More stories

  • in

    Methane emission from high latitude lakes: methane-centric lake classification and satellite-driven annual cycle of emissions

    1.
    IPCC: Climate Change 2014: Synthesis Report. Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [eds. Pachauri, R.K & L.A. Meyer) (IPCC, 2014).
    2.
    AMAP (Arctic Monitoring and Assessment Programme). Snow, Water (Ice and Permafrost in the Arctic, AMAP, Tromsø, 2017).
    Google Scholar 

    3.
    Wik, M., Varner, R. K., Walter Anthony, K., MacIntyre, S. & Bastviken, D. Climate-sensitive northern lakes and ponds are critical components of methane release. Nat. Geosci. 9, 99–105 (2016).
    ADS  CAS  Article  Google Scholar 

    4.
    Holgerson, M. A. & Raymond, P. A. Large contribution to inland water CO2 and CH4 emissions from very small ponds. Nat. Geosci. 9, 222–226 (2016).
    ADS  CAS  Article  Google Scholar 

    5.
    Bastviken, D., Cole, J., Pace, M. & Tranvik, L. Methane emissions from lakes: Dependence of lake characteristics, two regional assessments, and a global estimate. Global Biogeochem. Cyc. 18, 1–12 (2004).
    Article  Google Scholar 

    6.
    Bastviken, D., Tranvik, L. J., Downing, J. A., Crill, P. M. & Enrich-Prast, A. Freshwater methane emissions offset the continental carbon sink. Science 331, 50–51 (2011).
    ADS  CAS  Article  Google Scholar 

    7.
    Walter, K. M., Smith, L. C. & Chapin, F. S. Methane bubbling from northern lakes: present and future contributions to the global methane budget. Philos. Trans. R. Soc. A 365, 1657–1676 (2007).
    ADS  CAS  Article  Google Scholar 

    8.
    Walter Anthony, K. M. & Anthony, P. Constraining spatial variability of methane ebullition seeps in thermokarst lakes using point process models. J. Geophys. Res. Biogeosci. 118, 1015–1034. https://doi.org/10.1002/jgrg.20087 (2013).
    Article  Google Scholar 

    9.
    Tan, Z. & Zhuang, Q. Arctic lakes are continuous methane sources to the atmosphere under warming conditions. Environ. Res. Lett. 10, 1–9 (2015).
    Article  Google Scholar 

    10.
    Tan, Z. & Zhuang, Q. Methane emissions from pan-Arctic lakes during the 21st century: An analysis with process-based models of lake evolution and biogeochemistry. J. Geophys. Res. Biogeosci. 120, 2641 (2015).
    CAS  Article  Google Scholar 

    11.
    Lehner, B. & Doell, P. Development and validation of a global database of lakes, reservoirs and wetlands. J. Hydrol. 296, 1–22 (2004).
    ADS  Article  Google Scholar 

    12.
    Verpoorter, C., Kutser, T., Seekell, D. A. & Tranvik, L. J. A global inventory of lakes based on high-resolution satellite imagery. Geophys. Res. Lett. 41, 6396–6402 (2014).
    ADS  Article  Google Scholar 

    13.
    Messager, M. L., Lehner, B., Grill, G., Nedeva, I. & Schmitt, O. Estimating the volume and age of water stored in global lakes using a geo-statistical approach. Nat. Commun. 7, 13603 (2016).
    ADS  CAS  Article  Google Scholar 

    14.
    Downing, J. A. & Duarte, C. M. Abundance and size distribution of lakes, ponds, and impoundments. In Encyclopedia of Inland Water, 1 (ed. Likens, G. E.) 469–478 (Elsevier, Amsterdam, 2009).
    Google Scholar 

    15.
    McGinnis, D. F., Greinert, J., Artemov, Y., Beaubien, S. E. & Wüest, A. Fate of rising methane bubbles in stratified waters: how much methane reaches the atmosphere?. J. Geophys. Res. 111, C09007 (2006).
    ADS  Article  Google Scholar 

    16.
    Lamarche, C. et al. Compilation and validation of SAR and optical data products for a complete and global map of inland/ocean water tailored to the climate modeling community. Rem. Sens. 9, 36 (2017).
    ADS  Article  Google Scholar 

    17.
    Pekel, J.-F., Cottam, A., Gorelick, N. & Belward, A. S. High-resolution mapping of global surface water and its long-term changes. Nature 540, 418–422 (2016).
    ADS  CAS  Article  Google Scholar 

    18.
    Sanches, L. F., Guenet, B., Marinho, C. C., Barros, N. & de Assis Esteves, F. Global regulation of methane emission from natural lakes. Sci. Rep. 9, 255 (2019).
    ADS  Article  Google Scholar 

    19.
    Bruhwiler, L. et al. CarbonTracker-CH4: an assimilation system for estimating emissions of atmospheric methane. Atmos. Chem. Phys. 14, 8269–8293 (2014).
    ADS  Article  Google Scholar 

    20.
    Du, J., Kimball, J. S., Duguay, C., Kim, Y. & Watts, J. D. Satellite microwave assessment of Northern Hemisphere lake ice phenology from 2002 to 2015. Cryosphere 11, 47–63 (2017).
    ADS  Article  Google Scholar 

    21.
    Kim, Y., Kimball, J. S., McDonald, K. C. & Glassy, J. Developing a global data record of daily landscape freeze/thaw status using satellite microwave remote sensing, Version 4. IEEE Trans. Geosci. Rem. Sens. 49, 949–960 (2016).
    ADS  Article  Google Scholar 

    22.
    Brown, J., Ferrians, O.J., Heginbottom, J.A. & Melnikov, E.S. Circum-Arctic map of permafrost and ground-ice conditions, Version 2. Boulder, CO, National Snow and Ice Data Center/World Data Center for Glaciology. https://doi.org/10.3133/cp45 (2002).

    23.
    Obu, J. et al. Northern hemisphere permafrost map based on TTOP modelling for 2000–2016 at 1 km2 scale. Earth Sci. Rev. 193, 299–316 (2019).
    ADS  Article  Google Scholar 

    24.
    Harmonized World Soil Database (HWSD) https://daac.ornl.gov/SOILS/guides/HWSD.html.

    25.
    Allen, G. H. & Pavelsky, T. M. Global extent of rivers and streams. Science 361, 585–588 (2018).
    MathSciNet  CAS  Article  Google Scholar 

    26.
    Lehner, B. et al. High-resolution mapping of the world’s reservoirs and dams for sustainable river-flow management. Front. Ecol. Environ. 9(9), 494–502 (2011).
    Article  Google Scholar 

    27.
    Mulligan, M., Saenz-Cruz, L., van Soesbergen, A., Smith, V.T. & Zurita, L. The Global georeferenced Database of Dams (GOOD2), Version 1. Global dams database and geowiki. https://geodata.policysupport.org/dams (2009).

    28.
    Chau, Y. K., Snodgrass, W. J. & Wong, P. T. S. A sampler for collecting evolved gases from sediment. Water Res. 11, 807–809 (1977).
    CAS  Article  Google Scholar 

    29.
    Howard, D.L., Frea, J.I. & Pfister, R.M. The potential for methane carbon cycling in Lake Erie. In Paper Presented at 14th Conference on Great Lakes Research (Int. Assoc. of Great Lakes Res., Ann Arbor, Mich. 1971).

    30.
    Townsend-Small, A. et al. Quantifying emissions of methane derived from anaerobic organic matter respiration and natural gas extraction in Lake Erie. Limnol. Oceanogr. 61, S356–S366 (2016).
    CAS  Article  Google Scholar 

    31.
    Joung, D., Leonte, M. & Kessler, J. D. Methane sources in the waters of Lake Michigan and Lake Superior as revealed by natural radiocarbon measurements. Geophys. Res. Lett. 46, 5436–5444 (2019).
    ADS  CAS  Article  Google Scholar 

    32.
    Shimoda, Y. et al. Our current understanding of lake ecosystem response to climate change: what have we really learned from the north temperate deep lakes?. J. Great Lakes Res. 37, 173–193 (2011).
    CAS  Article  Google Scholar 

    33.
    Blenckner, T. R. et al. Large-scale climatic signatures in lakes across Europe: a meta-analysis. Glob. Change Biol. 13, 1314–1326 (2007).
    ADS  Article  Google Scholar 

    34.
    van Huissteden, J. et al. Methane emissions from permafrost thaw lakes limited by lake drainage. Nat. Clim. Change 1, 119–123 (2011).
    ADS  Article  Google Scholar 

    35.
    Kalff, J. Limnology, Inland Water Ecosystems (Prentice Hall, Upper Saddle River, 2002).
    Google Scholar 

    36.
    Kourzeneva, E., Asensio, H., Martin, E. & Faroux, S. Global gridded dataset of lake coverage and lake depth for use in numerical weather prediction and climate modelling. Tellus A 64, 1–14 (2012).
    Google Scholar  More

  • in

    Estimating illegal fishing from enforcement officers

    Experimental design
    Following five focus groups with SERNAPESCA’s head of enforcement and other personnel, we designed and implemented an online survey that targeted fisheries enforcement officers who are responsible for monitoring IUU activities in Chile. The survey was structured to capture expert knowledge on various aspects of illegal activities, as well as the relative experience of the officers. The survey defined illegal fishing as a fishing activity carried out in national jurisdiction waters by national or international boats that is in violation of the national fishing law, conducted without a legal permit, or activities that involve unreported or misreported captures to the authorities. The Director of SERNAPESCA delivered the survey via email to all SERNAPESCA enforcement officers. The list of officers was constructed by the Director (n = 86). The survey was anonymous in that the officers were not asked to report their name nor any information that could be used for identification (e.g., email). Answers to questions were not mandatory; that is, respondents could opt-out of answering particular questions and continue with the survey. The survey was available online for ten weeks, over which five reminder emails were sent to officers requesting them to complete the survey.
    The survey, in Spanish, consisted of two sections. First, we asked respondents to rank the magnitude of illegal activity for twenty fisheries on a nominal scale (1–5), along with their relative experience with each fishery (nominal scale, 1–5). The twenty fisheries were selected a priori based on our focus groups and known information about illegal activity. All fisheries were single species, with the exception of four that included multiple species: skates (2 species, Zearaja chilensis and Bathyraja macloviana), kelp (4 species: Lessonia spicate, L. berteroana, L. traberculata, Macrocystis pyrifera), red algae (3 species: Sarcothalia crispate, Gigartina skottsbergii, Mazzaella laminarioides), and crabs (10 species excluding southern king crab: Cancer edwardsi, C. porter, C. setosus, C. coronatus, Homalaspis plana, Ovalipes trimaculatus, Taliepus dentatus, T. marginatus, Mursia gaudichaudi, Hemigrapsus crenulatus). In the second part of the survey, we asked respondents additional questions for four focal fisheries: South Pacific hake (Merluccius gayi gayi), southern hake (M. australis), loco or Chilean abalone (Concholepas concholepas), and kelp. For each fishery, we asked respondents to score on a nominal scale (1–5),

    The frequency of six specific illegal activities in the industrial sector: size, gear, season, area, transshipment, and port.

    The frequency of six specific illegal activities in the small-scale sector: size, gear, season, area, transshipment, and port.

    The participation of illegal activity for six different stakeholders along the supply chain: fisher, purchaser, processor, wholesaler, exporter, and restaurateur.

    The utilization of seven infrastructure types in illegal activities: fishing boats, refrigeration trucks, processing plants, markets, transshipment boats, export vehicles, and restaurants.

    This study was approved by the Advanced Conservation Strategies and Pontificia Universidad Católica ethics institutional review boards and followed guidelines established by their ethics committees, which complies with national and international standards. The surveys included a written informed consent approved by all interviewees, which acknowledged research objectives and established that the survey was anonymous and that interviewees were free to choose to not answer questions. While all species have common names in Chile (which were used in the survey), we use Fishbase and Sealifebase as the taxonomic authority and for the common names reported here to facilitate comparisions34,35.
    Statistical analysis
    For both sections of the survey, we used a Bayesian cumulative multinomial logit model to predict illegal estimates. First, we fitted a model for illegal estimates for each of the twenty fisheries jointly. Second, we fitted models for illegal estimates for various aspects of the four focal fisheries (i.e., activities, stakeholders, and infrastructure) in a single analysis for each aspect. In both models, we included a random intercept term for respondent, along with a fixed effect for fishery. We evaluated the role of experience, as self-reported by the respondents, by comparing the difference between the illegal score by a respondent for a fishery and the model prediction for that fishery across respondents. If higher levels of expertise increased or decreased the value of a respondent’s scoring, there would be a relationship between the size of the differences and the level of experience reported for a fishery. Experience may also affect the difference in mean responses (i.e., bias), potentially due to more personal experience over a longer period of time, which would lead to a correlation between expertise and mean illegality scores. Depending on the patterns observed in the data, there are several ways to control for a respondent’s experience in illegality estimates. In our case, we used experience scores as a covariate in the model.
    For the twenty fisheries, we used the following model,

    $$Prleft{{S}_{ij}=kright}=phi left({tau }_{k}-left({varvec{beta}}{{varvec{x}}}_{{varvec{i}}}+{{varvec{z}}}_{{varvec{j}}}{{varvec{V}}}_{{varvec{i}}}right)right)-phi left({tau }_{k-1}-left({varvec{beta}}{{varvec{x}}}_{{varvec{i}}}+{{varvec{z}}}_{{varvec{j}}}{{varvec{V}}}_{{varvec{j}}}right)right)$$
    (1)

    in which the probability that the score for the level of illegal landings ({S}_{ij}) for the ith species by the jth respondent is equal to category k, can be represented as a latent continuous variable which is divided into K categories, by K − 1 thresholds at ({tau }_{k}). This latent continuous variable is represented by the cumulative normal distribution, (phi). For a given observation, the regression equation is composed of coefficients multiplied times predictor variables ({varvec{beta}}{{varvec{x}}}_{{varvec{i}}}) plus a design matrix for the random effect, multiplied times the error term for the jth respondent, ({{varvec{z}}}_{{varvec{j}}}{{varvec{V}}}_{{varvec{i}}}) . The probability of that observation falling in category k, (Prleft{{S}_{ij}=kright}), is thus the probability of it being in a category equal to or smaller than k, (phi left({tau }_{k}-left({varvec{beta}}{{varvec{x}}}_{{varvec{i}}}+{{varvec{z}}}_{{varvec{j}}}{{varvec{V}}}_{{varvec{i}}}right)right)), less the probability of the observation being in a category smaller than k, (phi left({tau }_{k-1}-left({varvec{beta}}{{varvec{x}}}_{{varvec{i}}}+{{varvec{z}}}_{{varvec{j}}}{{varvec{V}}}_{{varvec{j}}}right)right)). Implemented in the R statistical language, using the brms package36, the call to fit this model looks like the following:

    $${text{Score}}; , sim ;{text{Species}} + {text{Experience }} + left( {{1}|{text{Respondent}}} right),;{text{ data}} = {text{SurveyData}},;{text{family}} = {text{cumulative}}),$$

    where Score is ({S}_{ij}) in (1) above, the fixed effects, ({varvec{beta}}{{varvec{x}}}_{{varvec{i}}}) are the experience of the respondent and the species that was scored, and (1|Respondent) denotes a random intercept model, where each has a different intercept term, drawn from a shared error distribution. For more information on the application of this model to ordinal response data, see Burkner and Vuorre37.
    For the estimates for the various aspects of the four focal fisheries, we used the following model,

    $${text{Response}}; sim ;{text{Species}} + {text{Experience}} + left( {{1}|{text{Respondent}}} right),;{text{data}} = {text{SurveyData}},;{text{family}} = {text{cumulative}}),$$

    which is structured as per (1) above, but with the responses to the various focal species questions (i.e., activities per sector, stakeholders, and infrastructure) substituted for the species scores as in (1).
    We compared both models with simpler models, including a single-term null model using leave-one-out cross-validation. We did so in the R statistical language using the loo packages36,38,39. Prior distributions for all regression terms were improper flat priors over the real numbers, the default in the brms package for population parameters. The priors on the intercept and the random effects were student t3,0,10 distributions, as per the default for uninformative priors in the brms package.
    We carried out a Principal Components Analysis (PCA) with the four focal fisheries as categorical variables and the illegal activity, stakeholder, and infrastructure estimates from the Bayesian cumulative multinomial logit model. For each fishery, we used 10,000 estimates from the model, along with a qualitative variable that represented the different factors (e.g., restaurateur). The latter has no influence on the principal components of the analysis but helps to interpret the dimensions of variability. Principal Components Analysis is especially powerful as an approach to visualize patterns, such as clusters, clines, and outliers in a dataset40. In our case, we sought to visualize whether there were common illegal factors with similar set of scores and whether there was any association between high or low scores of illegal factors and the focal fisheries. We used the FactoMineR package in the R statistical language41. More

  • in

    Gymnosperm glandular trichomes: expanded dimensions of the conifer terpenoid defense system

    1.
    Bowe, L. M., Coat, G. & DePamphilis, C. W. Phylogeny of seed plants based on all three genomic compartments: Extant gymnosperms are monophyletic and Gnetales’ closest relatives are conifers. Proc. Natl. Acad. Sci. USA 97, 4092–4097 (2000).
    ADS  Article  CAS  Google Scholar 
    2.
    Bohlmann, J. & Keeling, C. I. Terpenoid biomaterials. Plant J. 54, 656–669 (2008).
    Article  CAS  Google Scholar 

    3.
    Celedon, J. M. & Bohlmann, J. Oleoresin defenses in conifers: Chemical diversity, terpene synthases, and limitations of oleoresin defense under climate change. New Phytol. 224, 1444–1463 (2019).
    Article  CAS  Google Scholar 

    4.
    Whitehill, J. G. A. et al. Functions of stone cells and oleoresin terpenes in the conifer defense syndrome. New Phytol. 221, 1503–1517 (2019).
    Article  CAS  Google Scholar 

    5.
    Hilker, M., Kobs, C., Varama, M. & Schrank, K. Insect egg deposition induces Pinus sylvestris to attract egg parasitoids. J. Exp. Biol. 205, 455–461 (2002).
    PubMed  Google Scholar 

    6.
    Mumm, R., Schrank, K., Wegener, R., Schulz, S. & Hilker, M. Chemical analysis of volatiles emitted by Pinus sylvestris after induction by insect oviposition. J. Chem. Ecol. 29, 1235–1252 (2003).
    Article  CAS  Google Scholar 

    7.
    Martin, D. M., Gershenzon, J. & Bohlmann, J. Induction of volatile terpene biosynthesis and diurnal emission by methyl jasmonate in foliage of Norway spruce. Plant Physiol. 132, 1586–1599 (2003).
    Article  CAS  Google Scholar 

    8.
    Miller, B., Madilao, L. L., Ralph, S. & Bohlmann, J. Insect-induced conifer defense. White pine weevil and methyl jasmonate induce traumatic resinosis, de novo formed volatile emissions, and accumulation of terpenoid synthase and putative octadecanoid pathway transcripts in Sitka spruce. Plant Physiol. 137, 369–382 (2005).

    9.
    Niinemets, U., Reichstein, M., Staudt, M., Seufert, G. & Tenhunen, J. D. Stomatal constraints may affect emission of oxygenated monoterpenoids from the foliage of Pinus pinea. Plant Physiol. 130, 1371–1385 (2002).
    Article  CAS  Google Scholar 

    10.
    Harley, P., Eller, A., Guenther, A. & Monson, R. K. Observations and models of emissions of volatile terpenoid compounds from needles of ponderosa pine trees growing in situ: Control by light, temperature and stomatal conductance. Oecologia 176, 35–55 (2014).
    ADS  Article  Google Scholar 

    11.
    Tissier, A. Plant secretory structures: More than just reaction bags. Curr. Opin. Biotechnol. 49, 73–79 (2018).
    Article  CAS  Google Scholar 

    12.
    Schilmiller, A. L., Last, R. L. & Pichersky, E. Harnessing plant trichome biochemistry for the production of useful compounds. Plant J. 54, 702–711 (2008).
    Article  CAS  Google Scholar 

    13.
    Wagner, G. J., Wang, E. & Shepherd, R. W. New approaches for studying and exploiting an old protuberance, the plant trichome. Ann. Bot. 93, 3–11 (2004).
    Article  CAS  Google Scholar 

    14.
    Lange, B. M. The evolution of plant secretory structures and the emergence of terpenoid chemical diversity. Annu. Rev. Plant Biol. 66, 139–159 (2015).
    Article  CAS  Google Scholar 

    15.
    Lange, B. M. et al. Probing essential oil biosynthesis and secretion by functional evaluation of expressed sequence tags from mint glandular trichomes. Proc. Natl. Acad. Sci. USA 97, 2934–2939 (2000).
    ADS  Article  CAS  Google Scholar 

    16.
    Wang, G. et al. Terpene biosynthesis in glandular trichomes of Hop. Plant Physiol. 148, 1254–1266 (2008).
    ADS  Article  CAS  Google Scholar 

    17.
    Nagel, J. et al. EST analysis of hop glandular trichomes identifies an o-methyltransferase that catalyzes the biosynthesis of xanthohumol. Plant Cell 20, 186–200 (2008).
    Article  CAS  Google Scholar 

    18.
    Czechowski, T. et al. Artemisia annua mutant impaired in Artemisinin synthesis demonstrates importance of nonenzymatic conversion in terpenoid metabolism. Proc. Natl. Acad. Sci. USA. 113, 15150–15155 (2016).
    Article  CAS  Google Scholar 

    19.
    Johnson, H. B. Plant pubescence: An ecological perspective. Bot. Rev. 41, 233–258 (1975).
    Article  Google Scholar 

    20.
    Fernald, M. L. Gray’s Manual of Botany. (American Book Company, 1950).

    21.
    Hernandez-Castillo, G. R., Stockey, R. A., Rothwell, G. W. & Mapes, G. Reconstructing Emporia lockardii (Voltziales: Emporiaceae) and initial thoughts on paleozoic conifer ecology. Int. J. Plant Sci. 170, 1056–1074 (2009).
    Article  Google Scholar 

    22.
    Rothwell, G. W., Mapes, G. & Hernandez-Castillo, G. R. Hanskerpia gen. nov. and phylogenetic relationships among the most ancient conifers (Voltziales). Taxon 54, 733–750 (2005).

    23.
    Heinrich, M. Das Harz der Nadelhölzer, seine Entstehung, Vertheilung, Bedeutung und Gewinnung (Springer, Berlin, 1894).
    Google Scholar 

    24.
    Tomlin, E. S., Antonejevic, E., Alfaro, R. I. & Borden, J. H. Changes in volatile terpene and diterpene resin acid composition of resistant and susceptible white spruce leaders exposed to simulated white pine weevil damage. Tree Physiol. 20, 1087–1095 (2000).
    Article  CAS  Google Scholar 

    25.
    Birol, I. et al. Assembling the 20 Gb white spruce (Picea glauca) genome from whole-genome shotgun sequencing data. Bioinformatics 29, 1492–1497 (2013).
    Article  CAS  Google Scholar 

    26.
    Warren, R. L. et al. Improved white spruce (Picea glauca) genome assemblies and annotation of large gene families of conifer terpenoid and phenolic defense metabolism. Plant J. 83, 189–212 (2015).
    Article  CAS  Google Scholar 

    27.
    Zulak, K. G., Dullat, H. K., Keeling, C. I., Lippert, D. & Bohlmann, J. Immunofluorescence localization of levopimaradiene/abietadiene synthase in methyl jasmonate treated stems of Sitka spruce (Picea sitchensis) shows activation of diterpenoid biosynthesis in cortical and developing traumatic resin ducts. Phytochemistry 71, 1695–1699 (2010).
    Article  CAS  Google Scholar 

    28.
    Whitehill, J. G. A., Henderson, H., Strong, W., Jaquish, B. & Bohlmann, J. Function of Sitka spruce stone cells as a physical defense against white pine weevil. Plant. Cell Environ. 39, 2545–2556 (2016).
    Article  CAS  Google Scholar 

    29.
    Franceschi, V. R., Krokene, P., Krekling, T. & Christiansen, E. Phloem parenchyma cells are involved in local and distant defense responses to fungal inoculation or bark-beetle attack in Norway spruce (Pinaceae). Am. J. Bot. 87, 314–326 (2000).
    Article  CAS  Google Scholar 

    30.
    Parent, G. J., Giguère, I., Mageroy, M., Bohlmann, J. & MacKay, J. J. Evolution of the biosynthesis of two hydroxyacetophenones in plants. Plant Cell Environ. 41, 620–629 (2018).
    Article  CAS  Google Scholar 

    31.
    Mageroy, M. H. et al. A conifer UDP-sugar dependent glycosyltransferase contributes to acetophenone metabolism and defense against insects. Plant Physiol. 175, 00611.2017 (2017).

    32.
    Tissier, A. Glandular trichomes: What comes after expressed sequence tags?. Plant J. 70, 51–68 (2012).
    Article  CAS  Google Scholar 

    33.
    Sacher, J. A. Structure and seasonal activity of the shoot apices of Pinus lambertiana and Pinus ponderosa. Am. J. Bot. 41, 749–759 (1954).
    Article  Google Scholar 

    34.
    De Simón, B. F., Vallejo, M. C. G., Cadahía, E., Miguel, C. A. & Martinez, M. C. Analysis of lipophilic compounds in needles of Pinus pinea L. Ann. For. Sci. 58, 449–454 (2001).
    Article  Google Scholar 

    35.
    Lange, W. & Weissman, G. Untersuchungen der Harzbalsame von Pinus resinosa Ait. und Pinus pinea L. Holz als Roh- und Werkst. 49, 476–480 (1991).

    36.
    Geisler, K., Jensen, N. B., Yuen, M. M. S., Madilao, L. & Bohlmann, J. Modularity of conifer diterpene resin acid biosynthesis: P450 enzymes of different CYP720B clades use alternative substrates and converge on the same products. Plant Physiol. 171, 152–164 (2016).
    Article  CAS  Google Scholar 

    37.
    Hamberger, B., Ohnishi, T., Hamberger, B., Séguin, A. & Bohlmann, J. Evolution of diterpene metabolism: Sitka spruce CYP720B4 catalyzes multiple oxidations in resin acid biosynthesis of conifer defense against insects. Plant Physiol. 157, 1677–1695 (2011).
    Article  CAS  Google Scholar 

    38.
    Ro, D., Arimura, G., Lau, S. Y. W., Piers, E. & Bohlmann, J. Loblolly pine abietadienol/abietadienal oxidase PtAO (CYP720B1) is a multifunctional, multisubstrate cytochrome P450 monooxygenase. Proc. Natl. Acad. Sci. USA 102, 8060–8065 (2005).
    ADS  Article  CAS  Google Scholar 

    39.
    Hilker, M., Stein, C., Schröder, R., Varama, M. & Mumm, R. Insect egg deposition induces defence responses in Pinus sylvestris: characterisation of the elicitor. J. Exp. Biol. 208, 1849–1854 (2005).
    Article  Google Scholar 

    40.
    Schuurink, R. & Tissier, A. Glandular trichomes: Micro-organs with model status?. New Phytol. https://doi.org/10.1111/nph.16283 (2019).
    Article  PubMed  Google Scholar 

    41.
    Huchelmann, A., Boutry, M. & Hachez, C. Plant glandular trichomes: Natural cell factories of high biotechnological interest. Plant Physiol. 175, 00727.2017 (2017).

    42.
    Sallaud, C. et al. Characterization of two genes for the biosynthesis of the labdane diterpene Z-abienol in tobacco (Nicotiana tabacum) glandular trichomes. Plant J. 72, 1–17 (2012).
    Article  CAS  Google Scholar 

    43.
    Liu, Y. et al. A geranylfarnesyl diphosphate synthase provides the precursor for sesterterpenoid (C25) formation in the glandular trichomes of the mint species Leucosceptrum canum. Plant Cell 28, 804–822 (2016).
    Article  CAS  Google Scholar 

    44.
    Reynolds, E. S. The use of lead citrate at high pH as an electron-opaque stain in electron microscopy. J. Cell Biol. 17, 208–212 (1963).
    Article  CAS  Google Scholar  More

  • in

    Changes in the vaginal microbiota across a gradient of urbanization

    1.
    Huttenhower, C. Structure, function and diversity of the healthy human microbiome. Nature 486, 207–214 (2012).
    ADS  CAS  Google Scholar 
    2.
    Nugent, R. P., Krohn, M. A. & Hillier, S. L. Reliability of diagnosing bacterial vaginosis is improved by a standardized method of gram stain interpretation. J Clin Microbiol 29, 297–301 (1991).
    CAS  PubMed  PubMed Central  Google Scholar 

    3.
    Amsel, R. et al. Nonspecific vaginitis: diagnostic criteria and microbial and epidemiologic associations. Am. J. Med. 74, 14–22 (1983).
    CAS  PubMed  Google Scholar 

    4.
    Boris, S. & Barbes, C. Role played by lactobacilli in controlling the population of vaginal pathogens. Microbes Infect 2, 543–546 (2000).
    CAS  PubMed  Google Scholar 

    5.
    Tyssen, D. et al. Anti-HIV-1 Activity of Lactic Acid in Human Cervicovaginal Fluid. mSphere 3, e00055–00018 (2018).

    6.
    Petrova, M. I., Reid, G., Vaneechoutte, M. & Lebeer, S. Lactobacillus iners: friend or foe?. Trends Microbiol. 25, 182–191 (2017).
    CAS  PubMed  Google Scholar 

    7.
    Schwebke, J. R. & Desmond, R. Natural history of asymptomatic bacterial vaginosis in a high-risk group of women. Sex. Transm. Dis. 34, 876–877 (2007).
    PubMed  Google Scholar 

    8.
    Larsson, P.-G., Platz-Christensen, J.-J., Thejls, H., Forsum, U. & Påhlson, C. Incidence of pelvic inflammatory disease after first-trimester legal abortion in women with bacterial vaginosis after treatment with metronidazole: a double-blind, randomized study. Am. J. Obstet. Gynecol. 166, 100–103 (1992).
    CAS  PubMed  Google Scholar 

    9.
    Hillier, S. L. et al. Association between bacterial vaginosis and preterm delivery of a low-birth-weight infant. N. Engl. J. Med. 333, 1737–1742 (1995).
    CAS  PubMed  Google Scholar 

    10.
    Donders, G. et al. Predictive value for preterm birth of abnormal vaginal flora, bacterial vaginosis and aerobic vaginitis during the first trimester of pregnancy. BJOG 116, 1315–1324 (2009).
    CAS  PubMed  Google Scholar 

    11.
    Borgdorff, H. et al. Lactobacillus-dominated cervicovaginal microbiota associated with reduced HIV/STI prevalence and genital HIV viral load in African women. ISME J. 8, 1781 (2014).
    CAS  PubMed  PubMed Central  Google Scholar 

    12.
    Cherpes, T. L., Hillier, S. L., Meyn, L. A., Busch, J. L. & Krohn, M. A. A delicate balance: risk factors for acquisition of bacterial vaginosis include sexual activity, absence of hydrogen peroxide-producing lactobacilli, black race, and positive herpes simplex virus type 2 serology. Sex. Transm. Dis. 35, 78–83 (2008).
    PubMed  Google Scholar 

    13.
    Dareng, E. O. et al. Prevalent high-risk HPV infection and vaginal microbiota in Nigerian women. Epidemiol. Infect. 144, 123–137 (2016).
    CAS  PubMed  Google Scholar 

    14.
    Brotman, R. M. et al. Interplay between the temporal dynamics of the vaginal microbiota and human papillomavirus detection. J. Infect. Dis. 210, 1723–1733 (2014).
    CAS  PubMed  PubMed Central  Google Scholar 

    15.
    Lee, J. E. et al. Association of the vaginal microbiota with human papillomavirus infection in a Korean twin cohort. PLoS ONE 8, e63514–e63514 (2013).
    ADS  PubMed  PubMed Central  Google Scholar 

    16.
    Rodriguez-Cerdeira, C., Sanchez-Blanco, E. & Alba, A. Evaluation of association between vaginal infections and high-risk human papillomavirus types in female sex workers in Spain. ISRN Obstet. Gynecol. https://doi.org/10.5402/2012/240190 (2012).
    Article  PubMed  PubMed Central  Google Scholar 

    17.
    Di Paola, M. et al. Characterization of cervico-vaginal microbiota in women developing persistent high-risk Human Papillomavirus infection. Sci. Rep. 7, 10200 (2017).
    PubMed  PubMed Central  Google Scholar 

    18.
    Cauci, S. Vaginal immunity in bacterial vaginosis. Curr. Infect. Dis. Rep. 6, 450–456 (2004).
    PubMed  Google Scholar 

    19.
    Stokholm, J. et al. Antibiotic use during pregnancy alters the commensal vaginal microbiota. Clin. Microbiol. Infect. 20, 629–635 (2014).
    CAS  PubMed  Google Scholar 

    20.
    Welch, J. S. Quantitative and qualitative effects of douche preparations on vaginal microflora. Obstet. Gynecol. 81, 320–321 (1993).
    CAS  PubMed  Google Scholar 

    21.
    Brotman, R. M. et al. The effect of vaginal douching cessation on bacterial vaginosis: a pilot study. Am. J. Obstet. Gynecol. 198(628), e621-628 (2008).
    Google Scholar 

    22.
    Schwebke, J. R., Richey, C. M. & Weiss, H. L. Correlation of behaviors with microbiological changes in vaginal flora. J. Infect. Dis. 180, 1632–1636 (1999).
    CAS  PubMed  Google Scholar 

    23.
    Beigi, R. H., Wiesenfeld, H. C., Hillier, S. L., Straw, T. & Krohn, M. A. Factors associated with absence of H2O2-producing Lactobacillus among women with bacterial vaginosis. J. Infect. Dis. 191, 924–929 (2005).
    PubMed  Google Scholar 

    24.
    Ahluwalia, N. & Grandjean, H. Nutrition, an under-recognized factor in bacterial vaginosis. J. Nutr. 137, 1997–1998 (2007).
    CAS  PubMed  Google Scholar 

    25.
    Neggers, Y. H. et al. Dietary intake of selected nutrients affects bacterial vaginosis in women. J. Nutr. 137, 2128–2133 (2007).
    CAS  PubMed  PubMed Central  Google Scholar 

    26.
    Brotman, R. M. et al. Association between cigarette smoking and the vaginal microbiota: a pilot study. BMC Infect. Dis. 14, 471 (2014).
    PubMed  PubMed Central  Google Scholar 

    27.
    van Houdt, R. et al. Lactobacillus iners-dominated vaginal microbiota is associated with increased susceptibility to Chlamydia trachomatis infection in Dutch women: a case–control study. Sex. Transm. Infect. 94, 117–123 (2017).
    PubMed  PubMed Central  Google Scholar 

    28.
    Noyes, N., Cho, K.-C., Ravel, J., Forney, L. J. & Abdo, Z. Associations between sexual habits, menstrual hygiene practices, demographics and the vaginal microbiome as revealed by Bayesian network analysis. PLoS ONE 13, e0191625 (2018).
    PubMed  PubMed Central  Google Scholar 

    29.
    Ravel, J. et al. Vaginal microbiome of reproductive-age women. Proc. Natl. Acad. Sci. USA 108(Suppl 1), 4680–4687 (2011).
    ADS  CAS  PubMed  Google Scholar 

    30.
    Zhou, X. et al. Differences in the composition of vaginal microbial communities found in healthy Caucasian and black women. ISME J. 1, 121–133 (2007).
    CAS  PubMed  Google Scholar 

    31.
    Zhou, X. et al. The vaginal bacterial communities of Japanese women resemble those of women in other racial groups. FEMS Immunol. Med. Microbiol. 58, 169–181 (2010).
    CAS  PubMed  Google Scholar 

    32.
    Borgdorff, H. et al. The association between ethnicity and vaginal microbiota composition in Amsterdam, the Netherlands. PLoS ONE 12, e0181135 (2017).
    PubMed  PubMed Central  Google Scholar 

    33.
    Verstraelen, H. et al. Characterisation of the human uterine microbiome in non-pregnant women through deep sequencing of the V1–2 region of the 16S rRNA gene. PeerJ 4, e1602 (2016).
    PubMed  PubMed Central  Google Scholar 

    34.
    Rick, A.-M. et al. Open forum infectious diseases. (Oxford University Press, Oxford).

    35.
    Anukam, K. C., Osazuwa, E. O., Ahonkhai, I. & Reid, G. Lactobacillus vaginal microbiota of women attending a reproductive health care service in Benin city, Nigeria. Sex. Transm. Dis. 33, 59–62 (2006).
    PubMed  Google Scholar 

    36.
    Bourgeon, L., Burke, A. & Higham, T. Earliest human presence in North America Dated to the Last Glacial Maximum: new radiocarbon dates from Bluefish Caves, Canada. PLoS ONE 12, e0169486 (2017).
    PubMed  PubMed Central  Google Scholar 

    37.
    Bortolini, M. C. et al. Y-chromosome evidence for differing ancient demographic histories in the Americas. Am. J. Hum. Genet. 73, 524–539 (2003).
    CAS  PubMed  PubMed Central  Google Scholar 

    38.
    Hurtado, A. M., Hurtado, I. & Hill, K. Public health and adaptive immunity among natives of South America. Lost Paradises and the Ethics of Research and Publication 164–192 (Oxford University Press, New York, 2003).

    39.
    Lindenau, J. et al. Distribution patterns of variability for 18 immune system genes in Amerindians–relationship with history and epidemiology. HLA 82, 177–185 (2013).
    CAS  Google Scholar 

    40.
    Bhatia, K. K., Black, F. L., Smith, T. A., Prasad, M. L. & Koki, G. N. Class I HLA antigens in two long-separated populations: Melanesians and South Amerinds. Am. J. Phys. Anthropol. 97, 291–305 (1995).
    CAS  PubMed  Google Scholar 

    41.
    Watkins, D. I. et al. New recombinant HLA-B alleles in a tribe of South American Amerindians indicate rapid evolution of MHC class I loci. Nature 357, 329–333 (1992).
    ADS  CAS  PubMed  Google Scholar 

    42.
    Ewerton, P. D., de Meira Leite, M., Magalhães, M., Sena, L. & dos Santos, E. J. M. Amazonian Amerindians exhibit high variability of KIR profiles. Immunogenetics 59, 625–630 (2007).
    CAS  PubMed  Google Scholar 

    43.
    Freire, G. & Tillett, A. Salud indígena en Venezuela. First volume. (Dirección de Salud Indígena, 2007).

    44.
    Contreras, M. et al. The bacterial microbiota in the oral mucosa of rural Amerindians. Microbiology 156, 3282–3287 (2010).
    CAS  PubMed  Google Scholar 

    45.
    Clemente, J. C. et al. The microbiome of uncontacted Amerindians. Sci. Adv. 1, e1500183 (2015).
    ADS  PubMed  PubMed Central  Google Scholar 

    46.
    Blaser, M. J. et al. Distinct cutaneous bacterial assemblages in a sampling of South American Amerindians and US residents. ISME J. 7, 86–95 (2012).
    Google Scholar 

    47.
    Yatsunenko, T. et al. Human gut microbiome viewed across age and geography. Nature 486, 222–222 (2012).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    48.
    Godinho, N. M. D. O. O Impacto das Migrações na Constituição Genética de Populações Latino-Americanas. (2008).

    49.
    Jespers, V. et al. The significance of Lactobacillus crispatus and L. vaginalis for vaginal health and the negative effect of recent sex: a cross-sectional descriptive study across groups of African women. BMC Infect. Dis. 15, 115 (2015).
    PubMed  PubMed Central  Google Scholar 

    50.
    Gosmann, C. et al. Lactobacillus-deficient cervicovaginal bacterial communities are associated with increased HIV acquisition in Young South African Women. Immunity 46, 29–37 (2017).
    CAS  PubMed  PubMed Central  Google Scholar 

    51.
    Godoy-Vitorino, F. et al. Cervicovaginal fungi and bacteria associated with cervical intraepithelial neoplasia and high-risk Human Papillomavirus infections in a Hispanic population. Front. Microbiol. 9, 2533 (2018).
    PubMed  PubMed Central  Google Scholar 

    52.
    Tarnberg, M., Jakobsson, T., Jonasson, J. & Forsum, U. Identification of randomly selected colonies of lactobacilli from normal vaginal fluid by pyrosequencing of the 16S rDNA variable V1 and V3 regions. APMIS 110, 802–810 (2002).
    PubMed  Google Scholar 

    53.
    Huse, S. M., Ye, Y., Zhou, Y. & Fodor, A. A. A core human microbiome as viewed through 16S rRNA sequence clusters. PLoS ONE 7, e34242 (2012).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    54.
    Shen, J. et al. Effects of low dose estrogen therapy on the vaginal microbiomes of women with atrophic vaginitis. Sci. Rep. 6, 24380 (2016).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    55.
    Witkin, S. S. et al. Vaginal biomarkers that predict cervical length and dominant bacteria in the vaginal microbiomes of pregnant women. mBio 10, e02242-129 (2019).
    Google Scholar 

    56.
    Ravel, J. et al. Daily temporal dynamics of vaginal microbiota before, during and after episodes of bacterial vaginosis. Microbiome 1, 29 (2013).
    PubMed  PubMed Central  Google Scholar 

    57.
    Smith, B. C. et al. The cervical microbiome over 7 years and a comparison of methodologies for its characterization. PLoS ONE 7, e40425 (2012).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    58.
    Gajer, P. et al. Temporal dynamics of the human vaginal microbiota. Sci. Transl. Med. 4, 132ra152 (2012).
    Google Scholar 

    59.
    Kim, T. K. et al. Heterogeneity of vaginal microbial communities within individuals. J. Clin. Microbiol. 47, 1181–1189 (2009).
    CAS  PubMed  PubMed Central  Google Scholar 

    60.
    Liu, M.-B. et al. Diverse vaginal microbiomes in reproductive-age women with vulvovaginal candidiasis. PLoS ONE 8, e79812 (2013).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    61.
    Verstraelen, H. et al. Longitudinal analysis of the vaginal microflora in pregnancy suggests that L. crispatus promotes the stability of the normal vaginal microflora and that L. gasseri and/or L. iners are more conducive to the occurrence of abnormal vaginal microflora. BMC Microbiol. 9, 116 (2009).
    PubMed  PubMed Central  Google Scholar 

    62.
    Byrne, E. H. et al. Cervicovaginal bacteria are a major modulator of host inflammatory responses in the female genital tract. Immunity 42, 965–976 (2015).
    PubMed  PubMed Central  Google Scholar 

    63.
    Mach, N. & Fuster-Botella, D. Endurance exercise and gut microbiota: a review. J. Sport Health Sci. 6, 179–197 (2017).
    PubMed  Google Scholar 

    64.
    Tasnim, N., Abulizi, N., Pither, J., Hart, M. M. & Gibson, D. L. Linking the gut microbial ecosystem with the environment: does gut health depend on where we live?. Front. Microbiol. 8, 1935 (2017).
    PubMed  PubMed Central  Google Scholar 

    65.
    Collins, M. D. The Prokaryotes 1013–1019 (Springer, Berlin, 2006).
    Google Scholar 

    66.
    Freitas, A. C. et al. Increased richness and diversity of the vaginal microbiota and spontaneous preterm birth. Microbiome 6, 117 (2018).
    PubMed  PubMed Central  Google Scholar 

    67.
    Marrazzo, J. M., Thomas, K. K., Fiedler, T. L., Ringwood, K. & Fredricks, D. N. Relationship of specific vaginal bacteria and bacterial vaginosis treatment failure in women who have sex with women. Ann. Intern. Med. 149, 20–28 (2008).
    PubMed  PubMed Central  Google Scholar 

    68.
    Si, J., You, H. J., Yu, J., Sung, J. & Ko, G. Prevotella as a hub for vaginal microbiota under the influence of host genetics and their association with obesity. Cell Host Microbe 21, 97–105 (2017).
    CAS  PubMed  Google Scholar 

    69.
    Srinivasan, S. et al. Bacterial communities in women with bacterial vaginosis: high resolution phylogenetic analyses reveal relationships of microbiota to clinical criteria. PLoS ONE 7, e37818 (2012).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    70.
    Yeoman, C. J. et al. A multi-omic systems-based approach reveals metabolic markers of bacterial vaginosis and insight into the disease. PLoS ONE 8, e56111 (2013).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    71.
    Oakley, B. B., Fiedler, T. L., Marrazzo, J. M. & Fredricks, D. N. Diversity of human vaginal bacterial communities and associations with clinically defined bacterial vaginosis. Appl. Environ. Microbiol. 74, 4898–4909 (2008).
    CAS  PubMed  PubMed Central  Google Scholar 

    72.
    Machado, A. & Cerca, N. Influence of biofilm formation by gardnerella vaginalis and other anaerobes on bacterial vaginosis. J. Infect. Dis. 212, 1856–1861 (2015).
    CAS  PubMed  Google Scholar 

    73.
    Vargas-Robles, D. et al. High rate of infection by only oncogenic human papillomavirus in Amerindians. mSphere 3, e00176-e118 (2018).
    PubMed  PubMed Central  Google Scholar 

    74.
    Song, D., Li, H., Li, H. & Dai, J. Effect of human papillomavirus infection on the immune system and its role in the course of cervical cancer. Oncol. Lett. 10, 600–606 (2015).
    CAS  PubMed  PubMed Central  Google Scholar 

    75.
    Briselden, A. M., Moncla, B. J., Stevens, C. E. & Hillier, S. L. Sialidases (Neuraminidases) in bacterial vaginosis and bacterial vaginosis-associated microflora. J. Clin. Microbiol. 30, 663–666 (1992).
    CAS  PubMed  PubMed Central  Google Scholar 

    76.
    Gillet, E. et al. Bacterial vaginosis is associated with uterine cervical human papillomavirus infection: a meta-analysis. BMC Infect. Dis. 11, 10 (2011).
    PubMed  PubMed Central  Google Scholar 

    77.
    Discacciati, M. G. et al. Presence of 20% or more clue cells: an accurate criterion for the diagnosis of bacterial vaginosis in Papanicolaou cervical smears. Diagn. Cytopathol. 34, 272–276 (2006).
    PubMed  Google Scholar 

    78.
    World Health Organization. Haemoglobin Concentrations for the Diagnosis of Anaemia and Assessment of Severity (World Health Organization, Geneva, 2011).
    Google Scholar 

    79.
    Magoč, T. & Salzberg, S. L. FLASH: fast length adjustment of short reads to improve genome assemblies. Bioinformatics 27, 2957–2963 (2011).
    PubMed  PubMed Central  Google Scholar 

    80.
    Callahan, B. J. et al. DADA2: High resolution sample inference from Illumina amplicon data. Nat. Methods 13, 581–583 (2016).
    CAS  PubMed  PubMed Central  Google Scholar 

    81.
    Allard, G., Ryan, F. J., Jeffery, I. B. & Claesson, M. J. SPINGO: a rapid species-classifier for microbial amplicon sequences. BMC Bioinform. 16, 324 (2015).
    Google Scholar 

    82.
    Cole, J. R. et al. Ribosomal Database Project: data and tools for high throughput rRNA analysis. Nucleic Acids Res. 42, D633–D642 (2013).
    PubMed  PubMed Central  Google Scholar 

    83.
    Morgulis, A. et al. Database indexing for production MegaBLAST searches. Bioinformatics 24, 1757–1764 (2008).
    CAS  PubMed  PubMed Central  Google Scholar 

    84.
    Muzny, C. A. et al. Characterization of the vaginal microbiota among sexual risk behavior groups of women with bacterial vaginosis. PLoS ONE 8, e80254 (2013).
    ADS  PubMed  PubMed Central  Google Scholar 

    85.
    Caporaso, J. G. et al. QIIME allows analysis of high-throughput community sequencing data. Nat. Methods 7, 335–336 (2010).
    CAS  PubMed  PubMed Central  Google Scholar 

    86.
    R: A language and environment for statistical computing (R Foundation for Statistical Computing, Vienna, Austria, 2016).

    87.
    Oksanen, J. et al. vegan: Community Ecology Package. R package version 2.4-4. R Development Core Team. R: A language and environment for statistical computing. Vienna: R Foundation for Statistical Computing (2010).

    88.
    Anderson, M. J. Permutation tests for univariate or multivariate analysis of variance and regression. Can. J. Fish. Aquat. Sci. 58, 626–639 (2001).
    Google Scholar 

    89.
    Bates, D., Maechler, M., Bolker, B. & Walker, S. Fitting Linear Mixed-Effects Models Using lme4. (2015).

    90.
    textmineR: Functions for Text Mining and Topic Modeling v. 2.0.6 (2017).

    91.
    Rousseeuw, P. J. Silhouettes: a graphical aid to the interpretation and validation of cluster analysis. J. Comput. Appl. Math. 20, 53–65 (1987).
    MATH  Google Scholar 

    92.
    Walther, R. T. A. G. Cluster validation by prediction strength. J. Comput. Graph. Stat. 14, 511–528 (2005).
    MathSciNet  Google Scholar 

    93.
    Philentropy: Information Theory and Distance Quantification with R (2018).

    94.
    Hsieh, T. C., Ma, K. H. & Chao, A. iNEXT: an R package for rarefaction and extrapolation of species diversity (Hill numbers). Methods Ecol Evol 7, 1451–1456 (2016).
    Google Scholar 

    95.
    Cohen, J. A coefficient of agreement for nominal scales. Educ. Psychol. Meas. 20, 37–46 (1960).
    Google Scholar 

    96.
    Viera, A. J. & Garrett, J. M. Understanding interobserver agreement: the kappa statistic. Fam. Med. 37, 360–363 (2005).
    PubMed  Google Scholar 

    97.
    Segata, N. et al. Metagenomic biomarker discovery and explanation. Genome Biol. 12, R60 (2011).
    PubMed  PubMed Central  Google Scholar 

    98.
    Nakazawa, M. fmsb: Functions for medical statistics book with some demographic data. R package version 0.4 (2014).

    99.
    Benjamini, Y. & Hochberg, Y. Controlling the false discovery rate: a practical and powerful approach to multiple testing. J. R. Stat. Soc. B 57, 289–300 (1995).
    MathSciNet  MATH  Google Scholar 

    100.
    Wickham, H. Reshaping Data with the reshape Package. J. Stat. Softw. 21, 1–20 (2007).
    Google Scholar 

    101.
    gplots: Various R Programming Tools for Plotting Data (R, 2016).

    102.
    Vázquez-Baeza, Y., Pirrung, M., Gonzalez, A. & Knight, R. EMPeror: a tool for visualizing high-throughput microbial community data. Gigascience 2, 16 (2013).
    PubMed  PubMed Central  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

  • in

    Climate and atmospheric deposition effects on forest water-use efficiency and nitrogen availability across Britain

    Site and sampling
    We selected twelve monoculture tree stands of the most common tree species in Britain, Scots pine (Pinus sylvestris L.), Sitka spruce (Picea sitchensis Bong. Carr.), pedunculate oak (Quercus robur L.) and common beech (Fagus sylvatica L.). The majority of the stands were experimental sites within the Level II- ICP intensive forest monitoring network (http://icp-forests.net/), with the exception of Covert Wood, Shobdon and Goyt. The Goyt site was added as a high Ndep site as a contrast to the low Ndep Sitka spruce site in Scotland (Fig. 1, Table 1, Supplementary Table 1). For each species, forests were selected with similar soil type and age, but with contrasting Ndep, Sdep and climate, particularly rainfall and temperature, as described in Fig. 1, Table 1 and Supplementary Table 1. Stand information (mean tree height, mean diameter at the breast height and basal area) as measured for target years and for some of the forest stands are shown in Fig. S4.
    At each ICP forest site, a plot of 0.25 ha was established in 1995 to carry out monitoring30 and a similar protocol was followed at the Goyt and Shobdon sites. Within each plot, 10 trees were selected for the collection of 3 wood cores per tree by using a 5 mm diameter increment borer, which were placed in paper straws for transport. Sampling was carried out between November 2010 and March 2011. Once in the laboratory, samples were dried at 70 °C for 48 h. Of the three wood cores sampled, one was kept for dendrochronology, with the other two used for stable isotope analyses.
    Climate and atmospheric deposition data
    Climate data (Temperature, T, Vapour Pressure Deficit, VPD, Precipitation, P) were obtained from automated weather stations at the sites and/or the nearest local meteorological stations (data were provided by the British Atmospheric Data Centre). Annual mean (Ta) and mean maximum (Tamax) values for temperature were calculated from monthly mean and maximum air temperature, T, respectively, and annual precipitation (Pa) was calculated as the sum of total monthly precipitations. Annual VPD was calculated from averaging monthly values obtained from mean monthly maximum temperature and minimum monthly relative humidity. For all the parameters, mean values were also calculated over the growing season, i.e., from May to September. We also considered the standardised precipitation-evapo-transpiration index, SPEI, relative to August, with 1 (SPEI8_1), 2 (SPEI8_2) and 3 (SPEI8_3) months time-scale and SPEI relative to December, with 1 and 12 months time-scale, the latter providing year-cumulated soil moisture conditions. SPEI values were obtained from the global database with 0.5 degrees spatial resolution available online at: https://sac.csic.es/spei/.
    Yearly wet nitrogen (Ndep) and sulphur deposition (Sdep) were obtained from measured bulk precipitation and throughfall water volumes at the sites and measured elemental concentrations (NO3−, NH4+ and SO2–4) as previously described30. For the spatial analyses, we considered mean of annual deposition (sNdep and sSdep), obtained as the sum of total (NH4-N + NO3-N for Ndep) wet and dry deposition. The latter were estimated as difference between throughfall and bulk precipitation N fluxes30. For Rogate only 1 year (2010) of monitoring was available. For Goyt site, atmospheric deposition data collected at Ladybower were considered, as the two sites are 30 km apart. For two sites (i.e., Shobdon and Covert Wood), which were not part of the regular ICP forest sites, the wet deposition obtained from the UK 5 × 5 km grid Ndep and Sdep dataset was used4. The estimate included wet and dry NHx-N (NH4, NH3), NOy-N (NO2, NO3, HNO3) and S (SOx = SO2 and SO4) deposition, modelled using FRAME with 2005 emissions data4. However, only the total wet deposition was included in the analyses here, as we previously reported a good agreement between modelled and measured wet Ndep50.
    For the temporal analyses, only wet deposition (as calculated from NO3−, NH4+ and SO2–4 concentrations in bulk precipitation) was considered (indicated as aNdep and aSdep), given the uncertainties associated with the quantification of the dry deposition. For instance, when differences between throughfall and bulk precipitation are  More

  • in

    Elevational is the main factor controlling the soil microbial community structure in alpine tundra of the Changbai Mountain

    1.
    Ruan, H., Zou, X., Scatena, F. & Zimmerman, J. Asynchronous fluctuation of soil microbial biomass and plant litterfall in a tropical wet forest. Plant Soil 260, 147–154 (2004).
    CAS  Google Scholar 
    2.
    Ibekwe, A. M. et al. Impact of fumigants on soil microbial communities. Appl. Environ. Microbiol. 67, 3245–3257 (2001).
    CAS  PubMed  PubMed Central  Google Scholar 

    3.
    Reiss, J., Bridle, J. R., Montoya, J. M. & Woodward, G. Emerging horizons in biodiversity and ecosystem functioning research. Trends Ecol. Evol. 24, 505–514 (2009).
    PubMed  Google Scholar 

    4.
    Fuhrman, J. A. & Fuhrman, J. A. Microbial community structure and its functional implications. Nature 459, 193–199 (2009).
    ADS  CAS  PubMed  Google Scholar 

    5.
    Lanzen, A. et al. Multi-targeted metagenetic analysis of the influence of climate and environmental parameters on soil microbial communities along an elevational gradient. Sci. Rep. 6, 28257. https://doi.org/10.1038/srep28257 (2016).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    6.
    Berendsen, R. L., Pieterse, C. M. J. & Bakker, P. A. H. M. The rhizosphere microbiome and plant health. Trends Plant Sci. 17, 478–486 (2012).
    CAS  PubMed  Google Scholar 

    7.
    Oh, Y. M. et al. Distinctive bacterial communities in the rhizoplane of four tropical tree species. Microb. Ecol. 64, 1018–1027 (2012).
    PubMed  Google Scholar 

    8.
    Walker, T. S., Bais, H. P., Grotewold, E. & Vivanco, J. M. Root exudation and rhizosphere biology. Plant Physiol. 132, 44–51 (2003).
    CAS  PubMed  PubMed Central  Google Scholar 

    9.
    Herre, E. A. Negative plant-soil feedback predicts tree-species relative abundance in a tropical forest. Nature 466, 752–755 (2010).
    ADS  PubMed  Google Scholar 

    10.
    Philippot, L., Raaijmakers, J. M., Lemanceau, P. & Wh, V. D. P. Going back to the roots: The microbial ecology of the rhizosphere. Nat. Rev. Microbiol. 11, 789–799 (2013).
    CAS  PubMed  Google Scholar 

    11.
    Hackl, E., Pfeffer, M., Donat, C., Bachmann, G. & Zechmeister-Boltenstern, S. Composition of the microbial communities in the mineral soil under different types of natural forest. Soil Biol. Biochem. 37, 661–671 (2005).
    CAS  Google Scholar 

    12.
    Park, S. et al. Principal component analysis and discriminant analysis (PCA–DA) for discriminating profiles of terminal restriction fragment length polymorphism (T-RFLP) in soil bacterial communities. Soil Biol. Biochem. 38, 2344–2349 (2006).
    CAS  Google Scholar 

    13.
    Mendes, R. et al. Deciphering the rhizosphere microbiome for disease-suppressive bacteria. Science 332, 1097–1100 (2011).
    ADS  CAS  PubMed  Google Scholar 

    14.
    Klimeš, L. Alpine plant life. Functional plant ecology of high mountain ecosystems by C. Körner. Folia Geobot. 41, 454–455 (2006).
    Google Scholar 

    15.
    Diaz, H. F., Grosjean, M. & Graumlich, L. Climate variability and change in high elevation regions: Past, present and future. Clim. Change 59, 1–4 (2003).
    Google Scholar 

    16.
    Schinner, F. & Gstraunthaler, G. Adaptation of microbial activities to the environmental conditions in alpine soils. Oecologia 50, 113–116 (1981).
    ADS  PubMed  Google Scholar 

    17.
    Cui, H.-J. et al. Soil microbial community composition and its driving factors in alpine grasslands along a mountain elevational gradient. J. Mt. Sci. 13, 1013–1023. https://doi.org/10.1007/s11629-015-3614-7 (2016).
    Article  Google Scholar 

    18.
    Zhang, B., Liang, C., He, H. & Zhang, X. Variations in soil microbial communities and residues along an altitude gradient on the northern slope of changbai mountain, china. PLoS ONE 8, e66184 (2013).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    19.
    Collins, C. G., Carey, C. J., Aronson, E. L., Kopp, C. W. & Diez, J. M. Direct and indirect effects of native range expansion on soil microbial community structure and function. J. Ecol. 104, 1271–1283 (2016).
    Google Scholar 

    20.
    Margesin, R., Jud, M., Tscherko, D. & Schinner, F. Microbial communities and activities in alpine and subalpine soils. FEMS Microbiol. Ecol. 67, 208–218 (2009).
    CAS  PubMed  Google Scholar 

    21.
    Fierer, N. & Mcculley, R. L. Reconstructing the microbial diversity and function of pre-agricultural tallgrass prairie soils in the United States. Science 342, 621–624 (2013).
    ADS  CAS  PubMed  Google Scholar 

    22.
    Shi, Y. et al. Multi-scale variability analysis reveals the importance of spatial distance in shaping Arctic soil microbial functional communities. Soil Biol. Biochem. 86, 126–134 (2015).
    CAS  Google Scholar 

    23.
    Yang, Y. et al. The microbial gene diversity along an elevation gradient of the Tibetan grassland. Isme J. Multidiscipl. J. Microb. Ecol. 8, 430–440 (2013).
    Google Scholar 

    24.
    Ding, J. et al. Integrated metagenomics and network analysis of soil microbial community of the forest timberline. Sci. Rep. 5, 7994 (2015).
    CAS  PubMed  PubMed Central  Google Scholar 

    25.
    Liu, J. et al. High throughput sequencing analysis of biogeographical distribution of bacterial communities in the black soils of northeast China. Soil Biol. Biochem. 70, 113–122 (2014).
    CAS  Google Scholar 

    26.
    Brockett, B. F. T., Prescott, C. E. & Grayston, S. J. Soil moisture is the major factor influencing microbial community structure and enzyme activities across seven biogeoclimatic zones in western Canada. Soil Biol. Biochem. 44, 9–20 (2012).
    CAS  Google Scholar 

    27.
    Uroz, S., Tech, J. J., Sawaya, N. A., Frey-Klett, P. & Leveau, J. H. J. Structure and function of bacterial communities in ageing soils: Insights from the Mendocino ecological staircase. Soil Biol. Biochem. 69, 265–274 (2014).
    CAS  Google Scholar 

    28.
    Shi, Y. et al. Vegetation-associated impacts on arctic tundra bacterial and microeukaryotic communities. Appl. Environ. Microbiol. 81, 492–501 (2014).
    PubMed  Google Scholar 

    29.
    Zeglin, L. H. & Myrold, D. D. Fate of decomposed fungal cell wall material in organic horizons of old-growth douglas-fir forest soils. Soil Sci. Soc. Am. J. 77, 489–500 (2013).
    ADS  CAS  Google Scholar 

    30.
    Wallander, H., Göransson, H. & Rosengren, U. Production, standing biomass and natural abundance of 15N and 13C in ectomycorrhizal mycelia collected at different soil depths in two forest types. Oecologia 139, 89–97 (2004).
    ADS  PubMed  Google Scholar 

    31.
    Colpaert, J. V., Laere, A. V. & Assche, J. A. V. Carbon and nitrogen allocation in ectomycorrhizal and non-mycorrhizal Pinus sylvestris L. seedlings. Tree Physiol. 16, 787–793 (1996).
    CAS  PubMed  Google Scholar 

    32.
    Zhang, M. et al. Distribution of soil organic carbon fractions along the altitudinal gradient in Changbai mountain, China. Pedosphere 21, 615–620 (2011).
    CAS  Google Scholar 

    33.
    Wenduo, X. U., Xingyuan, H. E., Chen, W. & Liu, C. Characteristics and succession rules of vegetation types in Changbai mountain. Chin. J. Ecol. 23, 162–174 (2004).
    Google Scholar 

    34.
    Mao, Y., Yannarell, A. C., Davis, S. C. & Mackie, R. I. Impact of different bioenergy crops on N-cycling bacterial and archaeal communities in soil. Environ. Microbiol. 15, 928–942 (2012).
    PubMed  Google Scholar 

    35.
    Grayston, S. J. et al. Assessing shifts in microbial community structure across a range of grasslands of differing management intensity using CLPP, PLFA and community DNA techniques. Appl. Soil. Ecol. 25, 63–84 (2004).
    Google Scholar 

    36.
    Lauber, C. L., Hamady, M., Knight, R. & Fierer, N. Pyrosequencing-based assessment of soil pH as a predictor of soil bacterial community structure at the continental scale. Appl. Environ. Microbiol. 75, 5111–5120 (2009).
    CAS  PubMed  PubMed Central  Google Scholar 

    37.
    Tedersoo, L. et al. Towards global patterns in the diversity and community structure of ectomycorrhizal fungi. Mol. Ecol. 21, 4160–4170 (2012).
    PubMed  Google Scholar 

    38.
    Davey, M. L., Heegaard, E., Halvorsen, R., Kauserud, H. & Ohlson, M. Amplicon-pyrosequencing-based detection of compositional shifts in bryophyte-associated fungal communities along an elevation gradient. Mol. Ecol. 22, 368–383 (2013).
    CAS  PubMed  Google Scholar 

    39.
    Mao, Y., Yannarell, A. C. & Mackie, R. I. Changes in N-transforming archaea and bacteria in soil during the establishment of bioenergy crops. PLoS ONE 6, e24750 (2011).
    ADS  CAS  PubMed  PubMed Central  Google Scholar 

    40.
    Heijden, M. G. A. V. D., Bardgett, R. D. & Straalen, N. M. V. The unseen majority: Soil microbes as drivers of plant diversity and productivity in terrestrial ecosystems. Ecol. Lett. 11, 296–310 (2008).
    PubMed  Google Scholar 

    41.
    Steltzer, H. & Bowman, W. D. Original articles: Differential influence of plant species on soil nitrogen transformations within moist meadow alpine Tundra. Ecosystems 1, 464–474 (1998).
    CAS  Google Scholar 

    42.
    Shen, C., Ni, Y., Liang, W., Wang, J. & Chu, H. Distinct soil bacterial communities along a small-scale elevational gradient in alpine tundra. Front. Microbiol. 6, 538–541 (2014).
    CAS  Google Scholar 

    43.
    Zhang, C., Liu, G. B., Xue, S. & Xiao, L. Effect of different vegetation types on the rhizosphere soil microbial community structure in the loess plateau of China. J. Integr. Agric. 12, 2103–2113 (2013).
    Google Scholar 

    44.
    Weand, M. P., Arthur, M. A., Lovett, G. M., Mcculley, R. L. & Weathers, K. C. Effects of tree species and N additions on forest floor microbial communities and extracellular enzyme activities. Soil Biol. Biochem. 42, 2161–2173 (2010).
    CAS  Google Scholar 

    45.
    Wu, Z. et al. Terminal restriction fragment length polymorphism analysis of soil bacterial communities under different vegetation types in subtropical area. PLoS ONE https://doi.org/10.1371/journal.pone.0129397 (2015).
    Article  PubMed  PubMed Central  Google Scholar 

    46.
    Singh, D. et al. Strong elevational trends in soil bacterial community composition on Mt. Halla, South Korea. Soil Biol. Biochem. 68, 140–149 (2014).
    ADS  CAS  Google Scholar 

    47.
    Fallen, M. Linking water and nutrients through the vadose zone: A fungal interface between the soil and plant systems. J. Arid Land 206, 155–163 (2011).
    Google Scholar 

    48.
    Grayston, S. J., Griffith, G. S., Mawdsley, J. L., Campbell, C. D. & Bardgett, R. D. Accounting for variability in soil microbial communities of temperate upland grassland ecosystems. Soil Biol. Biochem. 33, 533–551 (2001).
    CAS  Google Scholar 

    49.
    Studies, R. Forest floor properties across sharp compositional boundaries separating trembling aspen and jack pine stands in the southern boreal forest. Plant Soil 345, 353–364 (2011).
    Google Scholar 

    50.
    Ruzicka, S., Edgerton, D., Norman, M. & Hill, T. The utility of ergosterol as a bioindicator of fungi in temperate soils. Soil Biol. Biochem. 32, 989–1005 (2000).
    CAS  Google Scholar 

    51.
    Weete, J. D., Abril, M. & Blackwell, M. Phylogenetic distribution of fungal sterols. PLoS ONE 5, e10899 (2010).
    ADS  PubMed  PubMed Central  Google Scholar 

    52.
    Shen, C. et al. Dramatic increases of soil microbial functional gene diversity at the treeline ecotone of Changbai mountain. Front. Microbiol. https://doi.org/10.3389/fmicb.2016.01184 (2016).
    Article  PubMed  PubMed Central  Google Scholar 

    53.
    Quideau, S. A., Chadwick, O. A., Benesi, A., Graham, R. C. & Anderson, M. A. A direct link between forest vegetation type and soil organic matter composition. Geoderma 104, 41–60 (2001).
    ADS  CAS  Google Scholar 

    54.
    Tian, J. et al. Linkages between the soil organic matter fractions and the microbial metabolic functional diversity within a broad-leaved Korean pine forest. Eur. J. Soil Biol. 66, 57–64 (2015).
    CAS  Google Scholar 

    55.
    Yuan, Y., Si, G., Jian, W., Luo, T. & Zhang, G. Bacterial community in alpine grasslands along an altitudinal gradient on the Tibetan plateau. FEMS Microbiol. Ecol. 87, 121–132 (2013).
    PubMed  Google Scholar 

    56.
    Fierer, N. et al. Microbes do not follow the elevational diversity patterns of plants and animals. Ecology 92, 797–804 (2011).
    PubMed  Google Scholar 

    57.
    Shen, C., Ni, Y., Liang, W., Wang, J. & Chu, H. Distinct soil bacterial communities along a small-scale elevational gradient in alpine tundra. Front. Microbiol. 6, 582. https://doi.org/10.3389/fmicb.2015.00582 (2015).
    Article  PubMed  PubMed Central  Google Scholar 

    58.
    Hinsinger, P., Bengough, A. G., Vetterlein, D. & Young, I. M. Rhizosphere: Biophysics, biogeochemistry and ecological relevance. Plant Soil 321, 117–152 (2009).
    CAS  Google Scholar 

    59.
    Shen, C. et al. Contrasting elevational diversity patterns between eukaryotic soil microbes and plants. Ecology 95, 3190–3202 (2014).
    Google Scholar 

    60.
    Jarvis, S. G., Woodward, S. & Taylor, A. F. S. Strong altitudinal partitioning in the distributions of ectomycorrhizal fungi along a short (300 m) elevation gradient. New Phytol. 206, 1145–1155 (2015).
    CAS  PubMed  Google Scholar 

    61.
    Lanzén, A. et al. Multi-targeted metagenetic analysis of the influence of climate and environmental parameters on soil microbial communities along an elevational gradient. Sci. Rep. 6, 28257 (2016).
    ADS  PubMed  PubMed Central  Google Scholar 

    62.
    Shen, C. et al. Soil pH drives the spatial distribution of bacterial communities along elevation on Changbai mountain. Soil Biol. Biochem. 57, 204–211 (2013).
    CAS  Google Scholar 

    63.
    Siles, J. A. & Margesin, R. Abundance and diversity of bacterial, archaeal, and fungal communities along an altitudinal gradient in alpine forest soils: What are the driving factors?. Microb. Ecol. 72, 207–220 (2016).
    PubMed  PubMed Central  Google Scholar 

    64.
    Sagovamareckova, M., Cermak, L., Omelka, M., Kyselkova, M. & Kopecky, J. Bacterial diversity and abundance of a creek valleysites reflected soil pH and season. Open Life Sci. https://doi.org/10.1515/biol-2015-0007 (2015).
    Article  Google Scholar 

    65.
    Smith, J. L., Halvorson, J. J. & Bolton, H. Soil properties and microbial activity across a 500m elevation gradient in a semi-arid environment. Soil Biol. Biochem. 34, 1749–1757 (2002).
    CAS  Google Scholar 

    66.
    Yergeau, E., Kang, S., He, Z., Zhou, J. & Kowalchuk, G. A. Functional microarray analysis of nitrogen and carbon cycling genes across an Antarctic latitudinal transect. ISME J. 1, 163–179 (2007).
    CAS  PubMed  Google Scholar 

    67.
    Kai, X. et al. Warming alters expressions of microbial functional genes important to ecosystem functioning. Front. Microbiol. 7, 668 (2016).
    ADS  Google Scholar 

    68.
    Jing, C. et al. Available nitrogen is the key factor influencing soil microbial functional gene diversity in tropical rainforest. BMC Microbiol. 15, 397–398 (2015).
    Google Scholar 

    69.
    Griffiths, R. I. et al. The bacterial biogeography of British soils. Environ. Microbiol. 13, 1642–1654 (2011).
    PubMed  Google Scholar 

    70.
    Davey, M. L., Nybakken, L., Kauserud, H. & Ohlson, M. Fungal biomass associated with the phyllosphere of bryophytes and vascular plants. Mycol. Res. 113, 1254–1260 (2009).
    CAS  PubMed  Google Scholar 

    71.
    Meier, C. L., Rapp, J., Bowers, R. M., Silman, M. & Fierer, N. Fungal growth on a common wood substrate across a tropical elevation gradient: Temperature sensitivity, community composition, and potential for above-ground decomposition. Soil Biol. Biochem. 42, 1083–1090 (2010).
    CAS  Google Scholar 

    72.
    Tu, S., Sun, J., Guo, Z. & Gu, F. On relationship between root exudates and plant nutrition in rhizosphere. Soil Environ. 9, 64–67 (2000).
    Google Scholar 

    73.
    Zhang, C., Liu, G., Xue, S. & Song, Z. Rhizosphere soil microbial activity under different vegetation types on the Loess Plateau, China. Geoderma 161, 115–125 (2011).
    ADS  CAS  Google Scholar 

    74.
    Zeng, S., Zhiyao, S. U., Chen, B. & Yuanchun, Y. U. A review on the rhizosphere nutrition ecology research. J. Nanjing For. Univ. 27, 79 (2003).
    Google Scholar 

    75.
    Wei, Z., Xiaojuan, Q. I., Jianwei, L., Zhengxiang, Y. U. & Xia, C. Characterization of microbial community structure in rhizosphere soils of cowskin Azalea (Rhododendron aureum Georgi) on northern slope of Changbai mountains, China. Chin. Geogr. Sci. 26, 78–89 (2016).
    Google Scholar 

    76.
    Yang, X. & Wu, G. The strategy for conservation and sustainable utilization of biodiversity in Changbaishan biosphere reserve. J. For. Res. 9, 217–222 (1998).
    Google Scholar 

    77.
    Zong, S. et al. Analysis of the process and impacts of Deyeuxia angustifolia invasion on the Alpine Tundra, Changbai mountain. Acta Ecol. Sin. 34, 87–104 (2014).
    Google Scholar 

    78.
    Batten, K. M., Scow, K. M., Davies, K. F. & Harrison, S. P. Two invasive plants alter soil microbial community composition in serpentine grasslands. Biol. Invas. 8, 217–230 (2006).
    Google Scholar 

    79.
    Mebius, L. J. A rapid method for the determination of organic carbon in soil. Anal. Chim. Acta 22, 120–124 (1960).
    CAS  Google Scholar 

    80.
    Industry, D. I. Design in industry. Electr. Power 28, 228–228 (1982).
    Google Scholar 

    81.
    Murphy, J. & Riley, J. P. A modified single solution method for the determination of phosphate in natural waters. Anal. Chim. Acta 27, 31–36 (1962).
    CAS  Google Scholar 

    82.
    Brookes, P. C., Landman, A., Pruden, G. & Jenkinson, D. S. Chloroform fumigation and the release of soil nitrogen: A rapid direct extraction method to measure microbial biomass nitrogen in soil. Soil Biol. Biochem. 17, 837–842 (1985).
    CAS  Google Scholar 

    83.
    Schinner, F. & Mersi, W. V. Xylanase-, CM-cellulase- and invertase activity in soil: An improved method. Soil Biol. Biochem. 22, 511–515 (1990).
    CAS  Google Scholar 

    84.
    Johnson, J. L. & Temple, K. L. Some variables affecting the measurement of “catalase activity” in Soil1. Soil Sci. Soc. Am. J. 28, 207–209 (1964).
    ADS  CAS  Google Scholar 

    85.
    Klose, S. & Tabatabai, M. A. Urease activity of microbial biomass in soils as affected by cropping systems. Biol. Fertil. Soils 31, 191–199 (2000).
    CAS  Google Scholar 

    86.
    Vaughan, D. & Ord, B. G. An effect of soil organic matter on invertase activity in soil. Soil Biol. Biochem. 12, 449–450 (1980).
    CAS  Google Scholar 

    87.
    Djajakirana, G., Joergensen, R. G. & Meyer, B. Ergosterol and microbial biomass relationship in soil. Biol. Fertil. Soils 22, 299–304 (1996).
    CAS  Google Scholar 

    88.
    Levy-Booth, D. J., Prescott, C. E. & Grayston, S. J. Microbial functional genes involved in nitrogen fixation, nitrification and denitrification in forest ecosystems. Soil Biol. Biochem. 75, 11–25 (2014).
    CAS  Google Scholar 

    89.
    Yuan, H. et al. Abundance and composition of CO_2 fixating bacteria in relation to long-term fertilization of paddy soils. Acta Ecol. Sin. 32, 183–189 (2012).
    CAS  Google Scholar 

    90.
    Chen, J., Yu, Z., Michel, F. C. Jr., Wittum, T. & Morrison, M. Development and application of real-time PCR assays for quantification of erm genes conferring resistance to macrolides-lincosamides-streptogramin B in livestock manure and manure management systems. Appl. Environ. Microbiol. 73, 4407–4416. https://doi.org/10.1128/AEM.02799-06 (2007).
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    91.
    Edgar, R. C. UPARSE: highly accurate OTU sequences from microbial amplicon reads. Nat. Methods 10, 996–998 (2013).
    CAS  PubMed  Google Scholar 

    92.
    Edgar, R. C. Search and clustering orders of magnitude faster than BLAST. Bioinformatics 26, 2460–2461 (2010).
    CAS  PubMed  Google Scholar 

    93.
    Maidak, B. L. et al. The ribosomal database project (RDP). Nucleic Acids Res. 24, 82–85 (1996).
    CAS  PubMed  PubMed Central  Google Scholar 

    94.
    CoreTeam, R. R: A language and environment for statistical computing. (2013).

    95.
    Oksanen, B. J., Kindt, R., Legendre, P. & O’Hara, B. vegan: Community Ecology Package. R package version 1.8-6 (accessed 10 December 2019); https://CRAN.R-project.org/package=vegan.

    96.
    Oksanen, J. et al. vegan: Community ecology package. R package version 1.17-3. J. Stat. Softw. 48, 103–132 (2010).
    Google Scholar 

    97.
    Nd, S. A., Potvin, L. R. & Lilleskov, E. A. Fertility-dependent effects of ectomycorrhizal fungal communities on white spruce seedling nutrition. Mycorrhiza 25, 649–662 (2015).
    Google Scholar 

    98.
    Arbuckle, J. L. Amos 7.0 User’s Guide. (SPSS, 2006). More

  • in

    Biological characteristics of Trissolcus urichi (Crawford) (Hymenoptera: Scelionidae) on Euschistus heros (Fabricius) and Dichelops melacanthus (Dallas) (Hemiptera: Pentatomidae) Eggs

    1.
    Panizzi, A. R. Economic importance of stink bugs (Pentatomidae). In Heteroptera of economic importance (eds Schaefer, C. W. & Panizzi, A. R.) 421–474 (CRC Press, Boca Ratón, 2000).
    Google Scholar 
    2.
    Akin, S., Phillips, J. & Johnson, D.T. Biology, identification and management of the redbanded stink bug. Arkansas, US Cooperative Extension Service, University of Arkansas, U.S. Dept. of Agriculture, and county governments cooperating. FSA7078 (2011).

    3.
    Corrêa-Ferreira, B. S. & Azevedo, J. Soybean seed damage by diferente species of stink bugs. Agric. For. Entomol. 4, 145–150 (2002).
    Google Scholar 

    4.
    Panizzi, A. R. & Slansky, F. Jr. Review of phytophagous pentatomids (Hemiptera: Pentatomidae) associated with soybean in the Americas. Fla. Entomol. 68, 184–203 (1985).
    Google Scholar 

    5.
    Zerbino, M. S. & Panizzi, A. R. The underestimated role of pest pentatomid parasitoids in Southern South America. Arth. Plant Int. 13, 703–718 (2019).
    Google Scholar 

    6.
    Panizzi, A. R. & Corrêa-Ferreira, B. S. Dynamics in the insect fauna adaptation to soybean in the tropics. Trends Entomol. 1, 71–88 (1997).
    Google Scholar 

    7.
    Gomes, E. C., Hayashida, R. & Bueno, A. F. Dichelops melacanthus and Euschistus heros injury on maize: Basis for re-evaluating stink bug thresholds for IPM decisions. Crop Prot. 130, 105050 (2020).
    CAS  Google Scholar 

    8.
    Smaniotto, L. F. & Panizzi, A. R. Interactions of selected species of stink bugs (Hemiptera: Heteroptera: Pentatomidae) from leguminous crops with plants in the Neotropics. Florida Entomol. 98, 7–17 (2015).
    Google Scholar 

    9.
    Bueno, A. F., Bortolotto, O. C., Pomari-Fernandes, A. & França-Neto, J. B. Assessment of a more conservative stink bug economic threshold for managing stink bugs in Brazilian soybean. Crop Prot. 71, 132–137 (2015).
    Google Scholar 

    10.
    Sosa-Gómez, D.R., Corso, I.C. & Morales, L. Insecticide resistance to endosulfan, monocrotophos and methamidophos in the neotropical brown stink bug, Euschistus heros (F.) Neotrop. Entomol. 30, 317–320 (2001).

    11.
    Sosa-Gómez, D. R. & Silva, J. J. D. Neotropical brown stink bug (Euschistus heros) resistance to methamidophos in Paraná Brazil. Pesq. Agrop. Bras 45, 767–769 (2010).
    Google Scholar 

    12.
    Bueno, A. F. et al. Effects of integrated pest management, biological control and prophylactic use of insecticides on the management and sustainability of soybean. Crop Prot. 30, 937–945 (2011).
    Google Scholar 

    13.
    van Lenteren, J. C., Bolckmans, K., Köhl, J., Ravensberg, W. J. & Urbaneja, A. Biological control using invertebrates and microorganisms: plenty of new opportunities. Biocontrol 63, 39–59 (2018).
    Google Scholar 

    14.
    Koppel, A. L., Herbert, D. A. Jr., Kuhar, T. P. & Kamminga, K. Survey of stink bug (Hemiptera: Pentatomidae) egg parasitoids in wheat, soybean, and vegetable crops in southeast Virginia. Environ. Entomol. 38, 375–379 (2009).
    CAS  PubMed  Google Scholar 

    15.
    Laumann, R. A. et al. Egg parasitoid wasps as natural enemies of the Neotropical stink bug Dichelops melacanthus. Pesq. Agropec. Bras. 45, 442–449 (2010).
    Google Scholar 

    16.
    Corrêa-Ferreira, B. S. & Moscardi, F. Seasonal occurrence and host spectrum of egg parasitoids associated with soybean stink bugs. Biol. Control. 5, 196–202 (1995).
    Google Scholar 

    17.
    Cividanes, F. J. Development and emergence of Trissolcus brochymenae (Ashmead) and Telenomus podisi Ashmead (Hymenoptera: Scelionidae) at different temperatures. An. Soc. Entomol. Bras. 25, 207–211 (1996).
    Google Scholar 

    18.
    Silva, G. V., Bueno, A. F., Neves, P. M. O. J. & Favetti, B. M. Biological characteristics and parasitism capacity of Telenomus podisi (Hymenoptera: Platygastridae) on eggs of Euschistus heros (Hemiptera: Pentatomidae). J. Agric. Sci. 10, 210–220 (2018).
    Google Scholar 

    19.
    Laumann, R. A. et al. Comparative biology and functional response of Trissolcus spp. (Hymenoptera: Scelionidae) and implications for stink bugs (Hemiptera: Pentatomidae) biological control. Biol. Control. 44, 32–41 (2008).
    Google Scholar 

    20.
    Favetti, B. M., Krinski, D., Butnariu, A. R. & Loiácono, M. S. Egg parasitoids of Edessa meditabunda (Fabricius) (Pentatomidae) in lettuce crop. Rev. Bras. Entomol. 57, 236–237 (2013).
    Google Scholar 

    21.
    Margaría, C. B., Loiácono, M. S. & Lanteri, A. A. New geographic and host records for scelionid wasps (Hymenoptera: Scelionidae) parasitoids of insect pests in South America. Zootaxa 2314, 41–49 (2009).
    Google Scholar 

    22.
    Peres, W. A. A. & Corrêa-Ferreira, B. S. Methodology of mass multiplication of Telenomus podisi Ashmead and Trissolcus basalis (Hymenoptera: Scelionidae) on eggs of Euschistus heros (Hemiptera: Pentatomidae). Neotrop. Entomol. 33, 457–462 (2004).
    Google Scholar 

    23.
    Panizzi, A. R., Parra, J. R. P., Santos, C. H. & Carvalho, D. R. Rearing the southern green stink bug using artificial dry diet and artificial plant. Pesq. Agropec. Bras. 35, 1709–1715 (2000).
    Google Scholar 

    24.
    Thuler, R. T., Volpe, H. X. L., Bortoli, S. A., Goulart, R. M. & Viana, C. L. T. Metodologia para avaliação da preferência hospedeira de parasitoides do gênero Trichogramma Westood. Bol. San. Veg. 33, 333–340 (2007).
    Google Scholar 

    25.
    Queiroz, A. P., Taguti, E. A., Bueno, A. F., Grande, M. L. M. & Costa, C. O. Host preferences of Telenomus podisi (Hymenoptera: Scelionidae): parasitism on eggs of Dichelops melacanthus, Euschistus heros, and Podisus nigrispinus (Hemiptera: Pentatomidae). Neotrop. Entomol. 47, 543–552 (2018).
    CAS  PubMed  Google Scholar 

    26.
    van Lenteren, J. C. Quality control and production of biological control agents: theory and testing procedures 327 (CABI, Wallingford, 2003).
    Google Scholar 

    27.
    Shapiro, S. S. & Wilk, M. B. An analysis of variance test for normality (complete samples). Biometrika 52, 591–611 (1965).
    MathSciNet  MATH  Google Scholar 

    28.
    Burr, I. W. & Foster, L. A. A Test for Equality of Variances (University of Purdue, West Lafayette, 1972).
    Google Scholar 

    29.
    Institute, S. A. S. SAS User’s Guide: Statistics, Version 8e (SAS Institute, Cary, NC, 2009).
    Google Scholar 

    30.
    Sujii, E. R., Costa, M. L. M., Pires, C. S. S., Colazza, S. & Borges, M. Inter and intra-guild interactions in egg parasitoid species of the soybean stink bug complex. Pesq. Agropec. Bras. 37, 1541–1549 (2002).
    Google Scholar 

    31.
    Zhou, Y., Abram, P. K., Boivin, G. & Brodeur, J. Increasing host age does not have the expected negative effects on the fitness parameters of an egg parasitoid. Entomol. Exp. Appl. 151, 106–111 (2014).
    Google Scholar 

    32.
    Jones, T. S., Bilton, A. R., Mak, L. & Sait, S. M. Host switching in a generalist parasitoid: contrasting transient and transgenerational costs associated with novel and original host species. Ecol. Evol. 5, 459–465 (2015).
    PubMed  PubMed Central  Google Scholar 

    33.
    Orr, D. B. Scelionid wasps as biological control agents: a review. Florida Entomol. 71, 506–528 (1988).
    Google Scholar 

    34.
    Blackiston, D. J., Casey, E. S. & Weiss, M. R. Retention of memory through metamorphosis: can a moth remember what it learned as a caterpillar?. PlosOne 3, e1736 (2008).
    ADS  Google Scholar 

    35.
    Kaiser, L., Pham-Delegue, M. H. & Masson, C. Behavioural study of plasticity in host preferences of Trichogramma maidis (Hymenoptera: Trichogrammatidae). Physiol. Entomol. 14, 53–60 (1989).
    Google Scholar 

    36.
    Gandolfi, M., Mattiacci, L. & Dorn, S. Preimaginal learning determines adult response to chemical stimuli in a parasitic wasp. Proc. R. Soc. Lond. B 270, 2623–2629 (2003).
    Google Scholar 

    37.
    Corbet, S. A. Insect chemosensory responses: a chemical legacy hypothesis. Ecol. Entomol. 10, 143–153 (1985).
    Google Scholar 

    38.
    Pluke, R. W. H. & Leibee, G. L. Host preferences of Trichogramma pretiosum and the influence of prior ovipositional experience on the parasitism of Plutella xylostella and Pseudoplusia includes eggs. Biocontrol 51, 569–583 (2006).
    Google Scholar 

    39.
    Stephens, D. W. & Krebs, J. R. Foraging theory (Princeton University Press, Princeton, 1986).
    Google Scholar 

    40.
    Vinson, S. B. & Iwantsch, G. F. Host suitability for insect parasitoids. Annu. Rev. Entomol. 25, 397–419 (1980).
    Google Scholar 

    41.
    Bin, F., Vinson, S. B., Strand, M. R., Colazza, S. & Jones, W. A. Jr. Source of an egg kairomone for Trissolcus basalis, a parasitoid of Nezara viridula. Physiol. Entomol. 18, 7–15 (1993).
    Google Scholar 

    42.
    Borges, M. et al. Semiochemical and physical stimuli involved in host recognition by Telenomus podisi (Hymenoptera: Scelionidae) toward Euschistus heros (Heteroptera: Pentatomidae). Physiol. Entomol. 24, 227–233 (1999).
    Google Scholar 

    43.
    Borges, M. & Aldrich, J. R. Attractant pheromone for Nearctic stink bug, Euschistus obscurus (Heteroptera: Pentatomidae): insight in to a Neotropical relative. J. Chem. Ecol. 20, 1095–1102 (1994).
    CAS  PubMed  Google Scholar 

    44.
    Pomari, A. F., Bueno, A. F., Bueno, R. C. O. F. & Menezes Junior, A. O. Biological Characteristics and thermal requirements of the biological control agent Telenomus remus (Hymenoptera: Platygastridae) reared on eggs of different species of the genus Spodoptera (Lepidoptera: Noctuidae). Ann. Entomol. Soc. Am. 105, 73–81 (2012).
    Google Scholar 

    45.
    Bueno, R. C. O., Parra, J. R. P. & Bueno, A. F. Biological characteristics and thermal requirements of a Brazilian strain of the parasitoid Trichogramma pretiosum reared on eggs of Pseudoplusia includes and Anticarsia gemmatalis. Biol. Control. 51, 355–361 (2009).
    Google Scholar 

    46.
    Cônsoli, F. L., Kitajima, E. W. & Parra, J. R. P. Ultrastructure of the natural and factitious host eggs of Trichogramma galloi Zucchi and Trichogramma pretiosum Riley (Hymenoptera: Trichogrammatidae). Int. J. Insect. Morphol. Embriol. 28, 211–229 (1999).
    Google Scholar 

    47.
    Bai, B., Luck, R. F., Forster, L., Stephens, B. & Janssen, J. A. M. The effect of host size on quality attributes of the egg parasitoid Trichogramma pretiosum. Entomol. Exp. Appl. 64, 37–48 (1992).
    Google Scholar 

    48.
    Schwartz, A. & Gerling, D. Adult biology of Telenomus remus (Hymenoptera: Scelionidae) under laboratory conditions. Entomophaga 19, 482–492 (1974).
    Google Scholar 

    49.
    Charnov, E. L., Los-Den Hartogh, R. L., Jones, W. T. & Van Den Assem, J. Sex ratio evolution in a variable environment. Nature 289, 27–33 (1981).
    ADS  CAS  PubMed  Google Scholar 

    50.
    Houseweart, M. W., Jennings, D. T., Welty, C. & Southard, S. G. Progeny production by Trichogramma minutum (Hymenoptera: Trichogrammatidae) utilizing eggs for Choristoneura fumiferana (Lepidoptera: Tortricidae) and Sitotroga cerealella (Lepidoptera: Gelechiidae). Can. Entomol. 115, 1245–1252 (1983).
    Google Scholar 

    51.
    Sequeira, R. & Mackauer, M. Covariance of adult size and development time in the parasitoid wasp Aphidius ervi in relation to the size of its host Acyrthosiphon pisum. Evol. Ecol. 6, 34–44 (1992).
    Google Scholar 

    52.
    Mackauer, M. Sexual size dimorphism in solitary wasps: influence of host quality. Oikos 76, 265–272 (1996).
    Google Scholar  More