environment. Nat. Comm.10, 1329 (2019).
    Google Scholar 

    42.
    Johnstone, J. F. et al. Fire, climate change, and forest resilience in interior Alaska. Can. J. Res.40, 1302–1312 (2010).
    Google Scholar 

    43.
    Juszak, I., Eugster, W., Heijmans, M. M. & Schaepman-Strub, G. Contrasting radiation and soil heat fluxes in Arctic shrub and wet sedge tundra. Biogeosci.13, 404940–404964 (2016).
    Google Scholar 

    44.
    Shur, Y., Hinkel, K. M. & Nelson, F. E. The transient layer: implications for geocryology and climate‐change science. Permafr. Periglac. Proc.16, 5–17 (2005).
    Google Scholar 

    45.
    Barker, A. J. et al. Late season mobilization of trace metals in two small Alaskan arctic watersheds as a proxy for landscape scale permafrost active layer dynamics. Chem. Geo.381, 180–193 (2014).
    Google Scholar 

    46.
    Khosh, M. S. et al. Seasonality of dissolved nitrogen from spring melt to fall freezeup in Alaskan Arctic tundra and mountain streams. J. Geophys. Res. Biogeosci.122, 1718–1737 (2017).
    Google Scholar 

    47.
    Loiko, S. V. et al. Abrupt permafrost collapse enhances organic carbon, CO2, nutrient and metal release into surface waters. Chem. Geo.471, 153–165 (2017).
    Google Scholar 

    48.
    Jorgenson, M. T., Racine, C. H., Walters, J. C. & Osterkamp, T. E. Permafrost degradation and ecological changes associated with a warming climate in central Alaska. Clim. Change48, 551–579 (2001).
    Google Scholar 

    49.
    Douglas, T., Jones, M. C., Hiemstra, C. A. & Arnold, J. R. Sources and sinks of carbon in boreal ecosystems of interior Alaska: A review. Elem. Sci. Anth.12, 2 (2014).
    Google Scholar 

    50.
    Douglas, T. A. et al. Degrading permafrost mapped with electrical resistivity tomography, airborne imagery and LiDAR, and seasonal thaw measurements. Geophys81, WA71-WA85 (2015).
    Google Scholar  More

  • in

    Infrared spectroscopy refines chronological assessment, depositional environment and pyrolysis conditions of archeological charcoals

    The relevant bands that were used for sample evaluation are compiled in Table 1. Band positions are indicated according to Smith47 and Guo and Bustin48. The list is limited to visible band maxima. Aging/oxidation lead to interactions (e.g. H-bonds) and subsequently to broadening of bands49. Strong bands of inorganic components overlap smaller organic bands. Nevertheless, underlying features are included by multivariate data analysis of the spectral pattern. Data analyses were performed with selected wavenumber regions.
    Table 1 Wavenumber position and assignment of functional groups.
    Full size table

    Stepwise procedure in the interpretation of sample sets from additional sites
    Principal Component Analysis (PCA) of infrared spectra (wavenumber regions of 4,000–2,400 cm−1 and 1,800–400 cm−1) is performed for all samples (reference samples and dated sample sets). If their scores correspond to the age determined by the reference method, it is indicative that alteration proceeds according to the spectral pattern of the reference samples. Otherwise, further investigations are necessary as described below.
    Trend of the spectral pattern of natural charcoal aging
    Due to the determined age, it can be assumed that all samples were pyrolyzed at temperatures  > 400 °C. Charcoals produced at lower temperatures are more affected by microbial degradation as shown by incubation50, field experiments42 and chemical oxidation17. During the aging process a common succession of the spectral changes is noticeable. Figure 1a and b display the PCA (4,000–2,400 cm−1 and 1,800–400 cm−1) of reference samples from Austria (recent, A, B, C) and sample sets of charcoals from Brazil (Rio) and Germany (Wittnau WI and Iznang IZ). Despite the heterogeneity within the group, their scores in PC1 are proportional to age. The loadings plot (Fig. 1b) clearly reveals the relevance of the organic bands with regard to the aging process.
    Figure 1

    (a) Scores plot and (b) corresponding loadings plot of the PCA based on the wavenumber regions 4,000–2,400 cm−1 and 1,800–400 cm−1, (c) average spectra (wavenumber range 1,800–1,100 cm−1) of Austrian reference samples (rec, A1800 CE, B13th–early 15th cent. CE, C15th–13th cent. BCE) and samples from Brazil (Rio18th–19th cent. CE) and Germany (WI15th–17th cent. CE and IZ3280–3250 BCE).

    Full size image

    The average infrared spectra of each sample set are shown in Fig. 1c to support the loadings interpretation. They feature the characteristic development of relevant bands in the region from 1,800 cm−1 to 1,100 cm−1: the increase in intensity of the carboxylate bands (1,585–1,565 cm−1 and 1,385–1,375 cm−1) and the concomitant increase, followed by a relative decrease in the oldest samples, of the carboxylic acid bands at about 1705 cm−1 and 1,260–1,250 cm−1. This decline is observed starting from the fifteenth to the thirteenth century BCE (reference samples “C”). The samples from Brazil (Rio), Wittnau (WI) and Iznang (IZ) fit in the series according to their determined age. Emerging functional groups and changes of band intensities in the carboxylic/carboxylate region are paralleled by a corresponding increase of the O–H stretch band in the spectral region 3,500–2,500 cm−1, which is in accordance with the increasing hydrophilicity due to oxidation (as shown in Fig. 3). Despite different sampling sites (Germany, Brazil) and therefore different climatic conditions, the oxidation process follows a useful regularity. The degree of carbonization seems more important than some differences in the environment, as confirmed by litterbag experiments, where degradation was generally highest in 500 °C chars and lowest in 300 °C chars, independent of storage conditions such as soil surface, litter, or layer of limestone chips42. Changes regarding drought, humidity and temperatures might be counterbalanced over the long-lasting residence time in the environment. Nevertheless, some environmental conditions have a strong impact on alteration or preservation15. Heterogeneity within groups is pronounced from the nineteenth to the thirteenth century CE, whereas older groups become more uniform. Over a period of millennia, the relevance of individual degrees of carbonization or environmental exposure abates and only samples with high degrees of carbonization and appropriate environmental conditions remain.
    Archeological sites Bodnegg and Olzreute
    Two sample sets (Bodnegg and Olzreute) do not fit in the series of reference samples in terms of their spectral features. Despite the age of several thousand years (3950–3650 BCE and 2900–2820 BCE, respectively), sample positions in the scores plot of the PCA (not shown), based on all reference samples and the wavenumber regions 4,000–2,400 cm−1 and 1,800–400 cm−1, are close to recent samples (rec) or overlap samples from reference set “A” (about 1800 CE). This first PCA was the basis for the second PCA (Fig. 2), which emphasizes these two relevant periods. This second PCA used the wavenumber regions 4,000–2,400 cm−1 and 1,800–1,100 cm−1. The wavenumber region from 1,100 to 400 cm−1 was excluded, as mineral compounds were not visible in the spectra.
    Figure 2

    Scores plot of the PCA based on infrared spectra (wavenumber regions 4,000–2,400 cm−1 and 1,800–1,100 cm−1) of reference samples (rec, A) and (a) samples from Bodnegg (BO) and (b) samples from Olzreute (OL); squares indicate samples (BO1, BO2, OL1, OL2) for single spectra (Fig. 3) and 14C analyses (see below).

    Full size image

    Figure 3

    Average infrared spectra (wavenumber regions 4,000–2,400 cm−1 and 1,800–1,100 cm−1) of reference samples (rec, A, B) and single spectra of marked samples (Fig. 2) from Bodnegg (BO1 and BO2) and Olzreute (OL1 and OL2).

    Full size image

    According to their position in the scores plot, close to either recent samples (rec) or reference samples “A”, single spectra of the marked samples in Fig. 2 from both groups (BO1 and BO2, OL1 and OL2) are displayed in Fig. 3. The wide variability of intensities in the carboxyl/carboxylate region indicates different partial oxidation degrees among different BO and OL samples (within-group variation).
    As the spectra of BO2 and OL2 feature properties of reference samples C (increase of the bands at 1,585–1,565 cm−1 and 1,385–1,375 cm−1 and the concomitant decrease of the bands at 1,705 cm−1 and 1,260 cm−1), a SIMCA (Soft Independent Modeling of Class Analogy) was calculated to find out the membership of the samples.
    Most of the samples are located in the area “neither–nor”, but closer to the class “rec” than “C”. Only 7 out of 125 samples from Bodnegg, and none from Olzreute, are assigned to the class “C”, as would be expected from their determined age.
    The similarity to recent samples or reference samples “A” (i.e., low degree of partial oxidation) indicates some protective mechanism from aging or high charcoal recalcitrance, which could be provided by a high degree of carbonization. The minor evidence of aging and the discrepancy between the indicated age and the spectral signature required additional investigation.
    The scores plot of the PCA (Fig. 2) and the Coomans plot (Fig. 4) reveal the heterogeneity within these groups (BO and OL), which raises the question of whether charcoals with a wider range of age coexist in the same sites. In such cases FT-IR spectroscopy provides advantageous information as it allows analyses of huge sample sets due to low costs and a rapid procedure. The spectral feature can be used for sample screening to confirm previous dating results or to initiate additional investigations. Two samples from each site, Olzreute and Bodnegg, with the highest distance in the scores plot (see squares in Fig. 2) were subjected to 14C-dating, which confirmed the same age for both contrasting BO- and OL-samples. Therefore, we can conclude that a high degree of carbonization and/or special environmental conditions are responsible for the preservation.
    Figure 4

    Coomans plot representing the membership of samples from Bodnegg (BO) and Olzreute (OL); classification by a SIMCA model based on infrared spectra (wavenumber regions 4,000–2,400 cm−1 and 1,800–1,100 cm−1) with defined classes “rec” and “C”15th–13th cent. BCE; significance level 5% (black lines).

    Full size image

    In the next step, the carbonization temperature that the wood was exposed to, was determined based on spectral characteristics using an established prediction model51. It has to be emphasized that the prediction model has been calibrated with fresh charcoals from laboratory experiments51 and applied on recent traditional kiln processes52. The aging effect is not considered in the model. Prediction results for the current sample sets are presented in Fig. 5a and indicate that many charcoal samples from Olzreute (OL) and Bodnegg (BO) were subjected to similar temperatures as kiln samples52. Comparison of standard deviations of all sample sets (Fig. 5b) confirms that the application of the temperature prediction model is limited to charcoal samples that are similar to recent charcoals. High standard deviations indicate that the material departed from material characteristics of the calibrated recent charcoals.
    Figure 5

    Boxplots representing (a) the temperature prediction of recent samples and samples from Bodnegg (BO) and Olzreute (OL) and (b) the standard deviations of temperature prediction for all charcoal sample sets.

    Full size image

    It has to be emphasized that the sample set “rec” comprises samples with a high carbonization degree. It is not evident that such carbonization conditions can be presupposed for all pyrolysis processes to which charcoals were subjected at archeological sites. Reflectance measurements might provide more information about the production temperature, at least for  > 400 °C35,48. The boxplots show that most samples BO and OL had been exposed to temperatures  > 400 °C, half of them even  > 580 °C corresponding to high thermal alteration.
    Apart from the high degree of carbonization in the BO and OL samples, environmental conditions have to be considered. According to the archeological information about the sampling sites, charcoals from Bodnegg and Olzreute were embedded in a permanently wet peat. As oxygen availability and accessibility are essential factors for degradation, anoxic environments such as waterlogged ecosystems, peats and river sediments foster preservation15. Anaerobic environmental conditions over longer periods of time, together with the high degree of carbonization, seem to be responsible for the good state of preservation. Lab experiments with a strong chemical oxidative reagent revealed a considerable resistance of charcoals produced at 600 °C against oxidation17.
    Sample set with partially high content of mineral compounds (site Speckhau)
    The spectral pattern of the sample set from Speckhau (SP) conspicuously indicates the environment where the charcoals were buried. These charcoals originated from a tumulus and were embedded in a mineral matrix. Mineral components (silicates, clay) had permeated the pores and could not be removed without additional chemical methods. According to Huisman et al.53, who investigated remains of a Neolithic settlement, alkalinity is a main factor for charcoal alteration and pedological processes that lead to clay coatings. Figure 6a displays the average spectra of both samples containing high mineral contents (SP-H) and samples with a low content (SP-L).
    Figure 6

    (a) Average FTIR-ATR spectra of samples with low (SP-L) and high (SP-H) mineral content; (b) scores plot of the PCA based on infrared spectra (wavenumber range 4,000–2,400 cm−1 and 1,800–400 cm−1) for all samples (SP) and reference samples (rec, A, B, C); (c) corresponding loadings plot of the 1st and the 2nd PC; (d) scores plot of the Varimax-Rotation based on infrared spectra (wavenumber range 4,000–2,400 cm−1 and 1,800–400 cm−1) for all samples (SP) and reference samples (rec, A, B, C); (e) corresponding loadings plot of the rotated component 1 and rotated component 2.

    Full size image

    The high mineral content with intense infrared bands (e.g. Si–O at about 1,030 cm−1)54 obliterates other bands, with organic indicator bands disappearing almost completely. In 21 (SP-L) out of 40 samples the characteristic indicator bands of aging are at least visible. The PCA (Fig. 6b) based on infrared spectra of the whole sample set in the wavenumber range 4,000–2,400 cm−1 and 1,800–400 cm−1 illustrates the conspicuous heterogeneity due to different portions of mineral components. The corresponding loadings plot (Fig. 6c) of the 1st and the 2nd Principal Component (PC) reveals the spectral regions that are responsible for the main variance (explained variance by PC1 and PC2: 74% and 18%, respectively). Besides the characteristic spectral regions that represent the aging process ( > 1,200 cm−1), the dominant contribution of mineral components with atom pairs with large reduced mass, such as Si–O, Al–O, Fe–O etc., resulting in stretching bands at low wavenumbers (region  More

  • in

    Fungus-growing insects host a distinctive microbiota apparently adapted to the fungiculture environment

    1.
    Cragg, S. M. et al. Lignocellulose degradation mechanisms across the tree of Life. Curr. Opin. Chem. Bio. 29, 108–119. https://doi.org/10.1016/j.cbpa.2015.10.018 (2015).
    Article  CAS  Google Scholar 
    2.
    Sticklen, M. B. Plant genetic engineering for biofuel production: towards affordable cellulosic ethanol. Nat. Rev. Genet. 9, 433–443. https://doi.org/10.1038/nrg2336 (2008).
    Article  PubMed  CAS  Google Scholar 

    3.
    Guerriero, G., Hausman, J., Strauss, J., Ertan, H. & Siddiqui, K. S. Lignocellulosic biomass: biosynthesis, degradation, and industrial utilization. Eng. Life Sci. 16, 1–16. https://doi.org/10.1002/elsc.201400196 (2016).
    Article  CAS  Google Scholar 

    4.
    Morrison, M., Pope, P. B., Denman, S. E. & McSweeney, C. S. Plant biomass degradation by gut microbiomes: more of the same or something new?. Curr. Opin. Biotechnol. 20, 358–363. https://doi.org/10.1016/j.copbio.2009.05.004 (2009).
    Article  PubMed  CAS  Google Scholar 

    5.
    Karasov, W. H., del Rio, C. M. & Caviedes-Vidal, E. Ecological physiology of diet and digestive systems. Annu. Rev. Physiol. 73, 69–93. https://doi.org/10.1146/annurev-physiol-012110-142152 (2011).
    Article  PubMed  CAS  Google Scholar 

    6.
    Engel, P. & Moran, N. A. The gut microbiota of insects — diversity in structure and function. FEMS Microbiol. Rev. 37, 699–735. https://doi.org/10.1111/1574-6976.12025 (2013).
    Article  PubMed  CAS  Google Scholar 

    7.
    Hansen, A. K. & Moran, N. A. The impact of microbial symbionts on host plant utilization by herbivorous insects. Mol. Ecol. 23, 1473–1496. https://doi.org/10.1111/mec.12421 (2013).
    Article  PubMed  Google Scholar 

    8.
    Kohl, K. D., Connelly, J. W., Dearing, M. D. & Forbey, J. S. Microbial detoxification in the gut of a specialist avian herbivore, the Greater Sage-Grouse. FEMS Microbiol.Lett. 363, fnw144. https://doi.org/10.1093/femsle/fnw144 (2016).
    Article  PubMed  CAS  Google Scholar 

    9.
    Mueller, U. G., Gerardo, N. M., Aanen, D. K., Six, D. L. & Schultz, T. R. The evolution of agriculture in insects. Annu. Rev. Ecol. Evol. Syst. 36, 563–595. https://doi.org/10.1146/annurev.ecolsys.36.102003.152626 (2005).
    Article  Google Scholar 

    10.
    Mayhé-Nunes, A. J. & Jaffé, K. On the biogeography of Attini (Hymenoptera: Formicidae). Ecotropicos 11, 45–54 (1998).
    Google Scholar 

    11.
    Ward, P. S., Brady, S. G., Fisher, B. L. & Schultz, T. R. The evolution of myrmicine ants: phylogeny and biogeography of a hyperdiverse ant clade (Hymenoptera: Formicidae). Syst. Entomol. 40, 61–81. https://doi.org/10.1111/syen.12090 (2015).
    Article  Google Scholar 

    12.
    Jordal, B. H. & Cognato, C. Molecular phylogeny of bark and ambrosia beetles reveals multiple origins of fungus farming during periods of global warming. BMC Evol. Biol. 12, 133. https://doi.org/10.1186/1471-2148-12-133 (2012).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    13.
    Nobre, T., Rouland-Lefevre, C. & Aanen, D. K. Comparative biology of fungus cultivation in termites and ants. In Biology of termites: a modern synthesis, Chapter 8, 193–210 (eds Bignell, D. E. et al.) (Springer, Berlin, 2011).
    Google Scholar 

    14.
    Aylward, F. O. et al. Leucoagaricus gongylophorus produces diverse enzymes for the degradation of recalcitrant plant polymers in leaf-cutter ant fungus gardens. Appl. Environ. Microbiol. 79, 3770–3778. https://doi.org/10.1128/AEM.03833-12 (2013).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    15.
    Khadempour, L. et al. The fungal cultivar of leaf-cutter ants produces specific enzymes in response to different plant substrates. Mol. Ecol. 25, 5795–5805. https://doi.org/10.1111/mec.13872 (2016).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    16.
    Vigueras, G. et al. Growth and enzymatic activity of Leucoagaricus gongylophorus, a mutualistic fungus isolated from the leaf-cutting ant Atta mexicana, on cellulose and lignocellulosic biomass. Lett. Appl. Microbiol. 65, 173–181. https://doi.org/10.1111/lam.12759 (2017).
    Article  PubMed  CAS  Google Scholar 

    17.
    Poulsen, M. et al. Complementary symbiont contributions to plant decomposition in a fungus-farming termite. Proc. Natl. Acad. Sci. USA 111, 14500–14505. https://doi.org/10.1073/pnas.1319718111 (2014).
    ADS  Article  PubMed  CAS  Google Scholar 

    18.
    Hyodo, F., Inoue, T., Azuma, J. I., Tayasu, I. & Abe, T. Role of the mutualistic fungus in lignin degradation in the fungus-growing termite Macrotermes gilvus (Isoptera; Macrotermitinae). Soil Biol. Biochem. 32, 653–658. https://doi.org/10.1016/S0038-0717(99)00192-3 (2000).
    Article  CAS  Google Scholar 

    19.
    Hyodo, F. et al. Differential role of symbiotic fungi in lignin degradation and food provision for fungus-growing termites (Macrotermitinae: Isoptera). Funct. Ecol. 17, 186–193. https://doi.org/10.1046/j.1365-2435.2003.00718.x (2003).
    Article  Google Scholar 

    20.
    De Fine Lich, H. H. & Biedermann, P. H. W. Patterns of functional enzyme activity in fungus farming ambrosia beetles. Front. Zool. 9, 13. https://doi.org/10.1186/1742-9994-9-13 (2012).
    Article  CAS  Google Scholar 

    21.
    Lange, L. & Grell, M. N. The prominent role of fungi and fungal enzymes in the ant–fungus biomass conversion symbiosis. Appl. Microbiol. Biotechnol. 98, 4839–4851. https://doi.org/10.1007/s00253-014-5708-5 (2014).
    Article  PubMed  CAS  Google Scholar 

    22.
    Collins, N. M. The role of termites in the decomposition of wood and leaf litter in the Southern Guinea savanna of Nigeria. Oecologia 51, 389–399. https://doi.org/10.1007/BF00540911 (1981).
    ADS  Article  PubMed  CAS  Google Scholar 

    23.
    Beaver, R. A. Insect-fungus relationships in the bark and ambrosia beetles. In Insect-fungus interactions (eds Wilding, N. et al.) 121–143 (Academic Press, Cambridge, 1989).
    Google Scholar 

    24.
    Kok, L. T., Norrisd, M. & Chu, H. M. Sterol metabolism as a basis for mutualistic symbiosis. Nature 225, 661–662. https://doi.org/10.1038/225661b0 (1970).
    ADS  Article  PubMed  CAS  Google Scholar 

    25.
    Six, D. L. Ecological and evolutionary determinants of bark beetle-fungus symbioses. Insects 3, 339–366. https://doi.org/10.3390/insects3010339 (2012).
    Article  PubMed  PubMed Central  Google Scholar 

    26.
    Pinto-Tomás, A. A. et al. Symbiotic nitrogen fixation in the fungus gardens of leaf-cutter ants. Science 326, 1120–1123. https://doi.org/10.1126/science.1173036 (2009).
    ADS  Article  PubMed  CAS  Google Scholar 

    27.
    Suen, G. et al. An insect herbivore microbiome with high plant biomass degrading capacity. PLoS Genet. 6, e1001129. https://doi.org/10.1371/journal.pgen.1001129 (2010).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    28.
    Aylward, F. O. et al. Metagenomic and metaproteomic insights into bacterial communities in leaf-cutter ant fungus gardens. ISME J. 6, 1688–1701. https://doi.org/10.1038/ismej.2012.10 (2012).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    29.
    Haanstad, J. O. & Norris, D. M. Microbial symbiotes of the ambrosia beetle Xyletorinus politus. Microb. Ecol. 11, 267–276. https://doi.org/10.1007/BF02010605 (1985).
    Article  PubMed  CAS  Google Scholar 

    30.
    Grubbs, K. J. et al. Genome sequence of Streptomyces griseus  strain XyelbKG-1, an ambrosia beetle associated actinomycete. J. Bacteriol. 193, 2890–2891. https://doi.org/10.1128/JB.00330-11 (2011).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    31.
    Scott, J. J. et al. Bacterial protection of beetle-fungus mutualism. Science 322, 63. https://doi.org/10.1126/science.1160423 (2008).
    ADS  Article  PubMed  PubMed Central  CAS  Google Scholar 

    32.
    Boone, C. K. Bacteria associated with a tree-killing insect reduce concentrations of plant defense compounds. J. Chem. Ecol. 39, 1003–1006. https://doi.org/10.1007/s10886-013-0313-0 (2013).
    Article  PubMed  CAS  Google Scholar 

    33.
    Xu, L.-T., Lu, M. & Sun, J.-H. Invasive bark beetle-associated microbes degrade a host defensive monoterpene. Insect Sci. 23, 183–190. https://doi.org/10.1111/1744-7917.12255 (2016).
    Article  PubMed  CAS  Google Scholar 

    34.
    Um, S., Fraimout, A., Sapountzis, P., Oh, D.-C. & Poulsen, M. The fungus-growing termite Macrotermes natalensis harbors bacillaene-producing Bacillus sp. that inhibit potentially antagonistic fungi. Sci. Rep. 3, 3250. https://doi.org/10.1038/srep03250 (2013).
    Article  PubMed  PubMed Central  Google Scholar 

    35.
    Li, H. et al. Lignocellulose pretreatment in a fungus-cultivating termite. Proc. Natl. Acad. Sci. USA 114, 4709–4714. https://doi.org/10.1073/pnas.1618360114 (2017).
    ADS  Article  PubMed  CAS  Google Scholar 

    36.
    Aylward, F. O. et al. Convergent bacterial microbiotas in the fungal agricultural systems of insects. mBio 5, e02077-14. https://doi.org/10.1128/mBio.02077-14 (2014).
    Article  PubMed  PubMed Central  Google Scholar 

    37.
    Stayton, C. T. The definition, recognition, and interpretation of convergent evolution, and two new measures for quantifying and assessing the significance of convergence. Evolution 69, 2140–2153. https://doi.org/10.1111/evo.12729 (2015).
    Article  PubMed  Google Scholar 

    38.
    Arbuckle, K. & Speed, M. P. Analysing convergent evolution: a practical guide to methods. In Evolutionary biology: convergent evolution, evolution of complex traits, concepts and methods, Chapter 2, (ed. Pontarotti, P.) 23–36 (Springer, Berlin , 2016).
    Google Scholar 

    39.
    Martiny, J. B. H., Jones, S. E., Lennon, J. T. & Martiny, A. C. Microbiomes in light of traits: a phylogenetic perspective. Science 350, 9323. https://doi.org/10.1126/science.aac9323 (2015).
    ADS  Article  CAS  Google Scholar 

    40.
    Rabeling, C., Verhaagh, M. & Engels, W. Comparative study of nest architecture and colony structure of the fungus-growing ants, Mycocepurus goeldii and M. smithii. J. Insect. Sci. 7, 40. https://doi.org/10.1673/031.007.4001 (2007).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    41.
    Zanetti, R. et al. An overview of integrated management of leaf-cutting ants (Hymenoptera: Formicidae) in Brazilian forest plantations. Forests 5, 439–454. https://doi.org/10.3390/f5030439 (2014).
    Article  Google Scholar 

    42.
    Markowitz, V. M. et al. IMG/M-HMP: a metagenome comparative analysis system for the human microbiome project. PLoS ONE 7, e40151. https://doi.org/10.1371/journal.pone.0040151 (2012).
    ADS  Article  PubMed  PubMed Central  CAS  Google Scholar 

    43.
    Adams, A. S. et al. Mountain pine beetles colonizing historical and naïve host trees are associated with a bacterial community highly enriched in genes contributing to terpene metabolism. Appl. Environ. Microbiol. 79, 3468–3475. https://doi.org/10.1128/AEM.00068-13 (2013).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    44.
    Solheim, H. Oxygen deficiency and spruce resin inhibition of growth of blue stain fungi associated with Ips typographus. Mycol. Res. 95, 1387–1392. https://doi.org/10.1016/S0953-7562(09)80390-0 (1991).
    Article  Google Scholar 

    45.
    Schuurman, G. H. Ecosystem influences of fungus-growing termites in the dry Paleotropics. In Soil ecology and ecosystem services, Chapter 34 (eds Wall, D. H. et al.) 173–188 (Oxford University Press, Oxford, 2012).
    Google Scholar 

    46.
    Somera, A. F., Lima, A. M., Santos-Neto, A. J., Lanças, F. M. & Bacci, M. Jr. Leaf-cutter ant fungus gardens are biphasic mixed microbial bioreactors that convert plant biomass to polyols with biotechnological applications. Appl. Environ. Microbiol. 81, 4525–4535. https://doi.org/10.1128/AEM.00046-15 (2015).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    47.
    Ballard, R. W., Palleroni, N. J., Doudoroff, M., Stanier, R. Y. & Mandel, M. Taxonomy of the aerobic pseudomonads: Pseudomonas cepacia, P. marginata, P. alliicola and P. caryophylli. J. Gen. Microbiol. 60, 199–214. https://doi.org/10.1099/00221287-60-2-199 (1970).
    Article  PubMed  CAS  Google Scholar 

    48.
    O’Hara, C. M. Manual and automated instrumentation for identification of Enterobacteriaceae and other aerobic gram-negative Bacilli. Clin. Microbiol. Rev. 18, 147–162. https://doi.org/10.1128/CMR.18.1.147-162.2005 (2005).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    49.
    Brune, A., Miambi, E. & Breznak, J. A. Roles of oxygen and the intestinal microflora in the metabolism of lignin-derived phenylpropanoids and other monoaromatic compounds by termites. Appl. Environ. Microbiol. 61, 2688–2695 (1995).
    Article  CAS  Google Scholar 

    50.
    White, B. A., Lamed, R., Bayer, E. A. & Flint, H. J. Biomass utilization by gut microbiomes. Annu. Rev. Microbiol. 68, 279–296. https://doi.org/10.1146/annurev-micro-092412-155618 (2014).
    Article  PubMed  CAS  Google Scholar 

    51.
    de Vos, W. Microbial biofilms and the human intestinal microbiome. npj Biofilms Microbio. 1, 15005. https://doi.org/10.1038/npjbiofilms.2015.5 (2015).
    Article  CAS  Google Scholar 

    52.
    Koh, A., De Vadder, F., Kovatcheva-Datchary, P. & Bäckhed, F. From dietary fiber to host physiology: short-chain fatty acids as key bacterial metabolites. Cell 165, 1332–1345. https://doi.org/10.1016/j.cell.2016.05.041 (2016).
    Article  PubMed  CAS  Google Scholar 

    53.
    Leng, R. A. Biofilm compartmentalisation of the rumen microbiome: modification of fermentation and degradation of dietary toxins. Anim. Prod. Sci. 57, 2188–2203. https://doi.org/10.1071/AN17382 (2017).
    Article  CAS  Google Scholar 

    54.
    Kohl, K. D. et al. Metagenomic sequencing provides insights into microbial detoxification in the guts of small mammalian herbivores (Neotoma spp.). FEMS Microbiol. Ecol. 94, fiy184. https://doi.org/10.1093/femsec/fiy184 (2018).
    Article  CAS  Google Scholar 

    55.
    Burke, C., Steinberg, P., Rusch, D., Kjelleberg, S. & Thomas, T. Bacterial community assembly based on functional genes rather than species. Proc. Natl. Acad. Sci USA 108, 14288–14293. https://doi.org/10.1073/pnas.1101591108 (2011).
    ADS  Article  Google Scholar 

    56.
    Louca, S., Parfrey, L. W. & Doebeli, M. Decoupling function and taxonomy in the global ocean microbiome. Science 353, 1272–1277. https://doi.org/10.1126/science.aaf4507 (2016).
    ADS  Article  PubMed  CAS  Google Scholar 

    57.
    Louca, S. Function and functional redundancy in microbial systems. Nat. Ecol. Evol. 2, 936–943. https://doi.org/10.1038/s41559-018-0519-1 (2018).
    Article  PubMed  Google Scholar 

    58.
    Jurburg, S. D. & Salles, J. F. Functional redundancy and ecosystem function—the soil microbiota as a case study. In Biodiversity in ecosystems—linking structure and function (eds Lo, Y.-H. et al.) 29–49 (INTECH, New York, 2015).
    Google Scholar 

    59.
    Grell, M. N. et al. The fungal symbiont of Acromyrmex leaf-cutting ants expresses the full spectrum of genes to degrade cellulose and other plant cell wall polysaccharides. BMC Genomics 14, 928. https://doi.org/10.1186/1471-2164-14-928 (2013).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    60.
    Žifčáková, L. et al. Feed in summer, rest in winter: microbial carbon utilization in forest topsoil. Microbiome 5, 122. https://doi.org/10.1186/s40168-017-0340-0 (2017).
    Article  PubMed  PubMed Central  Google Scholar 

    61.
    Jing, T., Qi, F. & Wang, Z. Most dominant roles of insect gut bacteria: digestion, detoxification, or essential nutrient provision? Microbiome 8, 38. https://doi.org/10.1186/s40168-020-00823-y (2020).
    Article  PubMed  PubMed Central  Google Scholar 

    62.
    Howard, J. J., Cazin, J. & Wiemer, D. F. Toxicity of terpenoid deterrents to the leafcutting ant Atta cephalotes and its mutualistic fungus. J. Chem. Ecol. 14, 59–69. https://doi.org/10.1007/BF01022531 (1988).
    Article  PubMed  CAS  Google Scholar 

    63.
    Keeling, C. I. & Bohlmann, J. Diterpene resin acids in conifers. Phytochemistry 67, 2415–2423. https://doi.org/10.1016/j.phytochem.2006.08.019 (2006).
    Article  PubMed  CAS  Google Scholar 

    64.
    Zhu, L. et al. Potential mechanism of detoxification of cyanide compounds by gut microbiomes of bamboo-eating pandas. MSphere 3, e00229-18. https://doi.org/10.1128/mSphere.00229-18 (2018).
    Article  PubMed  PubMed Central  Google Scholar 

    65.
    Cheng, X. et al. Metagenomic analysis of the pinewood nematode microbiome reveals a symbiotic relationship critical for xenobiotics degradation. Sci. Rep. 3, 1869. https://doi.org/10.1038/srep01869 (2013).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    66.
    Flemming, H. et al. Biofilms: an emergent form of bacterial life. Nat. Rev. Microbiol. 14, 563–575. https://doi.org/10.1038/nrmicro.2016.94 (2016).
    Article  PubMed  CAS  Google Scholar 

    67.
    Sivadon, P., Barnier, C., Urios, L. & Grimaud, R. Biofilm formation as a microbial strategy to assimilate particulate substrates. Environ. Microbiol. Rep. 11, 749–764. https://doi.org/10.1111/1758-2229.12785 (2019).
    Article  PubMed  CAS  Google Scholar 

    68.
    Brethauer, S., Shahab, R. L. & Studer, M. H. Impacts of biofilms on the conversion of cellulose. Appl. Microbiol. Biotechnol. 104, 5201–5212. https://doi.org/10.1007/s00253-020-10595-y (2020).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    69.
    Macfarlane, S. & Macfarlane, G. T. Composition and metabolic activities of bacterial biofilms colonizing food residues in the human gut. Appl. Environ. Microbiol. 72, 6204–6211. https://doi.org/10.1128/AEM.00754-06 (2006).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    70.
    Deveau, A. et al. Bacterial–fungal interactions: ecology, mechanisms and challenges. FEMS Microbiol. Rev. 42, 335–352. https://doi.org/10.1093/femsre/fuy008 (2018).
    Article  PubMed  CAS  Google Scholar 

    71.
    Purahong, W. et al. Life in leaf litter: novel insights into community dynamics of bacteria and fungi during litter decomposition. Mol. Ecol. 25, 4059–4074. https://doi.org/10.1111/mec.13739 (2016).
    Article  PubMed  CAS  Google Scholar 

    72.
    Frey-Klett, P. et al. Bacterial-fungal interactions: hyphens between agricultural, clinical, environmental, and food microbiologists. Microbiol. Mol. Biol. Rev. 75, 583–609. https://doi.org/10.1128/MMBR.00020-11 (2011).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    73.
    Martin, M. M. Biochemical implications of insect mycophagy. Biol. Rev. 54, 1–21. https://doi.org/10.1111/j.1469-185X.1979.tb00865.x (1979).
    Article  CAS  Google Scholar 

    74.
    Brabcová, V., Nováková, M., Davidová, A. & Baldrian, P. Dead fungal mycelium in forest soil represents a decomposition hotspot and a habitat for a specific microbial community. New Phytol. 210, 1369–1381. https://doi.org/10.1111/nph.13849 (2016).
    Article  PubMed  CAS  Google Scholar 

    75.
    Brabcová, V., Štursová, M. & Baldrian, P. Nutrient content affects the turnover of fungal biomass in forest topsoil and the composition of associated microbial communities. Soil Biol. Biochem. 118, 187–198. https://doi.org/10.1016/j.soilbio.2017.12.012 (2018).
    Article  CAS  Google Scholar 

    76.
    de Boer, W. D., Folman, L. B., Summerbell, R. C. & Boddy, L. Living in a fungal world: impact of fungi on soil bacterial niche development. FEMS Microbiol. Rev. 29, 795–811. https://doi.org/10.1016/j.femsre.2004.11.005 (2005).
    Article  PubMed  CAS  Google Scholar 

    77.
    Leveau, J. H. & Preston, G. M. Bacterial mycophagy: definition and diagnosis of a unique bacterial–fungal interaction. New Phytol. 177, 859–876. https://doi.org/10.1111/j.1469-8137.2007.02325.x (2008).
    Article  PubMed  Google Scholar 

    78.
    Carrasco, J. & Preston, G. M. Growing edible mushrooms: a conversation between bacteria and fungi. Environ. Microbiol. 22, 858–872. https://doi.org/10.1111/1462-2920.14765 (2020).
    Article  PubMed  Google Scholar 

    79.
    Warmink, J. A., Nazir, R. & Van Elsas, J. D. Universal and species-specific bacterial ‘fungiphiles’ in the mycospheres of different basidiomycetous fungi. Environ. Microbiol. 11, 300–312. https://doi.org/10.1111/j.1462-2920.2008.01767.x (2009).
    Article  PubMed  CAS  Google Scholar 

    80.
    Guennoc, C., Rose, C., Labbé, J. & Deveau, A. Bacterial biofilm formation on the hyphae of ectomycorrhizal fungi: a widespread ability under controls?. FEMS Microbiol. Ecol. 94, 093. https://doi.org/10.1093/femsec/fiy093 (2018).
    Article  CAS  Google Scholar 

    81.
    Figueiredo, A. R. T. D. & Kramer, J. Cooperation and conflict within the microbiota and their effects on animal hosts. Front. Ecol. Evol. 8, 132. https://doi.org/10.3389/fevo.2020.00132 (2020).
    Article  Google Scholar 

    82.
    Coyte, K. Z., Schluter, J. & Foster, K. R. The ecology of the microbiome: networks, competition, and stability. Science 350, 663–666. https://doi.org/10.1126/science.aad2602 (2015).
    ADS  Article  PubMed  CAS  Google Scholar 

    83.
    Donaldson, G., Lee, S. & Mazmanian, S. Gut biogeography of the bacterial microbiota. Nat. Rev. Microbiol. 14, 20–32. https://doi.org/10.1038/nrmicro3552 (2016).
    Article  PubMed  CAS  Google Scholar 

    84.
    Tropini, C., Earle, K. A., Huang, K. C. & Sonnenburg, J. L. The gut microbiome: connecting spatial organization to function. Cell Host Microbe 21, 433–442. https://doi.org/10.1016/j.chom.2017.03.010 (2017).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    85.
    Adair, K. L. & Douglas, A. E. Making a microbiome: the many determinants of host-associated microbial community composition. Curr. Opin. Microbiol. 35, 23–29. https://doi.org/10.1016/j.mib.2016.11.002 (2017).
    Article  PubMed  Google Scholar 

    86.
    Shafquat, A., Joice, R., Simmons, S. & Huttenhower, C. Functional and phylogenetic assembly of microbial communities in the human microbiome. Trends Microbiol. 22, 261–266. https://doi.org/10.1016/j.tim.2014.01.011 (2014).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    87.
    Hernandez-Agreda, A., Gates, R. D. & Ainsworth, T. D. Defining the core microbiome in corals’ microbial soup. Trends Microbiol. 25, 125–140. https://doi.org/10.1016/j.tim.2016.11.003 (2017).
    Article  PubMed  CAS  Google Scholar 

    88.
    Ramadhar, T. et al. Bacterial symbionts in agricultural systems provide a strategic source for antibiotic discovery. J. Antibiot. 67, 53–58. https://doi.org/10.1038/ja.2013.77 (2014).
    Article  PubMed  CAS  Google Scholar 

    89.
    Van Arnam, E. B., Currie, C. R. & Clardy, J. Defense contracts: molecular protection in insect-microbe symbioses. Chem. Soc. Rev. 47, 1638–1651. https://doi.org/10.1039/C7CS00340D (2018).
    Article  PubMed  Google Scholar 

    90.
    Berasategui, A. et al. Potential applications of insect symbionts in biotechnology. Appl. Microbiol. Biotechnol. 100, 1567–1577. https://doi.org/10.1007/s00253-015-7186-9 (2016).
    Article  PubMed  CAS  Google Scholar 

    91.
    Cox, M. P., Peterson, D. A. & Biggs, P. J. SolexaQA: At-a-glance quality assessment of Illumina second-generation sequencing data. BMC Bioinformatics 11, 485. https://doi.org/10.1186/1471-2105-11-485 (2010).
    Article  PubMed  PubMed Central  Google Scholar 

    92.
    Li, D., Liu, C., Luo, R., Sadakane, K. & Lam, T.-W. MEGAHIT: an ultra-fast single-node solution for large and complex metagenomics assembly via succinct de Bruijn graph. Bioinformatics 31, 1674–1676. https://doi.org/10.1093/bioinformatics/btv033 (2015).
    Article  PubMed  CAS  Google Scholar 

    93.
    Schmieder, R. & Edwards, R. Quality control and preprocessing of metagenomic datasets. Bioinformatics 27, 863–864. https://doi.org/10.1093/bioinformatics/btr026 (2011).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    94.
    Markowitz, V. M. et al. IMG/M 4 version of the integrated metagenome comparative analysis system. Nucleic Acids Res. 42, D568–D573. https://doi.org/10.1093/nar/gkt919 (2014).
    Article  PubMed  CAS  Google Scholar 

    95.
    Patil, K. R., Roune, L. & MChardy, A. C. The PhyloPythiaS web server for taxonomic assignment of metagenome sequences. PLoS One 7, e38581. https://doi.org/10.1371/journal.pone.0038581 (2012).
    ADS  Article  PubMed  PubMed Central  CAS  Google Scholar 

    96.
    Engel, P., Martinson, V. G. & Moran, N. A. Functional diversity within the simple gut microbiota of the honey bee. Proc. Natl Acad. Sci. USA 109, 11002–11007. https://doi.org/10.1073/pnas.1202970109 (2012).
    ADS  Article  PubMed  Google Scholar 

    97.
    Kešnerová, L., Moritz, R. & Engel, P. Bartonella apis sp. Nov., a honey bee gut symbiont of the class Alphaproteobacteria. Int. J. Syst. Evol. Microbiol. 66, 414–421. https://doi.org/10.1099/ijsem.0.000736 (2016).
    Article  CAS  Google Scholar 

    98.
    Edgar, R. C. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 32, 1792–1797. https://doi.org/10.1093/nar/gkh340 (2004).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    99.
    Guindon, S. & Gascuel, O. A Simple, fast, and accurate algorithm to estimate large phylogenies by maximum likelihood. Syst. Biol. 52, 696–704. https://doi.org/10.1080/10635150390235520https://doi.org/10.1080/10635150390235520 (2003).
    Article  PubMed  Google Scholar 

    100.
    Hammer, Ř., Harper, D. A. T. & Ryan, P. D. PAST: Paleontological statistics software package for education and data analysis. Palaeontol. Electronica 4, 1. https://palaeo-electronica.org/2001_1/past/issue1_01.htm (2001).

    101.
    Parks, D. H., Tyson, G. W., Hugenholtz, P. & Beiko, R. G. STAMP: Statistical analysis of taxonomic and functional profiles. Bioinformatics 30, 3123–3124. https://doi.org/10.1093/bioinformatics/btu494https://doi.org/10.1093/bioinformatics/btu494 (2014).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    102.
    Fan, H., Ives, A. R., Surget-Groba, Y. & Cannon, C. H. An assembly and alignment-free method of phylogeny reconstruction from next-generation sequencing data. BMC Genomics 16, 522. https://doi.org/10.1186/s12864-015-1647-5 (2015).
    Article  PubMed  PubMed Central  CAS  Google Scholar 

    103.
    Letunic, I. & Bork, P. Interactive tree of life (iTOL): an online tool for phylogenetic tree display and annotation. Bioinformatics 23, 127–128. https://doi.org/10.1093/bioinformatics/btl529 (2007).
    Article  PubMed  CAS  Google Scholar 

    104.
    Kanehisa, M., Goto, S., Sato, Y., Furumichi, M. & Tanabe, M. KEGG for integration and interpretation of large scale molecular data sets. Nucleic Acids Res. 40, D109–D114. https://doi.org/10.1093/nar/gkr988 (2012).
    Article  PubMed  CAS  Google Scholar 

    105.
    White, J. R. et al. Statistical methods for detecting differentially abundant features in clinical metagenomic samples. PLoS Comput. Biol. 5, 1000352. https://doi.org/10.1371/journal.pcbi.1000352 (2009).
    Article  CAS  Google Scholar 

    106.
    Cantarel, B. L. et al. The carbohydrate-active enZymes database (CAZy): an expert resource for glycogenomics. Nucleic Acids Res. 37, D233–D238. https://doi.org/10.1093/nar/gkn663 (2009).
    Article  PubMed  CAS  Google Scholar 

    107.
    Zhang, H. et al. dbCAN2: a meta server for automated carbohydrate-active enzyme annotation. Nucleic Acids Res. 46, W95–W101. https://doi.org/10.1093/nar/gky418 (2018).
    ADS  Article  PubMed  PubMed Central  CAS  Google Scholar 

    108.
    Kanehisa, M., Sato, Y. & Morishima, K. BlastKOALA & GhostKOALA: KEGG tools for functional characterization of genome and metagenome sequences. J. Mol. Biol. 428, 726–731. https://doi.org/10.1016/j.jmb.2015.11.006 (2016).
    Article  PubMed  CAS  Google Scholar 

    109.
    Barcoto, M. O. Fungus-growing insects host a convergent microbiome with functional similarities to other lignocellulose-feeding insects. Masters dissertation, São Paulo State University (2017). More

  • in

    A marine virus as foe and friend

    1.
    Zhao, Y. et al. Nature 494, 357–360 (2013).
    CAS  Article  Google Scholar 
    2.
    Giovannoni, S. J. Annu. Rev. Mar. Sci. 9, 231–255 (2017).
    Article  Google Scholar 

    3.
    Morris, R. M., Cain, K. R., Hvorecny, K. L. & Kollman, J. M. Nat. Microbiol. https://doi.org/10.1038/s41564-020-0725-x (2020).
    Article  PubMed  Google Scholar 

    4.
    Zhao, Y. et al. Environ. Microbiol. 21, 1989–2001 (2019).
    CAS  Article  Google Scholar 

    5.
    Howard-Varona, C., Hargreaves, K. R., Abedon, S. T. & Sullivan, M. B. ISME J. 11, 1511–1520 (2017).
    Article  Google Scholar 

    6.
    Nanda, A. M., Thormann, K. & Frunzke, J. J. Bacteriol. 197, 410–419 (2015).
    Article  Google Scholar 

    7.
    Liu, X. et al. Environ. Microbiol. 21, 4212–4232 (2019).
    CAS  Article  Google Scholar 

    8.
    Feiner, R. et al. Nat. Rev. Microbiol. 13, 641–650 (2015).
    CAS  Article  Google Scholar 

    9.
    Våge, S., Storesund, J. E. & Thingstad, T. F. Nature 499, E3–E4 (2013).
    Article  Google Scholar  More

  • in

    Selectivity of deltamethrin doses on Palmistichus elaeisis (Hymenoptera: Eulophidae) parasitizing Tenebrio molitor (Coleoptera: Tenebrionidae)

    1.
    Zanuncio, J. C. et al. Population dynamics of Lepidoptera pests in Eucalyptus urophylla plantations in the Brazilian Amazonia. Forests 5, 72–87 (2014).
    Google Scholar 
    2.
    Macedo-Reis, L., Soares, L., Faria, M., Espírito-Santo, M. & Zanuncio, J. Survival of a lepidopteran defoliator of Eucalyptus is influenced by local hillside and forest remnants in Brazil. Fla. Entomol. 96, 941–947 (2013).
    Google Scholar 

    3.
    Loetti, V. & Bellocq, I. Effects of the insecticides methoxyfenozide and cypermethrin on non-target arthropods: a field experiment. Austral Entomol. 56, 255–260 (2017).
    Google Scholar 

    4.
    Lundström, N. L. P., Zhang, H. & Brännström, Å. Pareto-efficient biological pest control enable high efficacy at small costs. Ecol. Model. 364, 89–97 (2017).
    Google Scholar 

    5.
    Costa, L. G., Giordano, G., Guizzetti, M. & Vitalone, A. Neurotoxicity of pesticides: a brief review. Front. Biosci. 13, 1240–1249 (2008).
    CAS  PubMed  Google Scholar 

    6.
    de S Pereira, K., Guedes, N. M. P., Serrão, J. E., Zanuncio, J. C. & Guedes, R. N. C. Superparasitism, immune response and optimum progeny yield in the gregarious parasitoid Palmistichus elaeisis. Pest Manag. Sci. 73, 1101–1109 (2017).
    PubMed  Google Scholar 

    7.
    Pereira, F. F., Zanuncio, T. V., Zanuncio, J. C., Pratissoli, D. & Tavares, M. T. Species of lepidoptera defoliators of eucalyptus as new host for the parasitoid Palmistichus elaeisis (Hymenoptera: Eulophidae). Braz. Arch. Biol. Technol. 51, 259–262 (2008).
    Google Scholar 

    8.
    Barbosa, R. H., Zanuncio, J. C., Pereira, F. F., Kassab, S. O. & Rossoni, C. Foraging activity of Palmistichus elaeisis (Hymenoptera: Eulophidae) at various densities on pupae of the Eucalyptus defoliator Thyrinteina arnobia (Lepidoptera: Geometridae). Fla. Entomol. 99, 686–690 (2016).
    Google Scholar 

    9.
    Zanuncio, J., Pereira, F., Jacques, G., Tavares, M. & Serrão, J. Tenebrio molitor Linnaeus (Coleoptera: Tenebrionidae), a new alternative host to rear the pupae parasitoid Palmistichus elaeisis Delvare & Lasalle (Hymenoptera: Eulophidae). Coleopt. Bull. 62, 64–66 (2008).
    Google Scholar 

    10.
    Roubos, C. R., Rodriguez-Saona, C. & Isaacs, R. Mitigating the effects of insecticides on arthropod biological control at field and landscape scales. Biol. Control 75, 28–38 (2014).
    CAS  Google Scholar 

    11.
    Addison, P. J. & Barker, G. M. Effect of various pesticides on the non-target species Microctonus hyperodae, a biological control agent of Listronotus bonariensis. Entomol. Exp. Appl. 119, 71–79 (2006).
    CAS  Google Scholar 

    12.
    Banks, J. E., Stark, J. D., Vargas, R. I. & Ackleh, A. S. Parasitoids and ecological risk assessment: can toxicity data developed for one species be used to protect an entire guild?. Biol. Control 59, 336–339 (2011).
    Google Scholar 

    13.
    Bayram, A., Salerno, G., Onofri, A. & Conti, E. Lethal and sublethal effects of preimaginal treatments with two pyrethroids on the life history of the egg parasitoid Telenomus busseolae. Biocontrol 55, 697–710 (2010).
    CAS  Google Scholar 

    14.
    Desneux, N., Decourtye, A. & Delpuech, J.-M. The sublethal effects of pesticides on beneficial arthropods. Annu. Rev. Entomol. 52, 81–106 (2007).
    CAS  PubMed  Google Scholar 

    15.
    Delpuech, J. M. & Delahaye, M. The sublethal effects of deltamethrin on Trichogramma behaviors during the exploitation of host patches. Sci. Total Environ. 447, 274–279 (2013).
    ADS  CAS  PubMed  Google Scholar 

    16.
    De La Cruz, R. A. et al. Side-effects of pesticides on the generalist endoparasitoid Palmistichus elaeisis (Hymenoptera: Eulophidae). Sci. Rep. 7, 1–8 (2017).
    CAS  Google Scholar 

    17.
    Wang, Y. et al. Susceptibility to selected insecticides and risk assessment in the insect egg parasitoid Trichogramma confusum (Hymenoptera: Trichogrammatidae). J. Econ. Entomol. 106, 142–149 (2013).
    CAS  PubMed  Google Scholar 

    18.
    Zanuncio, T. V. et al. Fertility and life expectancy of the predator Supputius cincticeps (Heteroptera: Pentatomidae) exposed to sublethal doses of permethrin. Biol. Res. 38, 31–39 (2005).
    CAS  PubMed  Google Scholar 

    19.
    Wang, Y. et al. Toxicity risk of insecticides to the insect egg parasitoid Trichogramma evanescens Westwood (Hymenoptera: Trichogrammatidae). Pest Manag. Sci. 70, 398–404 (2014).
    CAS  PubMed  Google Scholar 

    20.
    Biondi, A., Zappalà, L., Stark, J. D. & Desneux, N. Do biopesticides affect the demographic traits of a parasitoid wasp and its biocontrol services through sublethal effects?. PLoS ONE 8, 1–11 (2013).
    Google Scholar 

    21.
    Cônsoli, F. L., Botelho, P. S. M. & Parra, J. R. P. Selectivity of insecticides to the egg parasitoid Trichogramma galloi Zucchi, 1988, (Hym., Trichogrammatidae). J. Appl. Entomol. 125, 37–43 (2001).
    Google Scholar 

    22.
    Soderlund, D. M. Toxicology and mode of action of pyrethroid insecticides. In Hayes’ Handbook of Pesticide Toxicology 1665–1686 (Elsevier Inc., 2010). doi:10.1016/B978-0-12-374367-1.00077-X.

    23.
    Youssef, A. I. et al. The side-effects of plant protection products used in olive cultivation on the hymenopterous egg parasitoid Trichogramma cacoeciae Marchal. J. Appl. Entomol. 128, 593–599 (2004).
    CAS  Google Scholar 

    24.
    Alix, A., Cortesero, A. M., Nénon, J. P. & Anger, J. P. Selectivity assessment of chlorfenvinphos reevaluated by including physiological and behavioral effects on an important beneficial insect. Environ. Toxicol. Chem. 20, 2530–2536 (2001).
    CAS  PubMed  Google Scholar 

    25.
    Fontes, J., Roja, I. S., Tavares, J. & Oliveira, L. Lethal and sublethal effects of various pesticides on Trichogramma achaeae (Hymenoptera: Trichogrammatidae). J. Econ. Entomol. 111, 1219–1226 (2018).
    CAS  PubMed  Google Scholar 

    26.
    Bayram, A., Salerno, G., Onofri, A. & Conti, E. Sub-lethal effects of two pyrethroids on biological parameters and behavioral responses to host cues in the egg parasitoid Telenomus busseolae. Biol. Control 53, 153–160 (2010).
    CAS  Google Scholar 

    27.
    Zantedeschi, R. et al. Toxicity of soybean-registered agrochemicals to Telenomus podisi and Trissolcus basalis immature stages. Phytoparasitica 46, 203–212 (2018).
    CAS  Google Scholar 

    28.
    Thubru, D. P., Firake, D. M. & Behere, G. T. Assessing risks of pesticides targeting lepidopteran pests in cruciferous ecosystems to eggs parasitoid, Trichogramma brassicae (Bezdenko). Saudi J. Biol. Sci. 25, 680–688 (2018).
    CAS  PubMed  Google Scholar 

    29.
    Fernández, M. D. M., Medina, P., Fereres, A., Smagghe, G. & Viñuela, E. Are mummies and adults of Eretmocerus mundus (Hymenoptera: Aphelinidae) compatible with modern insecticides?. J. Econ. Entomol. 108, 2268–2277 (2015).
    Google Scholar 

    30.
    Liu, F., Zhang, X., Gui, Q. Q. & Xu, Q. J. Sublethal effects of four insecticides on Anagrus nilaparvatae (Hymenoptera: Mymaridae), an important egg parasitoid of the rice planthopper Nilaparvata lugens (Homoptera: Delphacidae). Crop Prot. 37, 13–19 (2012).
    Google Scholar 

    31.
    Pastori, P. L. et al. Reproduction of Trichospilus diatraeae (Hymenoptera: Eulophidae) in pupae of two lepidopterans defoliators of eucalypt. Rev. Colomb. Entomol. 38, 91–93 (2012).
    Google Scholar 

    32.
    Harvey, J. A. & Gols, R. Effects of plant-mediated differences in host quality on the development of two related endoparasitoids with different host-utilization strategies. J. Insect Physiol. 107, 110–115 (2018).
    CAS  PubMed  Google Scholar 

    33.
    Pereira, F. F. et al. The density of females of Palmistichus elaeisis Delvare and Lasalle (Hymenoptera: Eulophidae) affects their reproductive performance on pupae of Bombyx mori L. (Lepidoptera: Bombycidae). An. Acad. Bras. Cienc. 82, 323–331 (2010).
    PubMed  Google Scholar 

    34.
    Desneux, N., Denoyelle, R. & Kaiser, L. A multi-step bioassay to assess the effect of the deltamethrin on the parasitic wasp Aphidius ervi. Chemosphere 65, 1697–1706 (2006).
    ADS  CAS  PubMed  Google Scholar 

    35.
    Desneux, N., Ramirez-Romero, R. & Kaiser, L. Multistep bioassay to predict recolonization potential of emerging parasitoids after a pesticide treatment. Environ. Toxicol. Chem. 25, 2675–2682 (2006).
    CAS  PubMed  Google Scholar 

    36.
    Agrolink. Bula Decis 25 EC. www.agrolink.com.br https://www.agrolink.com.br/agrolinkfito/produto/decis-25-ec_8.html (2018).

    37.
    Khan, M. A., Khan, H. & Ruberson, J. R. Lethal and behavioral effects of selected novel pesticides on adults of Trichogramma pretiosum (Hymenoptera: Trichogrammatidae). Pest Manag. Sci. 71, 1640–1648 (2015).
    CAS  PubMed  Google Scholar 

    38.
    Stark, J. D., Banks, J. E. & Vargas, R. How risky is risk assessment: the role that life history strategies play in susceptibility of species to stress. Proc. Natl. Acad. Sci. 101, 732–736 (2004).
    ADS  CAS  PubMed  Google Scholar 

    39.
    Rieth, J. P. & Levin, M. D. The repellent effect of two pyrethroid insecticides on the honey bee. Physiol. Entomol. 13, 213–218 (1988).
    CAS  Google Scholar 

    40.
    Valente, E. C. N., Broglio, S. M. F., dos Passos, E. M. & de Lima, A. S. T. Desempenho de Trichogramma galloi (Hymenoptera: Trichogrammatidae) sobre ovos de Diatraea spp. (Lepidoptera: Crambidae). Pesqui. Agropecu. Bras. 51, 293–300 (2016).
    Google Scholar 

    41.
    Rossoni, C. et al. Development of Eulophidae (Hymenoptera) parasitoids in Diatraea saccharalis (Lepidoptera: Crambidae) pupae exposed to entomopathogenic fungi. Can. Entomol. 148, 716–723 (2016).
    Google Scholar 

    42.
    Harvey, J. A., Poelman, E. H. & Tanaka, T. Intrinsic inter- and intraspecific competition in parasitoid wasps. Annu. Rev. Entomol. 58, 333–351 (2013).
    CAS  PubMed  Google Scholar 

    43.
    Jervis, M. A., Ellers, J. & Harvey, J. A. Resource acquisition, allocation, and utilization in parasitoid reproductive strategies. Annu. Rev. Entomol. 53, 361–385 (2008).
    CAS  PubMed  Google Scholar 

    44.
    Souza, D., Monteiro, A. B. & Faria, L. D. B. Morphometry, allometry, and fluctuating asymmetry of egg parasitoid Trichogramma pretiosum under insecticide influence. Entomol. Exp. Appl. 166, 298–303 (2018).
    CAS  Google Scholar 

    45.
    Rasool, S. et al. Effect of host size on larval competition of the gregarious parasitoid Bracon hebetor (Say.) (Hymenoptera: Braconidae). Pak. J. Zool. 49, 1085–1085 (2017).
    Google Scholar 

    46.
    Menezes, C. W. G. et al. Reproductive and toxicological impacts of herbicides used in Eucalyptus culture in Brazil on the parasitoid Palmistichus elaeisis (Hymenoptera: Eulophidae). Weed Res. 52, 520–525 (2012).
    CAS  Google Scholar 

    47.
    Andrade, G. S. et al. Oogenesis pattern and type of ovariole of the parasitoid Palmistichus elaeisis (Hymenoptera: Eulophidae). An. Acad. Bras. Cienc. 84, 767–774 (2012).
    PubMed  Google Scholar 

    48.
    Menezes, C. W. G. et al. Palmistichus elaeisis (Hymenoptera: Eulophidae) as an indicator of toxicity of herbicides registered for corn in Brazil. Chil. J. Agric. Res. 74, 361–365 (2014).
    Google Scholar 

    49.
    I. R. A. C. Method No: 007: Leaf eating Lepidoptera and Coleoptera. Available online: uploads/ 2009, (2010).

    50.
    Finney, D. J. P. A. Probit Analysis 318 (Cambridge University Press, London , 1971).
    Google Scholar  More

  • in

    Savanna tree evolutionary ages inform the reconstruction of the paleoenvironment of our hominin ancestors

    1.
    Domínguez-Rodrigo, M. Is the “Savanna Hypothesis” a dead concept for explaining the emergence of the earliest hominins?. Curr. Anthropol. 55, 59–81 (2014).
    Google Scholar 
    2.
    Cerling, T. E. et al. Woody cover and hominin environments in the past 6 million years. Nature 476, 52–56 (2011).
    ADS  Google Scholar 

    3.
    Potts, R. Hominin evolution in settings of strong environmental variability. Q. Sci. Rev. 73, 1–13 (2013).
    ADS  Google Scholar 

    4.
    Magill, C. R., Ashley, G. M. & Freeman, K. H. Ecosystem variability and early human habitats in eastern Africa. Proc. Natl Acad. Sci. USA 110, 1167–1174 (2013).
    ADS  CAS  PubMed  Google Scholar 

    5.
    Levin, N. E. Environment and climate of early human evolution. Annu. Rev. Earth Planet. Sci. 43, 405–429 (2015).
    ADS  CAS  Google Scholar 

    6.
    Uno, K. T., Polissar, P. J., Jackson, K. E. & deMenocal, P. B. Neogene biomarker record of vegetation change in eastern Africa. Proc. Natl Acad. Sci. USA 113, 6355–6363 (2016).
    ADS  CAS  PubMed  Google Scholar 

    7.
    Jacobs, B. F. Paleobotanical studies from tropical Africa: relevance to the evolution of forest, woodland, and savannah biomes. Phil. Trans. R. Soc. B 359, 1573–1583 (2004).
    PubMed  Google Scholar 

    8.
    Beerling, D. J. & Osborne, C. P. The origin of the savanna biome. Glob. Chang. Biol. 12, 2023–2031 (2006).
    ADS  Google Scholar 

    9.
    Bond, W. J. What limits trees in C4 grasslands and savannas?. Annu. Rev. Ecol. Evol. Syst. 39, 641–659 (2008).
    Google Scholar 

    10.
    Cerling, T. E. et al. Global vegetation change through the Miocene/Pliocene boundary. Nature 389, 153–158 (1997).
    ADS  CAS  Google Scholar 

    11.
    Ehleringer, J. R., Cerling, T. E. & Helliker, B. R. C4 photosynthesis, atmospheric CO2, and climate. Oecologia 112, 285–299 (1997).
    ADS  PubMed  Google Scholar 

    12.
    Beerling, D. J. & Royer, D. L. Convergent cenozoic CO2 history. Nat. Geosci. 4, 418–420 (2011).
    ADS  CAS  Google Scholar 

    13.
    Pagani, M., Zachos, J. C., Freeman, K. H., Tipple, B. & Bohaty, S. Marked decline in atmospheric carbon dioxide concentrations during the Paleogene. Science 309, 600–603 (2005).
    ADS  CAS  PubMed  Google Scholar 

    14.
    Pagani, M., Freeman, K. H. & Arthur, M. A. Late Miocene atmospheric CO2 concentrations and the expansion of C4 grasses. Science 285, 876–879 (1999).
    CAS  PubMed  Google Scholar 

    15.
    Bolton, C. T. et al. Decrease in coccolithophore calcification and CO2 since the middle Miocene. Nat. Commun. 7, 10284 (2016).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    16.
    Herbert, T. D. et al. Late Miocene global cooling and the rise of modern ecosystems. Nat. Geosci. 9, 843–847 (2016).
    ADS  CAS  Google Scholar 

    17.
    Sponheimer, M. et al. Isotopic evidence of early hominin diets. Proc. Natl Acad. Sci. USA 110, 10513–10518 (2013).
    ADS  CAS  Google Scholar 

    18.
    Feakins, S. J. et al. Northeast African vegetation change over 12 my. Geology 41, 295–298 (2013).
    ADS  Google Scholar 

    19.
    Ségalen, L., Lee-Thorp, J. A. & Cerling, T. E. Timing of C4 grass expansion across sub-Saharan Africa. J. Hum. Evol. 53, 549–559 (2007).
    PubMed  Google Scholar 

    20.
    Pennington, R. T., Cronk, Q. C. B. & Richardson, J. A. Introduction and synthesis: plant phylogeny and the origin of major biomes. Phil. Trans. R. Soc. B. 359, 1455–1464 (2004).
    PubMed  Google Scholar 

    21.
    Pennington, R. T., Richardson, J. E. & Lavin, M. Insights into the historical construction of species-rich biomes from dated plant phylogenies, neutral ecological theory and phylogenetic community structure. New Phytol. 172, 605–616 (2006).
    PubMed  Google Scholar 

    22.
    Bytebier, B., Antonelli, A., Bellstedt, D. U. & Linder, H. P. Estimating the age of fire in the Cape flora of South Africa from an orchid phylogeny. Proc. R. Soc. B 278, 188–195 (2010).
    PubMed  Google Scholar 

    23.
    Crisp, M. D., Burrows, G. E., Cook, L. G., Thornhill, A. H. & Bowman, D. M. Flammable biomes dominated by eucalypts originated at the Cretaceous-Palaeogene boundary. Nat. Comm. 2, 193 (2011).
    ADS  Google Scholar 

    24.
    Vicentini, A., Barber, J. C., Aliscioni, S. S., Giussani, L. M. & Kellogg, E. A. The age of the grasses and clusters of origins of C4 photosynthesis. Glob. Chang. Biol. 14, 2963–2977 (2008).
    ADS  Google Scholar 

    25.
    Scheiter, S. et al. Fire and fire-adapted vegetation promoted C4 expansion in the Late Miocene. New Phytol. 195, 653–666 (2012).
    PubMed  Google Scholar 

    26.
    Ramírez, S. R., Gravendeel, B., Singer, R. B., Marshall, C. R. & Pierce, N. E. Dating the origin of the Orchidaceae from a fossil orchid with its pollinator. Nature 448, 1042–1045 (2007).
    ADS  Google Scholar 

    27.
    Losos, J. B. & Schluter, D. Analysis of an evolutionary species–area relationship. Nature 408, 847–850 (2000).
    ADS  CAS  PubMed  Google Scholar 

    28.
    Linder, H. P. & Verboom, G. A. The evolution of regional species richness: the history of the southern African flora. Annu. Rev. Ecol. Evol. Syst. 46, 393–412 (2015).
    Google Scholar 

    29.
    Simon, M. F. et al. Recent assembly of the Cerrado, a neotropical plant diversity hotspot, by in situ evolution of adaptations to fire. Proc. Natl Acad. Sci. USA 106, 20359–20364 (2009).
    ADS  CAS  PubMed  Google Scholar 

    30.
    Cardoso, D. et al. A molecular-dated phylogeny and biogeography of the monotypic legume genus Haplormosia, a missing African branch of the otherwise American-Australian Brongniartieae clade. Mol. Phylogenet. Evol. 107, 431–442 (2017).
    PubMed  Google Scholar 

    31.
    Fritz, S. A. & Purvis, A. Selectivity in mammalian extinction risk and threat types: a new measure of phylogenetic signal strength in binary traits. Conserv. Biol. 24, 1042–1051 (2010).
    PubMed  Google Scholar 

    32.
    Linder, H. P. East African Cenozoic vegetation history. Evol. Anthropol. 26, 300–312 (2017).
    PubMed  Google Scholar 

    33.
    Retallack, G. J. Middle Miocene fossil plants from Fort Ternan (Kenya) and evolution of African grasslands. Paleobiology 18, 383–400 (1992).
    Google Scholar 

    34.
    Uno, K. T. et al. Late Miocene to Pliocene carbon isotope record of differential diet change among East African herbivores. Proc. Natl. Acad. Sci. USA 108, 6509–6514 (2011).
    ADS  CAS  PubMed  Google Scholar 

    35.
    Dart, R. A. Australopithecus africanus: the man-ape of South Africa. Nature 115, 195–199 (1925).
    ADS  Google Scholar 

    36.
    Dembo, M. et al. The evolutionary relationships and age of Homo naledi: An assessment using dated Bayesian phylogenetic methods. J. Hum. Evol. 97, 17–26 (2016).
    PubMed  Google Scholar 

    37.
    Davies, T. J. & Buckley, L. B. Phylogenetic diversity as a window into the evolutionary and biogeographic histories of present-day richness gradients for mammals. Phil. Trans. R. Soc. B 366, 2414–2425 (2011).
    PubMed  Google Scholar 

    38.
    Maurin, O. et al. Savanna fire and the origins of the ‘underground forests’ of Africa. New Phytol. 204, 201–214 (2014).
    PubMed  Google Scholar 

    39.
    Charles-Dominique, T. et al. Spiny plants, mammal browsers, and the origin of African savannas. Proc. Natl Acad. Sci. USA 113, E5572–E5579 (2016).
    CAS  PubMed  Google Scholar 

    40.
    Bibi, F. A multi-calibrated mitochondrial phylogeny of extant Bovidae (Artiodactyla, Ruminantia) and the importance of the fossil record to systematics. BMC Evol. Biol. 13, 166 (2013).
    PubMed  PubMed Central  Google Scholar 

    41.
    Nyakatura, K. & Bininda-Emonds, O. Updating the evolutionary history of Carnivora (Mammalia): a new species-level supertree complete with divergence time estimates. BMC Biol. 10, 12 (2012).
    PubMed  PubMed Central  Google Scholar 

    42.
    Kyalangalilwa, B., Boatwright, J. S., Daru, B. H., Maurin, O. & van der Bank, M. Phylogenetic position and revised classification of Acacia s.l. (Fabaceae: Mimosoideae) in Africa, including new combinations Vachellia and Senegalia. Bot. J. Linn. Soc. 172, 500–523 (2013).
    Google Scholar 

    43.
    Silvestro, D. & Michalak, I. raxmlGUI: a graphical front-end for RAxML. Org. Divers. Evol. 12, 335–337 (2012).
    Google Scholar 

    44.
    Webb, C. O. & Donoghue, M. J. Phylomatic: tree assembly for applied phylogenetics. Mol. Ecol. Notes 5, 181–183 (2005).
    Google Scholar 

    45.
    Drummond, A. J., Suchard, M. A., Xie, D. & Rambaut, A. Bayesian Phylogenetics with BEAUti and the BEAST 1.7. Mol. Biol. Evol. 29, 1969–1973 (2012).
    CAS  PubMed  PubMed Central  Google Scholar 

    46.
    Bell, C. D., Soltis, D. E. & Soltis, P. S. The age and diversification of the angiosperms re-revisited. Am. J. Bot. 97, 1296–1303 (2010).
    PubMed  Google Scholar 

    47.
    Phillips, S. J. et al. Maximum entropy modeling of species geographic distributions. Ecol. Model. 190, 231–259 (2006).
    Google Scholar 

    48.
    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).
    Google Scholar 

    49.
    Liu, C., Berry, P. M., Dawson, T. P. & Pearson, R. G. Selecting thresholds of occurrence in the prediction of species distributions. Ecography 28, 385–393 (2005).
    Google Scholar 

    50.
    Blach-Overgaard, A., Svenning, J. C., Dransfield, J., Greve, M. & Balslev, H. Determinants of palm species distributions across Africa: the relative roles of climate, non-climatic environmental factors, and spatial constraints. Ecography 33, 380–391 (2010).
    Google Scholar 

    51.
    Ratnam, J. et al. When is a ‘forest’ a savanna, and why does it matter?. Glob. Ecol. Biogeogr. 20, 653–660 (2011).
    Google Scholar 

    52.
    Koenker, R. quantreg: Quantile Regression. R package version 5.05 (https://CRAN.R-project.org/package=quantreg, 2013).

    53.
    Richardson, J. E. et al. Rapid and recent origin of species richness in the Cape flora of South Africa. Nature 412, 181–183 (2001).
    ADS  CAS  PubMed  Google Scholar 

    54.
    Verboom, G. A. et al. Origin and diversification of the Greater Cape flora: ancient species repository, hot-bed of recent radiation, or both?. Mol. Phylog. Evol. 51, 44–53 (2009).
    Google Scholar 

    55.
    Dynesius, M. & Jansson, R. Evolutionary consequences of changes in species’ geographical distributions driven by Milankovitch climate oscillations. Proc. Natl Acad. Sci. USA 97, 9115–9120 (2000).
    ADS  CAS  PubMed  Google Scholar 

    56.
    deMenocal, P. B. African climate change and faunal evolution during the Pliocene-Pleistocene. Earth Planet. Sci. Lett. 220, 3–24 (2004).
    ADS  CAS  Google Scholar  More

  • in

    Exploring the upper pH limits of nitrite oxidation: diversity, ecophysiology, and adaptive traits of haloalkalitolerant Nitrospira

    Community composition of Nitrospira in the saline-alkaline lakes
    Members of the genus Nitrospira are the most diverse and widespread known NOB. However, reports of Nitrospira occurrence in alkaline habitats are scarce [23, 30], and a systematic assessment of their presence and activity in such extreme environments is missing. In this study, we discovered and investigated unusually alkalitolerant Nitrospira in saline-alkaline lakes of the national park “Neusiedler See-Seewinkel”, Burgenland, Austria using targeted amplicon profiling of the 16S rRNA gene and nxrB, of which the latter encodes the beta-subunit of nitrite oxidoreductase (the key enzyme for nitrite oxidation). In sediment samples from nine lakes, we detected phylogenetically diverse Nitrospira phylotypes which were affiliated with Nitrospira lineages I, II and IV (Fig. 2) [1].
    Fig. 2: Phylogenetic maximum likelihood analysis based on the 16S rRNA gene sequences of selected representatives from the genus Nitrospira and of the Nitrospira members detected in sediments from nine saline-alkaline lakes.

    Sequences obtained in this study are printed in bold. “Ca. N. alkalitolerans” is the Nitrospira species cultured and further analyzed in this study. The tree was constructed using full length sequences and a 50% conservation filter resulting in 1310 valid alignment positions. Shorter sequences from this study, generated through amplicon and Sanger sequencing were added to the tree using the Evolutionary Placement Algorithm (EPA) without changing the overall tree topology. Numbers in brackets behind these sequences firstly denote the likelihood score of the exact placement and secondly the cumulative likelihood score of the placement within the cluster. Filled, gray, and open circles denote branches with ≥90%, ≥70% and ≥50% bootstrap support, respectively. Leptospirillum ferrooxidans (AJ237903), Ca. Magnetobacterium bavaricum (FP929063), Thermodesulfovibrio yellowstonii DSM 11347 (CP001147), and Ca. Methylomirabilis oxyfera (FP565575) were used as outgroup. The scale bar indicates 6% estimated sequence divergence.

    Full size image

    The genomes of sequenced Nitrospira possess one to six paralogous copies of nxrB, and the nxrB copy numbers per genome remain unknown for the majority of uncultured Nitrospira [42]. This large variability likely affects relative abundance estimations of Nitrospira OTUs based on nxrB amplicon data. In contrast, all sequenced Nitrospira genomes contain only one ribosomal RNA (rrn) operon. Therefore, our further assessment of the Nitrospira community structures relies on the 16S rRNA gene amplicon datasets.
    The estimated alpha-diversity of Nitrospira 16S rRNA gene phylotypes was compared across the nine examined lakes (Fig. S1). The inverse Simpson’s index of the Nitrospira communities was negatively correlated with pH and the nitrite concentration (p = 0.00004, Tau-b = −0.53 for pH and p = 0.03, Tau-b = −0.36 for nitrite). The decrease of Nitrospira diversity with increasing pH may indicate that only specific Nitrospira phylotypes tolerate highly alkaline conditions.
    The Nitrospira communities clustered into two distinct major groups (Fig. 3). Group 1 mainly comprised the communities from those lakes, which are located closely to the shore of the much larger Lake Neusiedl, whereas group 2 contained the communities from the remaining lakes that are farther away from Lake Neusiedl (Fig. 1). The average pH and salinity in the water of lakes from the group 1 cluster were 9.97 ± 0.24. and 6.1 ± 4.1 g/l, respectively. These values were significantly higher (Welch’s t-test; p = 0.00001 for pH and p = 0.017 for salinity) than the mean pH of 9.37 ± 0.26 and salinity of 2.74 ± 0.88 g/l in the group 2 lakes (Table 1). None of the other determined lake properties at time of sampling differed significantly between the two groups. The Nitrospira phylotypes with the highest relative abundance in the sediments from group 1 were OTU1 and OTU20, both affiliated with Nitrospira lineage IV, whereas these OTUs were nearly absent from the sediments of the lakes in group 2 (Fig. 3). In contrast, the predominant phylotypes in the group 2 lake sediments were affiliated with Nitrospira lineage II (Fig. 3). Consistent with these results, a principal coordinate analysis showed a clear separation of the Nitrospira communities with the same two groups separated on the first axis of the ordination (Fig. S2). These results indicate a strong influence of pH and salinity on the composition of the Nitrospira communities. Members of Nitrospira lineage IV are adapted to saline conditions and are commonly found in marine ecosystems [15, 43,44,45,46,47]. However, to date no Nitrospira species have been described to tolerate elevated pH conditions. Our results show that a substantial diversity of Nitrospira is able to colonize alkaline environments. The data also indicate a niche differentiation between lineages IV and II in saline-alkaline lakes, which likely includes a higher tolerance of the detected lineage IV organisms toward an elevated pH and salinity.
    Fig. 3: Normalized abundances of Nitrospira 16S rRNA gene phylotypes detected in triplicate sediment samples from nine saline-alkaline lakes.

    Nitrospira communities are grouped by hierarchical clustering on the y-axis, and OTUs are grouped by phylogenetic affiliation on the x-axis. Lake names are abbreviated as in  Fig. 1. Lin. IV, Nitrospira lineage IV ; Lin. II, Nitrospira lineage II; I, Nitrospira lineage I; Freq normalized frequency counts; Grp.1, group 1 lakes; Grp.2, group 2 lakes (see also Fig. 1).

    Full size image

    Metagenome sequencing and physiology of alkalitolerant Nitrospira enrichments
    Following the inoculation of mineral nitrite medium flasks with sediment and/or water samples from four saline-alkaline lakes (LL, WW, KS and OEW; abbreviations see Table 1), we initially obtained 17 enrichment cultures that oxidized nitrite to nitrate. Based on FISH analyses with Nitrospira-specific 16S rRNA gene-targeted probes and Sanger sequencing of cloned 16S rRNA genes, several of these preliminary enrichment cultures contained co-existing phylotypes from Nitrospira lineages I, II, and IV as well as from the genus Nitrobacter (data not shown). Members of the genera Nitrotoga and Nitrospina were screened for by FISH or PCR, but were not detected.
    We used three of the enrichments which contained only Nitrospira NOB and originated from different lakes (referred to as EN_A from lake OEW, EN_B from lake LL, and EN_C from lake WW comprising ~35% Nitrospira in relation to the total microbial community based on FISH analysis) to determine the pH range for activity of the enriched Nitrospira members. Enrichment cultures EN_A and EN_C contained phylotypes from Nitrospira lineages I and II, while EN_B contained phylotypes from lineages I, II, and IV as determined by 16 rRNA gene amplicon cloning and Sanger sequencing (Fig. 2). The continued presence of these Nitrospira phylotypes for more than 2 years, despite several serial dilution transfers, demonstrates their tolerance to the alkaline incubation conditions and suggests that they were native to the saline-alkaline environment which they were sampled from. Hence, we conclude that at least the highly similar uncultured Nitrospira OTUs detected by amplicon sequencing (Fig. 2) were most likely also native inhabitants of the saline-alkaline lakes. Aliquots of each enrichment culture were incubated with nitrite as the sole added energy source for six weeks at pH 7.61–7.86 and 9–9.04, respectively. During this period, pH had no significant effect on nitrite utilization (Pearson correlation coefficient ≥0.96 with, p ≤ 0.01 for all three enrichments) and nitrate production (Pearson correlation coefficient ≥0.98 with, p ≤ 0.01 for all three enrichments) over time for any of the three enrichments (Fig. S3). Subsequently, the enrichment culture aliquots that had been incubated at pH 9–9.04 were sequentially incubated at pH 9.97–10, 10.24–10.52, and 10.72–11.02 for eight to nine days at each pH (Table S1). For all three enrichments, the observed nitrate production tended to be slower at pH 9.97–10 and 10.24–10.52 than at pH 9–9.04 (Fig. S3 and S4). At pH 10.72–11.02, no nitrite consumption was detected (Fig. S4). The trends observed at pH 10.24–10.52 and above were in stark contrast to the persistently high nitrite-oxidizing activity of the enrichments when routinely cultured at pH 9–10 for several weeks. While it was not possible to determine based on our data whether all Nitrospira phylotypes present in the three enrichments responded equally to the tested pH conditions, we can conclude that the activity of at least some Nitrospira remained unaffected up to pH 9 and had an upper limit between pH 10.5 and 10.7. This is remarkable, because previously enriched or isolated Nitrospira strains were not cultivated above pH 8.0 except for two Nitrospira cultures from geothermal springs, which showed activity up to pH 8.8 [4] or pH 9.0 [7]. To our knowledge, this is the first report of nitrite oxidation by Nitrospira at pH values above 9 and as high as 10.5.
    Further analyses focused on one additional enrichment, which had been inoculated with sediment from lake Krautingsee, belonging to the group 2 of the analyzed lakes (KS, Table 1). In contrast to the other enrichment cultures, this enrichment contained only lineage IV Nitrospira based on FISH analysis (Fig. 4a). Nitrospira-specific, 16S rRNA gene and nxrB-targeted PCR and phylogeny detected one phylotype from Nitrospira lineage IV that was related to other phylotypes detected from the lakes, specifically OTU 5 and EN_B_1 (16S rRNA gene, 100% and 98% nucleotide sequence identity, respectively; Fig. 2) and OTU 2 (nxrB, 98.5% nucleotide sequence identity; Fig. S5). Both these OTU phylotypes occurred in most of the analyzed lakes (Fig. 3). Thus, the closely related enrichment from lake KS may represent Nitrospira that could adapt to a relatively broad range of conditions, while some of the other OTUs were more abundant in specific lakes only (Fig. 3). The enriched Nitrospira reached a high relative abundance in the enrichment culture of ~60% of all bacteria based on metagenomic read abundance (see below) and observation by FISH.
    Fig. 4: Visualization and metagenomic analysis of the “Ca. N. alkalitolerans” enrichment.

    a FISH image showing dense cell clusters of “Ca. N. alkalitolerans” in the enrichment culture. The “Ca. N. alkalitolerans” cells appear in red (labeled by probe Ntspa1151 which has 1 mismatch at the 3’ end to the 16S rRNA gene sequence of “Ca. N. alkalitolerans”; the absence of lineage II Nitrospira in the enrichment culture was confirmed by the application of the competitor oligonucleotides c1Ntspa1151 and c2Ntspa1151 as indicated in the Supplementary text). Other organisms were stained by DAPI and are shown in light gray. Scale bar, 25 µm. b Phylogenetic affiliation of the metagenome scaffolds from the “Ca. N. alkalitolerans” enrichment, clustered based on sequence coverage and the GC content of DNA. Closed circles represent scaffolds, scaled by the square root of their length. Clusters of similarly colored circles represent potential genome bins.

    Full size image

    High-throughput metagenome sequencing, scaffold assembly, and binning revealed that the enrichment contained three Nitrospira strains that could be separated into three genome bins based on sequence coverage data (Table S2, Fig. S6). No other NOB were identified in the metagenome, and the three Nitrospira bins represented the most abundant organisms in the enrichment culture (Fig. 4b). Since the genome-wide average nucleotide identity (gANI) values were above the current species threshold of 95% [48] (Table S2), the three bins likely represented very closely related strains of the same Nitrospira lineage IV species with unique genetic components. From the predominant (based on coverage data) Nitrospira sequence bin, an almost complete metagenome-assembled genome (MAG) was reconstructed, which met the criteria for a “high-quality draft” genome [49] (Table S2), and used for comparative genomic analysis. Genome-wide, pairwise comparison of the gANI and average amino acid (gAAI) identity between this MAG and Nitrospira marina as the only other genome-sequenced and cultured Nitrospira lineage IV representative resulted in values of 80.1 and 77.3, respectively. The 16S rRNA gene, which had been retrieved from the MAG, was 97.90% identical to the 16S rRNA gene of N. marina, 97.87% identical to “N. strain Ecomares 2.1”, 94.92% to “Ca. N. salsa”, and 94.51% to “Nitrospira strain Aa01”, which are the other cultured members of Nitrospira lineage IV [15, 43, 46, 47]. These values are below the current species threshold of 98.7–99% for 16S rRNA genes [50]. Based on the low gANI and 16S rRNA gene sequence identities to described Nitrospira species, and additionally considering the distinct haloalkalitolerant phenotype (see also below), we conclude that the enriched Nitrospira represent a new species and propose “Ca. Nitrospira alkalitolerans” as the tentative name.
    The enrichment culture was maintained at a pH of 9–10 and a salt concentration of 2 g/l, resembling the natural conditions in the saline-alkaline lakes based on available data from 5 years. “Ca. N. alkalitolerans” grew in dense flocks (Fig. 4a), thereby possibly relieving the pH stress [51]. Its nitrite-oxidizing activity was not affected when the pH in the cultivation medium decreased below 8. However, no nitrite oxidation was observed when the enrichment culture was transferred into medium with 4× to 8× higher salt concentrations, the latter resembling marine conditions. Thus, “Ca. N. alkalitolerans” is best described as a facultatively haloalkalitolerant organism that oxidizes nitrite as an energy source over a wide range of pH and under hyposaline conditions. This phenotype is certainly advantageous in the investigated saline-alkaline lakes, as these lakes are prone to evaporation in summer, which causes a temporarily elevated salinity and alkalinity in the remaining water body and the sediment [35].
    The enrichment culture of “Ca. N. alkalitolerans” oxidized nitrite over a broad range of initial nitrite concentrations tested, although an extended lag phase of 10–15 days occurred at the higher concentrations of 0.7 and 1 mM nitrite (Fig. S7). Similarly, a lag phase at elevated nitrite concentrations was also observed for the Nitrospira lineage II member Nitrospira lenta [52]. A preference for low nitrite levels is consistent with the presumed ecological role of nitrite-oxidizing Nitrospira as slow-growing K-strategists, which are adapted to low nitrite concentrations [50, 52, 53].
    Genomic adaptations to the saline-alkaline environment
    As described below, comparative genomic analysis of “Ca. N. alkalitolerans” revealed several features that distinguish this organism from other known NOB and likely form the basis of its tolerance toward elevated alkalinity and salinity (Fig. 5).
    Fig. 5: Cell metabolic cartoon constructed from the genome annotation of “Ca. N. alkalitolerans”.

    Features putatively involved in the adaptation to high alkalinity and salinity, and selected core metabolic pathways of chemolithoautotrophic nitrite-oxidizing Nitrospira, are shown. Note that the transport stoichiometry of the ion transporters in “Ca. N. alkalitolerans” remains unknown. Colors of text labels indicate whether adaptive features are present (i.e., have homologs) in the genomes of other NOB (red, feature is not present in any other characterized NOB; blue, feature is present only in the marine Nitrospina gracilis; purple, feature is present in several other characterized NOB).

    Full size image

    Cytoplasmic pH and ion homeostasis
    At high pH, alkaliphilic and alkalitolerant microbes maintain a higher transmembrane electrical potential (ΔΨ) component of the proton motive force (PMF) than usually found in neutrophiles. The high ΔΨ is required to maintain PMF, because the ΔpH component of the PMF is reversed when the extracellular pH is higher than the intracellular pH [54]. Like in neutrophiles, the ΔΨ of alkaliphiles is negative inside the cell relative to the outside [54]. Furthermore, the intracellular pH must be kept below the (extremely) alkaline extracellular pH. At elevated salinity, resistance against high salt concentrations is an additional, fundamental necessity for survival. All this requires a tightly regulated pH and ion homeostasis, in which cation transmembrane transporters play key roles [54,55,56]. The “Ca. N. alkalitolerans” genome codes for various Na+-dependent transporters (Fig. 5, Table S3) including secondary Na+/H+ antiporters that are involved in pH homeostasis in other organisms: two copies of a group 3 Mrp-type Na+/H+ antiporter [57, 58] encoded by the seven genes mrpA-G, and monovalent cation-proton antiporters of the types NhaA and NhaB, each of which is encoded by a single gene [59]. The Mrp antiporter is crucial for growth at high pH and elevated salinity in alkaliphilic Halomonas spp. and Bacillus spp., where it exports Na+ and imports H+, thus contributing to the maintenance of a lower intracellular pH compared to the environment (e.g., cytoplasmic pH 8.3 at external pH ~ 10.5) [[60] and references cited therein, [55]]. The Mrp proteins may form a large surface at the outside of the cytoplasmic membrane that could support proton capture under alkaline conditions [54, 57]. Nha-type antiporters are widely distributed among non-extremophilic and extremophilic organisms [55]. Being involved in the homeostasis of Na+ and H+, they are important for survival under saline and/or alkaline conditions [56]. In E. coli, NhaA is regulated by the cytoplasmic pH and it catalyzes the import of 2H+ with the concurrent export of one Na+. This electrogenic activity is driven by ΔΨ and maintains pH homeostasis at elevated external pH [[52] and references cited therein]. The simultaneous presence of the two antiporters NhaA and NhaB has been associated with halophilic or haloalkaliphilic phenotypes in other organisms [55, 59]. Although the regulation and cation transport stoichiometry of the homologs in “Ca. N. alkalitolerans” remain unknown, the Mrp- and Nha-family antiporters most likely exhibit important physiological roles in this organism and support its survival under haloalkaline conditions. Possibly, “Ca. N. alkalitolerans” can even combine its growth in dense flocks with the extrusion of protons by its numerous proton transporters thereby lowering the pH inside the flock [51].
    One of the two nhaB genes present in the “Ca. N. alkalitolerans” genome is located in an interesting genomic region that also contains all genes encoding the group 3 Mrp-type Na+/H+ antiporter (Fig. S8). The two genes downstream from mrpD display sequence similarity to the NADH dehydrogenase (complex I) subunits NuoM and NuoL. However, based on the genomic context they are more likely additional mrpA- and/or mrpD-like genes, as these Na+/H+ antiporter subunits are evolutionary related to NuoM and NuoL [61]. Multiple copies of subunits NuoM and NuoL of the NADH dehydrogenase are encoded elsewhere in the genome, partially in larger nuo operons (see Table S3). Moreover, the locus contains one gene coding for the low-affinity, high flux Na+/HCO3− uptake symporter BicA [62] and gene motB encoding a H+-translocating flagellar motor component (Fig. S8). In the haloalkalitolerant cyanobacterium Aphanothece halophytica, a similar clustering of bicA with genes coding for Na+/H+ antiporters has been described. The authors proposed a model of cooperation between these transporters, where Na+ extruded by the Na+/H+ antiporters could drive the uptake of HCO3− by BicA under alkaline conditions when CO2 becomes limiting [63]. Sodium-driven import of HCO3− could be an essential feature for “Ca. N. alkalitolerans”, because bicarbonate is the main source of inorganic carbon for autotrophic organisms, but becomes less accessible at high pH >10 [55]. A carbonic anhydrase, which is also present in the genome (Fig. 5, Table S3), can convert the imported HCO3− to CO2 for carbon fixation via the reductive tricarboxylic acid cycle (Fig. 5).
    Since cytoplasmic K+ accumulation may compensate for Na+ toxicity at elevated intracellular pH [64], many alkaliphiles retain an inward directed K+ gradient [55]. The potassium uptake transporters of the Trk family contribute to pH and K+ homeostasis of halo- and/or alkaliphiles [55]. TrkAH catalyzes the NAD+-regulated uptake of K+ possibly coupled with H+ import [65]. Moreover, kinetic experiments revealed that TrkAH of the gammaproteobacterium Alkalimonas amylolytica is salt-tolerant and functions optimally at pH > 8.5 [66]. “Ca. N. alkalitolerans” encodes a TrkAH complex (Fig. 5, Table S3), which may be a specific adaptation to its haloalkaline environment as no homologous K+ transporter has been identified yet in any other NOB genome. Under more neutral pH conditions, Kef-type K+ efflux pumps, which are present in two copies in the “Ca. N. alkalitolerans” genome, could excrete excess K+ (Fig. 5, Table S3).
    Adaptations of the energy metabolism
    Aside from the different cation transporters (see above), “Ca. N. alkalitolerans” also encodes several mechanisms for cation homeostasis that are linked to membrane-bound electron transport and energy conservation. Like in other aerobic alkaliphiles [56], ATP synthesis is likely catalyzed by a canonical, H+-translocating F1FO-ATPase (Fig. 5, Table S3). In addition, the genome contains all genes of a predicted Na+-translocating N-ATPase [67] (Fig. 5, Fig. S9, Table S3). N-ATPases form a separate subfamily of F-type ATPases and have been suggested to be ATP-driven ion pumps that extrude Na+ cations [67] or H+ [68]. The c subunit of the N-ATPase in the genome of “Ca. N. alkalitolerans” contains the typical amino acid motifs for Na+ binding and transport [67] (Fig. S10). Subunits a and c of the N-ATPase, which are involved in ion transport, are most similar to homologs from the halotolerant, sulfate-reducing Desulfomicrobium baculatum (81.5% AA identity) and the haloalkalitolerant, sulfur-oxidizing Sulfuricella denitrificans (88.2% AA identity), respectively. Hence, in “Ca. N. alkalitolerans”, the N-ATPase may contribute to the maintenance of ΔΨ, the generation of a sodium motive force (SMF), and salt resistance (Fig. 5).
    The genome of “Ca. N. alkalitolerans” encodes two different types of NADH:quinone oxidoreductase (complex I of the electron transport chain) (Fig. 5, Table S3). Firstly, the organism possesses all 14 genes of type I NADH dehydrogenase (nuoA to nuoN). They are present in one to three copies each. The nuo genes are mostly clustered at several genomic loci (Table S3) and are most similar to either of the two nuo operons present in Nitrospira defluvii [39], with AA identities between 41% and 90%. As mentioned above, nuoL/M-like genes at loci without other nuo genes might represent subunits of cation antiporters.
    The genome furthermore contains a locus encoding all six subunits of a Na+-dependent NADH:quinone oxidoreductase (Nqr or type III NAD dehydrogenase) (Fig. 5, Table S3). The locus is situated on a single contig in the vicinity of transposase genes, indicating that “Ca. N. alkalitolerans” might have received this type of complex I by lateral gene transfer. The gene of subunit E, which takes part in Na+ translocation [69], is most similar to a homolog in the ammonia-oxidizing bacterium Nitrosomonas nitrosa (86% AA identity).
    The metabolic model for N. defluvii [39] assumes that two different versions of the H+-dependent complex I (Nuo) are used for forward or reverse electron transport, respectively. Nitrospira possess a canonical Nuo that is likely used for PMF generation during the forward flow of low-potential electrons from the degradation of intracellular glycogen or from hydrogen as an alternative substrate (see also below). In addition, reverse electron transport is essential in NOB to generate reducing power for CO2 fixation. In Nitrospira, a second (modified) form of Nuo with duplicated proton-translocating NuoM subunits might use PMF to lift electrons from quinol to ferredoxin [70]. The reduced ferredoxin is required for CO2 fixation via the rTCA cycle. As expected, “Ca. N. alkalitolerans” possesses these two Nuo forms that are conserved in other characterized Nitrospira members. In addition, the Na+-dependent Nqr complex might function in two directions in “Ca. N. alkalitolerans” as well. During forward electron flow, Nqr would contribute to SMF generation (Fig. 5). Reverse operation of the Nqr could generate NADH while importing Na+, thus utilizing SMF for the reduction of NAD+ with electrons derived from quinol (Fig. 5). Hence, the two types of complex I are likely involved in essential electron transport and the fine-tuning of PMF and SMF. They probably cooperate with the Na+- and the H+-translocating ATPases and the various cation transporters (see above) to adjust the cytoplasmic ion concentrations and the membrane potential in response to the environmental salinity and pH.
    In addition to a novel “bd-like” cytochrome c oxidase, which is commonly found in Nitrospira genomes [16, 39], the genome of “Ca. N. alkalitolerans” contains a locus with fused genes for a cbb3-type cytochrome c oxidase (Fig. 5, Table S3) similar to the one present in the marine nitrite oxidizer Nitrospina gracilis [41]. The cbb3-type terminal oxidases usually exhibit high affinities for O2 [71] and may allow “Ca. N. alkalitolerans” to sustain respiration at low oxygen levels.
    Interestingly, “Ca. N. alkalitolerans” encodes two different hydrogenases and the accessory proteins for hydrogenase maturation (Fig. 5, Table S3). First, it possesses a group 2a uptake hydrogenase that is also found in N. moscoviensis, which can grow autotrophically on H2 as the sole energy source [16]. Second, “Ca. N. alkalitolerans” codes for a putative bidirectional group 3b (sulf)hydrogenase that also occurs in other NOB and in comammox Nitrospira [18, 41] but has not been functionally characterized in these organisms. Experimental confirmation of H2 utilization as an alternative energy source and electron donor by “Ca. N. alkalitolerans” is pending. However, we assume that this capability would confer ecophysiological flexibility, especially if nitrite concentrations fluctuate and H2 is available at oxic-anoxic boundaries in biofilms or upper sediment layers. While electrons from the group 2a hydrogenase are probably transferred to quinone [16], the group 3b hydrogenase might reduce NAD+ [41] and fuel forward electron transport through the Nuo and Nqr complexes (see above).
    Osmoadaptation
    The intracellular accumulation of compatible solutes is an important mechanism allowing microorganisms to withstand the high osmotic pressure in saline habitats [55]. “Ca. N. alkalitolerans” has the genetic capacity to synthesize or import the compatible solutes trehalose, glycine betaine, and glutamate (Fig. 5). For trehalose synthesis the gene treS of trehalose synthase (Table S3), which enables trehalose synthesis from maltose, is present. The genes opuD and opuCB for glycine betaine import (Table S3) have been identified in the marine Nitrospina gracilis [41], but not yet in any Nitrospira species. For glutamate synthesis, the genes gltB and gltD were identified (Table S3). They code for the alpha and beta subunits of glutamate synthase, which catalyzes L-glutamate synthesis from L-glutamine and 2-oxoglutarate with NADPH as cofactor. In addition, we identified adaptations of “Ca. N. alkalitolerans” to the low availability of iron and the presence of toxic arsenite in saline-alkaline systems (Supplementary text). More