More stories

  • in

    Rare species disproportionally contribute to functional diversity in managed forests

    Hooper, D. U. et al. Effects of biodiversity on ecosystem functioning: a consensus of current knowledge. Ecol. Monogr. 75, 3–35 (2005).Article 

    Google Scholar 
    Schleuter, D., Daufresne, M., Massol, F. & Argillier, C. A user’s guide to functional diversity indices. Ecol. Monogr. 80, 469–484 (2010).Article 

    Google Scholar 
    Petchey, O. L. & Gaston, K. J. Extinction and the loss of functional diversity. Proc. R. Soc. B Biol. Sci. 269, 1721–1727 (2002).Article 

    Google Scholar 
    Tilman, D. et al. The influence of functional diversity and composition on ecosystem processes. Science (80-. ). 277, 1300–1302 (1997).Díaz, S. & Cabido, M. Vive la différence: Plant functional diversity matters to ecosystem processes. Trends Ecol. Evol. 16, 646–655 (2001).Article 

    Google Scholar 
    Tilman, D. Functional diversity. in Encyclopedia of Biodiversity, Volume 3 (ed. Levin, S. A.) 109–120 (Academic Press, 2001).McGill, B. J., Enquist, B. J., Weiher, E. & Westoby, M. Rebuilding community ecology from functional traits. Trends Ecol. Evol. 21, 178–185 (2006).PubMed 
    Article 

    Google Scholar 
    Cadotte, M. W., Carscadden, K. & Mirotchnick, N. Beyond species: Functional diversity and the maintenance of ecological processes and services. J. Appl. Ecol. 48, 1079–1087 (2011).Article 

    Google Scholar 
    Petchey, O. L., Hector, A. & Gaston, K. J. How do different measures of functional diversity perform?. Ecology 85, 847–857 (2004).Article 

    Google Scholar 
    Cardinale, B. J. et al. Biodiversity loss and its impact on humanity. Nature 486, 59–67 (2012).CAS 
    Article 
    ADS 

    Google Scholar 
    Petchey, O. L. & Gaston, K. J. Functional diversity (FD), species richness and community composition. Ecol. Lett. 5, 402–411 (2002).Article 

    Google Scholar 
    Halpern, B. S. & Floeter, S. R. Functional diversity responses to changing species richness in reef fish communities. Mar. Ecol. Prog. Ser. 364, 147–156 (2008).Article 
    ADS 

    Google Scholar 
    Seymour, C. L., Simmons, R. E., Joseph, G. S. & Slingsby, J. A. On bird functional diversity: Species richness and functional differentiation show contrasting responses to rainfall and vegetation structure in an arid landscape. Ecosystems 18, 971–984 (2015).Article 

    Google Scholar 
    Müller, J., Jarzabek-Müller, A., Bussler, H. & Gossner, M. M. Hollow beech trees identified as keystone structures for saproxylic beetles by analyses of functional and phylogenetic diversity. Anim. Conserv. 17, 154–162 (2014).Article 

    Google Scholar 
    Ulrich, W. et al. Species assortment or habitat filtering: A case study of spider communities on lake islands. Ecol. Res. 25, 375–381 (2010).Article 

    Google Scholar 
    Mouillot, D., Dumay, O. & Tomasini, J. A. Limiting similarity, niche filtering and functional diversity in coastal lagoon fish communities. Estuar. Coast. Shelf Sci. 71, 443–456 (2007).Article 
    ADS 

    Google Scholar 
    Cadotte, M. W. & Tucker, C. M. Should environmental filtering be abandoned?. Trends Ecol. Evol. 32, 429–437 (2017).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    Flynn, D. F. B. et al. Loss of functional diversity under land use intensification across multiple taxa. Ecol. Lett. 12, 22–33 (2009).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    Rader, R., Bartomeus, I., Tylianakis, J. M. & Laliberté, E. The winners and losers of land use intensification: Pollinator community disassembly is non-random and alters functional diversity. Divers. Distrib. 20, 908–917 (2014).Article 

    Google Scholar 
    Sol, D. et al. The worldwide impact of urbanisation on avian functional diversity. Ecol. Lett. 23, 962–972 (2020).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    Bihn, J. H., Gebauer, G. & Brandl, R. Loss of functional diversity of ant assemblages in secondary tropical forests. Ecology 91, 782–792 (2010).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    Balestrieri, R. et al. A guild-based approach to assessing the influence of beech forest structure on bird communities. For. Ecol. Manage. 356, 216–223 (2015).Article 

    Google Scholar 
    Basile, M., Mikusiński, G. & Storch, I. Bird guilds show different responses to tree retention levels: A meta-analysis. Glob. Ecol. Conserv. 18, e00615 (2019).Article 

    Google Scholar 
    Czeszczewik, D. et al. Effects of forest management on bird assemblages in the Bialowieza Forest, Poland. iForest – Biogeosciences For. 8, 377–385 (2015).Article 

    Google Scholar 
    Wesołowski, T. Primeval conditions—What can we learn from them? Ibis (Lond. 1859). 149, 64–77 (2007).Paillet, Y. et al. Biodiversity differences between managed and unmanaged forests: meta-analysis of species richness in europe. Conserv. Biol. 24, 101–112 (2010).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    Götzenberger, L. et al. Ecological assembly rules in plant communities-approaches, patterns and prospects. Biol. Rev. 87, 111–127 (2012).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    Fox, J. W. & Kerr, B. Analyzing the effects of species gain and loss on ecosystem function using the extended Price equation partition. Oikos 121, 290–298 (2012).Article 

    Google Scholar 
    Fox, J. W. Using the Price Equations to partition the effects of biodiversity loss on ecosystem function. Ecology 87, 2687–2696 (2006).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    Winfree, R. W., Fox, J., Williams, N. M., Reilly, J. R. & Cariveau, D. P. Abundance of common species, not species richness, drives delivery of a real-world ecosystem service. Ecol. Lett. 18, 626–635 (2015).PubMed 
    Article 

    Google Scholar 
    Storch, I. et al. Evaluating the effectiveness of retention forestry to enhance biodiversity in production forests of Central Europe using an interdisciplinary, multi‐scale approach. Ecol. Evol. ece3.6003 (2020) https://doi.org/10.1002/ece3.6003.Pommerening, A. & Murphy, S. T. A review of the history, definitions and methods of continuous cover forestry with special attention to afforestation and restocking. Forestry 77, 27–44 (2004).Article 

    Google Scholar 
    Bauhus, J., Puettmannn, K. J. & Kühne, C. Close-to-nature forest management in Europe: does it support complexity and adaptability of forest ecosystems? in Managing Forests as Complex Adaptive Systems: Building Resilience to the Challenge of Global Change 187–213 (Routledge/The Earthscan Forest Library, 2013). https://doi.org/10.4324/9780203122808.Bauhus, J., Puettmannn, K. J. & Kühne, C. Is Close-to-Nature Forest Management in Europe Compatible with Managing Forests as Complex Adaptive Forest Ecosystems? in Managing Forests as Complex Adaptive Systems: Building Resilience to the Challenge of Global Change (eds. Messier, C., Puettmannn, K. J. & Coates, K. D.) 187–213 (Routledge/The Earthscan Forest Library, 2013).Balestrieri, R., Basile, M., Posillico, M., Altea, T. & Matteucci, G. Survey effort requirements for bird community assessment in forest habitats. Acta Ornithol. 52, 1–9 (2017).Article 

    Google Scholar 
    Wilman, H. et al. EltonTraits 1.0: Species-level foraging attributes of the world’s birds and mammals. Ecology 95, 2027 (2014).Article 

    Google Scholar 
    Laliberte, E. & Legendre, P. A distance-based framework for measuring functional diversity from multiple traits. Ecology 91, 299–305 (2010).PubMed 
    Article 

    Google Scholar 
    Gower, J. C. A general coefficient of similarity and some of its properties. Biometrics 27, 857 (1971).Article 

    Google Scholar 
    Kahl, T. & Bauhus, J. An index of forest management intensity based on assessment of harvested tree volume, tree species composition and dead wood origin. Nat. Conserv. 7, 15–27 (2014).Article 

    Google Scholar 
    Paillet, Y. et al. Quantifying the recovery of old-growth attributes in forest reserves: A first reference for France. For. Ecol. Manage. 346, 51–64 (2015).Article 

    Google Scholar 
    Burrascano, S., Lombardi, F. & Marchetti, M. Old-growth forest structure and deadwood: Are they indicators of plant species composition? A case study from central Italy. Plant Biosyst. 142, 313–323 (2008).Article 

    Google Scholar 
    Van Wagner, C. E. Practical aspects of the line intersect method. (Minister of Supply and Services Canada, 1982).Larrieu, L. et al. Tree related microhabitats in temperate and Mediterranean European forests: A hierarchical typology for inventory standardization. Ecol. Indic. 84, 194–207 (2018).Article 

    Google Scholar 
    Asbeck, T., Pyttel, P., Frey, J. & Bauhus, J. Predicting abundance and diversity of tree-related microhabitats in Central European montane forests from common forest attributes. For. Ecol. Manage. 432, 400–408 (2019).Article 

    Google Scholar 
    Paillet, Y. et al. The indicator side of tree microhabitats: A multi-taxon approach based on bats, birds and saproxylic beetles. J. Appl. Ecol. https://doi.org/10.1111/1365-2664.13181 (2018).Article 

    Google Scholar 
    Basile, M. et al. What do tree-related microhabitats tell us about the abundance of forest-dwelling bats, birds, and insects?. J. Environ. Manage. 264, 110401 (2020).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    Wang, Q., Adiku, S., Tenhunen, J. & Granier, A. On the relationship of NDVI with leaf area index in a deciduous forest site. Remote Sens. Environ. 94, 244–255 (2005).Article 
    ADS 

    Google Scholar 
    Rafique, R., Zhao, F., De Jong, R., Zeng, N. & Asrar, G. R. Global and regional variability and change in terrestrial ecosystems net primary production and NDVI: A model-data comparison. Remote Sens. 8, 1–16 (2016).Article 

    Google Scholar 
    Bates, D. et al. Package ‘lme4’. R Found. Stat. Comput. Vienna 12, (2014).Zuur, A. F., Ieno, E. N., Walker, N., Saveliev, A. A. & Smith, G. M. Mixed effects models and extensions in ecology with R. (Springer, New York, 2009). https://doi.org/10.1007/978-0-387-87458-6.Hartig, F. DHARMa: Residual Diagnostics for Hierarchical (Multi-Level / Mixed) Regression Models. (2019).R Core Team. R: A language and environment for statistical computing. (2021).Mayfield, M. M. et al. What does species richness tell us about functional trait diversity? Predictions and evidence for responses of species and functional trait diversity to land-use change. Glob. Ecol. Biogeogr. 19, 423–431 (2010).
    Google Scholar 
    Pavoine, S. & Bonsall, M. B. Measuring biodiversity to explain community assembly: a unified approach. Biol. Rev. 86, 792–812 (2011).CAS 
    PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    Mayfield, M. M., Boni, M. F., Daily, G. C. & Ackerly, D. Species and functional diversity of natie and human-dominated plant communities. Ecology 86, 2365–2372 (2005).Article 

    Google Scholar 
    Holdaway, R. J. & Sparrow, A. D. Assembly rules operating along a primary riverbed-grassland successional sequence. J. Ecol. 94, 1092–1102 (2006).Article 

    Google Scholar 
    Matuoka, M. A., Benchimol, M., de Almeida-Rocha, J. M. & Morante-Filho, J. C. Effects of anthropogenic disturbances on bird functional diversity: A global meta-analysis. Ecol. Indic. 116, 106471 (2020).Article 

    Google Scholar 
    Leaver, J., Mulvaney, J., Ehlers-Smith, D. A., Ehlers-Smith, Y. C. & Cherry, M. I. Response of bird functional diversity to forest product harvesting in the Eastern Cape, South Africa. For. Ecol. Manage. 445, 82–95 (2019).Article 

    Google Scholar 
    Poos, M. S., Walker, S. C. & Jackson, D. A. Functional-diversity indices can be driven by methodological choices and species richness. Ecology 90, 341–347 (2009).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    Mayfield, M. M., Boni, M. F., Daily, G. C. & Ackerly, D. Species and functional diversity of native and human-dominated plant communities. Ecology 86, 2365–2372 (2005).Article 

    Google Scholar 
    Tsianou, M. A. & Kallimanis, A. S. Different species traits produce diverse spatial functional diversity patterns of amphibians. Biodivers. Conserv. 25, 117–132 (2016).Article 

    Google Scholar 
    Gregory, R. D., Skorpilova, J., Vorisek, P. & Butler, S. An analysis of trends, uncertainty and species selection shows contrasting trends of widespread forest and farmland birds in Europe. Ecol. Indic. 103, 676–687 (2019).Article 

    Google Scholar 
    Peña, R. et al. Biodiversity components mediate the response to forest loss and the effect on ecological processes of plant–frugivore assemblages. Funct. Ecol. 34, 1257–1267 (2020).Article 

    Google Scholar 
    Chase, J. M., Blowes, S. A., Knight, T. M., Gerstner, K. & May, F. Ecosystem decay exacerbates biodiversity loss with habitat loss. Nature 584, 238–243 (2020).CAS 
    PubMed 
    Article 
    ADS 
    PubMed Central 

    Google Scholar 
    Fedrowitz, K. et al. Can retention forestry help conserve biodiversity? A meta-analysis. J. Appl. Ecol. 51, 1669–1679 (2014).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Horák, J. et al. Green desert?: Biodiversity patterns in forest plantations. For. Ecol. Manage. 433, 343–348 (2019).Article 

    Google Scholar 
    Ameztegui, A. et al. Bird community response in mountain pine forests of the Pyrenees managed under a shelterwood system. For. Ecol. Manage. 407, 95–105 (2017).Article 

    Google Scholar 
    Basile, M., Balestrieri, R., de Groot, M., Flajšman, K. & Posillico, M. Conservation of birds as a function of forestry. Ital. J. Agron. 11, 42–48 (2016).
    Google Scholar 
    Uezu, A. & Metzger, J. P. Vanishing bird species in the Atlantic Forest: Relative importance of landscape configuration, forest structure and species characteristics. Biodivers. Conserv. 20, 3627–3643 (2011).Article 

    Google Scholar 
    Endenburg, S. et al. The homogenizing influence of agriculture on forest bird communities at landscape scales. Landsc. Ecol. 34, 1–15 (2019).Article 

    Google Scholar 
    Reif, J. et al. Changes in bird community composition in the Czech Republic from 1982 to 2004: Increasing biotic homogenization, impacts of warming climate, but no trend in species richness. J. Ornithol. 154, 359–370 (2013).Article 

    Google Scholar 
    Morelli, F. et al. Evidence of evolutionary homogenization of bird communities in urban environments across Europe. Glob. Ecol. Biogeogr. 25, 1284–1293 (2016).Article 

    Google Scholar 
    Devictor, V., Julliard, R., Couvet, D., Lee, A. & Jiguet, F. Functional homogenization effect of urbanization on bird communities. Conserv. Biol. 21, 741–751 (2007).PubMed 
    Article 

    Google Scholar 
    Doxa, A., Paracchini, M. L., Pointereau, P., Devictor, V. & Jiguet, F. Preventing biotic homogenization of farmland bird communities: The role of High Nature Value farmland. Agric. Ecosyst. Environ. 148, 83–88 (2012).Article 

    Google Scholar 
    Van Turnhout, C. A. M., Foppen, R. P. B., Leuven, R. S. E. W., Siepel, H. & Esselink, H. Scale-dependent homogenization: Changes in breeding bird diversity in the Netherlands over a 25-year period. Biol. Conserv. 134, 505–516 (2007).Article 

    Google Scholar 
    Clavero, M. & Brotons, L. Functional homogenization of bird communities along habitat gradients: Accounting for niche multidimensionality. Glob. Ecol. Biogeogr. 19, 684–696 (2010).
    Google Scholar 
    Gustafsson, L. et al. Retention as an integrated biodiversity conservation approach for continuous-cover forestry in Europe. Ambio 49, 85–97 (2020).PubMed 
    Article 

    Google Scholar 
    Lelli, C. et al. Biodiversity response to forest structure and management: Comparing species richness, conservation relevant species and functional diversity as metrics in forest conservation. For. Ecol. Manage. 432, 707–717 (2019).Article 

    Google Scholar 
    Aquilué, N., Messier, C., Martins, K. T., Dumais-Lalonde, V. & Mina, M. A simple-to-use management approach to boost adaptive capacity of forests to global uncertainty. For. Ecol. Manage. 481, (2021).Manes, F., Ricotta, C., Salvatori, E., Bajocco, S. & Blasi, C. A multiscale analysis of canopy structure in Fagus sylvatica L. and Quercus cerris L. old-growth forests in the Cilento and Vallo di Diano National Park. Plant Biosyst. 144, 202–210 (2010).Article 

    Google Scholar 
    Tscharntke, T. et al. Landscape moderation of biodiversity patterns and processes – eight hypotheses. Biol. Rev. 87, 661–685 (2012).PubMed 
    Article 
    PubMed Central 

    Google Scholar 
    Kirsch, J.-J. et al. The use of water-filled tree holes by vertebrates in temperate forests. Wildlife Biol. 2021, wlb.00786
    (2021). More

  • in

    MiDAS 4: A global catalogue of full-length 16S rRNA gene sequences and taxonomy for studies of bacterial communities in wastewater treatment plants

    The MiDAS global consortium was established in 2018 to coordinate the sampling and collection of metadata from WWTPs across the globe (Supplementary Data 1). Samples were obtained in duplicates from 740 WWTPs in 425 cities, 31 countries on six continents (Fig. 1a). The majority of the WWTPs were configured with the activated sludge process (69.7%) (Fig. 1b), and these were the main focus of the subsequent analyses. Nevertheless, WWTPs based on biofilters, moving bed bioreactors (MBBR), membrane bioreactors (MBR), and granular sludge were also sampled to cover the microbial diversity in other types of WWTPs. The activated sludge plants were designed for carbon removal only (C; 22.1%), carbon removal with nitrification (C,N; 9.5%), carbon removal with nitrification and denitrification (C,N,DN; 40.9%), and carbon removal with nitrogen removal and enhanced biological phosphorus removal, EBPR (C,N,DN,P; 21.7%) (Fig. 1c). The first type represents the simplest design whereas the latter represents the most advanced process type with varying oxic and anoxic stages or compartments.Fig. 1: Sampling of WWTPs across the world.a Geographical distribution of WWTPs included in the study and their process configuration. b Distribution of plant types. MBBR moving bed bioreactor, MBR membrane bioreactor. c Distribution of process types for the activated sludge plants. C carbon removal, C,N carbon removal with nitrification, C,N,DN carbon removal with nitrification and denitrification, C,N,DN,P carbon removal with nitrogen removal and enhanced biological phosphorus removal (EBPR). The values next to the bars are the number of WWTPs in each group.Full size imageMiDAS 4: a global 16S rRNA gene catalogue and taxonomy for WWTPsMicrobial community profiling at high taxonomic resolution (genus- and species-level) using 16S rRNA gene amplicon sequencing requires a reference database with high-identity reference sequences (≥99% sequence identity) for the majority of the bacteria in the samples and a complete seven-rank taxonomy (domain to species) for all reference sequences16,20. To create such a database for bacteria in WWTPs globally, we applied synthetic long-read full-length 16S rRNA gene sequencing20,21 on samples from all WWTPs included in this study.More than 5.2 million full-length 16S rRNA gene sequences were obtained after quality filtering and primer trimming. The sequences were processed with AutoTax20 to yield 80,557 full-length 16S rRNA gene amplicon sequence variant (FL-ASVs). These reference sequences were added to our previous MiDAS 3 database16, providing a combined database (MiDAS 4) with a total of 90,164 unique, chimera-free FL-ASV reference sequences. The absence of detectable chimeric sequences is a unique feature of the database and is achieved due to the attachment of unique molecular identifiers (UMIs) to each end of the original template molecules before any PCR amplification steps21. This allows filtering of true biological sequences from chimera already in the synthetic long-read assembly20,21. The novelty of the FL-ASVs were determined based on the percent identity shared with their closest relatives in the SILVA 138 SSURef NR99 database and the threshold for each taxonomic rank proposed by Yarza et al.22. Out of all FL-ASVs, 88% had relatives above the genus-level threshold (≥94.5% identity) and 56% above the species-level threshold (≥98.7% identity) (Fig. 2 and Table 1).Fig. 2: Novel sequences and de novo taxa defined in the MiDAS 4 reference database.The phylogenetic trees are based on a multiple alignment of all MiDAS 4 reference sequences, which were first aligned against the global SILVA 138 alignment using the SINA aligner, and subsequently pruned according to the ssuref:bacteria positional variability by parsimony filter in ARB to remove hypervariable regions. The eight phyla with most FL-ASVs are highlighted in different colours. Sequence novelty was determined by the percent identity between each FL-ASV and their closest relative in the SILVA_138_SSURef_Nr99 database according to Usearch mapping and the taxonomic thresholds proposed by Yarza et al.22 shown in Table 1. Taxonomy novelty was defined based on the assignment of de novo taxa by AutoTax20.Full size imageTable 1 Novel sequences and de novo taxa observed in the MiDAS 4 reference database.Full size tableMiDAS 4 provides placeholder names for many environmental taxaAlthough only a small percentage of the reference sequences in MiDAS 4 represented new putative taxa at higher ranks (phylum, class, or order) according to the sequence identity thresholds proposed by Yarza et al.22, a large number of sequences lacked lower-rank taxonomic classifications and was assigned de novo placeholder names by AutoTax20 (Fig. 2 and Table 1). In total, de novo taxonomic names were generated by AutoTax for 26 phyla (30.6% of observed), 83 classes (37.2% of observed), 297 orders (46.8% of observed), and more than 8000 genera (86.3% of observed). Without the de novo taxonomy we would not be able to discuss these taxa across studies to unveil their potential role in wastewater treatment systems.Phylum-specific phylogenetic trees were created to determine if the FL-ASV reference sequences that were assigned to de novo phyla were actual phyla or simply artifacts related to the naive sequence identity-based assignment of de novo placeholder taxonomies (Supplementary Fig. 1a). The majority (65 FL-ASVs) created deep branches from within the Alphaproteobacteria together with 16S rRNA gene sequences from mitochondria, suggesting they represented divergent mitochondrial genes rather than true novel phyla. We also observed several FL-ASVs assigned to de novo phyla that branched from the classes Parcubacteria (3 FL-ASVs) and Microgenomatis (22 FL-ASVs) within the Patescibacteria phylum. These two classes were originally proposed as superphyla due to an unusually high rate of evolution of their 16S rRNA genes23,24. It is, therefore, likely that these de novo phyla are also artefacts due to the simple taxonomy assignment approach, which does not take different evolutionary rates into account20. Most of the class- and order-level novelty was found within the Patescibacteria, Proteobacteria, Firmicutes, Planctomycetota and Verrucomicrobiota. (Supplementary Fig. 1b). At the family- and genus-level, we also observed many de novo taxa affiliated to Bacteroidota, Bdellovibrionota and Chloroflexi.MiDAS 4 provides a common taxonomy for the fieldThe performance of the MiDAS 4 database was evaluated based on an independent amplicon dataset from the Global Water Microbiome Consortium (GWMC) project2, which covers ~1200 samples from 269 WWTPs. The raw GWMC amplicon data of the 16S rRNA gene V4 region was resolved into ASVs, and the percent identity to their best hits in MiDAS 4 and other reference databases was calculated (Fig. 3). The MiDAS 4 database had high-identity hits (≥99% identity) for 72.0 ± 9.5% (mean ± SD) of GWMC ASVs with ≥0.01% relative abundance, compared to 57.9 ± 8.5% for the SILVA 138 SSURef NR99 database, which was the best of the universal reference databases (Fig. 3). The relative abundance cutoff selects taxa that likely have a quantitative impact on the ecosystem while filtering out the rare biosphere which includes many bacteria introduced with the influent wastewaters25. Similar analyses of ASVs obtained from the samples included in this study showed, not surprisingly, even better performance with high-identity hits for 90.7 ± 7.9% of V1–V3 ASVs and 90.0 ± 6.6% of V4 ASVs with ≥0.01% relative abundance, compared to 60.6 ± 11.9% and 73.9 ± 10.3% for SILVA (Supplementary Fig. 2a). Although the sampling of WWTPs was focused towards activated sludge plants, the MiDAS 4 database also includes high-identity references for most ASVs in other plant types (granules, biofilters, etc.) (Supplementary Fig. 2b). This suggests that most taxa were shared across plant types, although often present in other relative abundances.Fig. 3: Database evaluation based on amplicon data from the Global Water Microbiome Consortium project.Raw amplicon data from the Global Water Microbiome Consortium project2 was processed to resolve ASVs of the 16S rRNA gene V4 region. The ASVs for each of the samples were filtered based on their relative abundance (only ASVs with ≥0.01% relative abundance were kept) before the analyses. The percentage of the microbial community represented by the remaining ASVs after the filtering was 88.35 ± 2.98% (mean ± SD) across samples. High-identity (≥99%) hits were determined by the stringent mapping of ASVs to each reference database. Classification of ASVs was done using the SINTAX classifier. The violin and box plots represent the distribution of percent of ASVs with high-identity hits or genus/species-level classifications for each database across n = 1165 biologically independent samples. Box plots indicate median (middle line), 25th, 75th percentile (box) and the min and max values after removing outliers based on 1.5x interquartile range (whiskers). Outliers have been removed from the box plots to ease visualisation. Different colours are used to distinguish the different databases.Full size imageUsing MiDAS 4 with the SINTAX classifier, it was possible to obtain genus-level classifications for 75.0 ± 6.9% of the GWMC ASVs with ≥0.01% relative abundance (Fig. 3). In comparison, SILVA 138 SSURef NR99, which was the best of the universal reference databases, could only classify 31.4 ± 4.2% of the ASVs to genus-level. When MiDAS 4 was used to classify amplicons from this study, we obtained genus-level classification for 92.0 ± 4.0% of V1–V3 ASVs and 84.8 ± 3.6% of V4 ASVs (Supplementary Fig. 2a). This is close to the theoretical limit set by the phylogenetic signal provided by each amplicon region analyzed20. Improved classifications were also observed for archaeal V4 ASVs (93.3 ± 10.6% for MiDAS 4 vs 69.3 ± 21.3% for SILVA), although no additional archaeal reference sequences were added to the MiDAS database in this study.MiDAS 4 was also able to assign species-level classifications to 40.8 ± 7.1% of the GWMC ASVs. In contrast, the 16S rRNA gene reference database obtained from GTDB SSU r89, which is the only universal reference database that contains a comprehensive species-level taxonomy, only classified 9.9 ± 2.0% of the ASVs (Fig. 3). For the ASVs created in this study, MiDAS 4 provided a species-level classification for 68.4 ± 6.1% of the V1–V3 and 48.5 ± 6.0% of the V4 ASVs (Supplementary Fig. 2a).Based on the large number of WWTPs sampled, their diversity, and the independent evaluation based on the GWMC dataset2, we expect that the MiDAS 4 reference database essentially covers the large majority of bacteria in WWTPs worldwide. Therefore, the MiDAS 4 taxonomy should act as a shared vocabulary for wastewater treatment microbiologists, providing opportunities for cross-study comparisons and ecological studies at high taxonomic resolution.Comparison of the V1–V3 and V4 primer sets for community profiling of WWTPsBefore investigating what factors shape the activated sludge microbiota, we compared short-read amplicon data created for all activated sludge samples belonging to the four main process types (C; C,N; C,N,DN and C,N,DN,P) collected in the Global MiDAS project using two commonly used primer sets that target the V1–V3 or V4 variable region of the 16S rRNA gene. The V1–V3 primers were chosen because the corresponding region of the 16S rRNA gene provides the highest taxonomic resolution of common short-read amplicons20,26, and these primers have previously shown great correspondence with metagenomic data and quantitative fluorescence in situ hybridisation (FISH) results for wastewater treatment systems17. The V4 region has a lower phylogenetic signal, but the primers used for amplification have better theoretical coverage of the bacterial diversity in the SILVA database20,26.The majority of genera (62%) showed less than twofold difference in relative abundances between the two primer sets, and the rest were preferentially detected with either the V1–V3 or the V4 primer (19% for both) (Fig. 4). We observed that several genera of known importance detected in high abundance by V1–V3 were hardly observed by V4, including Acidovorax, Rhodoferax, Ca. Villigracilis, Sphaerotilus and Leptothrix. Similarly, we observed genera abundant with V4 but strongly underestimated by V1–V3, such as Acinetobacter and Prosthecobacter. A complete list of differentially detected genera (Supplementary Data 2) serves as a valuable tool in combination with in silico primer evaluation for deciding which primer pair to use for targeted studies of specific taxa.Fig. 4: Comparison of relative genus abundance based on V1–V3 and V4 region 16S rRNA gene amplicon data.a Mean relative abundance was calculated based on 709 activated sludge samples. Genera present at ≥0.001% relative abundance in V1–V3 and/or V4 datasets are considered. Genera with less than twofold difference in relative abundance between the two primer sets are shown with gray circles, and those that are overrepresented by at least twofold with one of the primer sets are shown in red (V4) and blue (V1–V3). The twofold difference is an arbitrary choice; however, it relates to the uncertainty we usually encounter in amplicon data. Genus names are shown for all taxa present at a minimum of 0.1% mean relative abundance (excluding those with de novo names). b Heatmaps of the most abundant genera with more than twofold relative abundance difference between the two primer sets.Full size imageBecause the V1–V3 primers provide better classification rates at the genus- and species-level (Supplementary Fig. 2a), we primarily focused on this dataset for the following analyses. It should be noted that the V1–V3 primer set performs poorly on anammox bacteria27,28 and does not target archaea at all. To determine the importance of these groups, we estimated their relative read abundance using the V4 amplicon data. Ca. Brocadia and Ca. Anammoximicrobium were the only anammox genera detected, and the latter was never more than 0.6% abundant. Ca. Brocadia was observed in MBBR reactors and granular sludge in anammox reactors with relative read abundances reaching 29%, but it was below 0.1% relative abundance in all but two of the activated sludge samples investigated. For archaea, the relative read abundance was generally low (median = 0.18%), but for a few WWTPs high (up to 11.7%), so archaea should not be neglected in these cases.Process and environmental factors affecting the activated sludge microbiotaAlpha diversity analysis revealed that the rarefied (10,000 read per sample) richness and diversity in activated sludge plants were most strongly affected by process type, industrial load and continent (Supplementary Fig. 3 and Supplementary Note 1). The richness and diversity increased with the complexity of the treatment process, as found in other studies, reflecting the increased number of niches29. In contrast, it decreased with high industrial loads, presumably because industrial wastewater often is less complex and therefore promotes the growth of fewer specialised species7. The effect of continents is presumably caused by the necessary unbalanced sampling of WWTPs and confounded by the effects of plant types and industrial loads.Distance decay relationship (DDR) analyses were used to determine the effect of geographic distance on the microbial community similarity of activated sludge plants with the four main process types (Supplementary Fig. 4 and Supplementary Note 2). We found that distance decay was only effective within shorter geographical distances (2500 km) at the ASV-level, but higher similarities with OTUs clustered at 97% and even more at the genus-level. This suggests that many ASVs are geographically restricted and functionally redundant in the activated sludge microbiota, so different strains or species from the same genus across the world may provide similar functions.To gain a deeper understanding of the factors that shape the activated sludge microbiota, we examined the genus-level taxonomic beta-diversity using principal coordinate analysis (PCoA) and permutational multivariate analysis of variance (PERMANOVA) analyses (Fig. 5 and Supplementary Note 3). We have chosen taxonomic diversity instead of phylogenetic diversity (UniFrac) because many of the important traits are categorical (yes/no) and only conserved at lower taxonomic ranks (genus/species). The analysis was made at the genus-level due to the high classification rate achieved with MiDAS 4 and because genera were less affected by DDR compared to ASVs. We found that the overall microbial community was most strongly affected by continent and temperature in the WWTPs. However, process type, industrial load and the climate zone also had significant impacts. The percentage of total variation explained by each parameter was generally low, indicating that the global WWTPs microbiota represents a continuous distribution rather than distinct states, as observed for the human gut microbiota30.Fig. 5: Effects of process and environmental factors on the activated sludge microbial community structure. Principal coordinate analyses of Bray–Curtis and Soerensen beta-diversity for genera based on V1–V3 amplicon data. Samples are coloured based on metadata.The fraction of variation in the microbial community explained by each variable in isolation was determined by PERMANOVA (Adonis R2-values). Exact P values 0.1% relative abundance in 80% (strict core), 50% (general core) and 20% (loose core) of all activated sludge plants (Fig. 6a).Fig. 6: Identification of core and conditionally rare or abundant taxa based on V1–V3 amplicon data.a Identification of strict, general and loose core genera based on how often a given genus was observed at a relative abundance above 0.1% in WWTPs. b Identification of conditionally rare or abundant (CRAT) genera based on whether a given genus was observed at a relative abundance above 1% in at least one WWTP. The cumulative genus abundance is based on all ASVs classified at the genus-level. All core genera were removed before identification of the CRAT genera. c, d Number of genera and species, respectively, and their abundance in different process types across the global WWTPs. Values for genera and species are divided into strict core, general core, loose core, CRAT, other taxa and unclassified ASVs. The relative abundance of different groups was calculated based on the mean relative abundance of individual genera or species across samples. C carbon removal, C,N carbon removal with nitrification, C,N,DN carbon removal with nitrification and denitrification, C,N,DN,P carbon removal with nitrogen removal and enhanced biological phosphorus removal (EBPR).Full size imageIn addition to the core taxa, we also identified conditionally rare or abundant taxa (CRAT)32 (Fig. 6b). These are taxa typically present in low abundance but occasionally become prevalent, including taxa related to process disturbances, such as bacteria causing activated sludge foaming or those associated with the degradation of specific residues in industrial wastewater. CRAT have only been studied in a single WWTP treating brewery wastewater, despite their potential effect on performance32,33. CRAT are here defined as taxa which are not part of the core, but present in at least one WWTP with a relative abundance above 1%.Core taxa and CRAT were identified for both the V1–V3 and V4 amplicon data to ensure that critical taxa were not missed due to primer bias. We identified 250 core genera (15 strict, 65 general and 170 loose) and 715 CRAT genera (Supplementary Data 4). The strict core genera (Fig. 7) mainly contained genera with versatile metabolisms found in several environments, including Flavobacterium, Novosphingobium and Haliangium. The general core (Fig. 7) included many known bacteria associated with nitrification (Nitrosomonas and Nitrospira), polyphosphate accumulation (Tetrasphaera, Ca. Accumulibacter) and glycogen accumulation (Ca. Competibacter). The loose core contained well-known filamentous bacteria (Ca. Microthrix, Ca. Promineofilum, Ca. Sarcinithrix, Gordonia, Kouleothrix and Thiothrix), but also Nitrotoga, a less common nitrifier in WWTPs.Fig. 7: Percent relative abundance of strict and general core taxa across process types.The taxonomy for the core genera indicates phylum and genus. For general core species, genus names are also provided. De novo taxa in the core are highlighted in red. C carbon removal, C,N carbon removal with nitrification, C,N,DN carbon removal with nitrification and denitrification, C,N,DN,P carbon removal with nitrogen removal and enhanced biological phosphorus removal (EBPR).Full size imageBecause MiDAS 4 allowed for species-level classification, we also identified core and CRAT species based on the same criteria as for genera (Supplementary Fig. 7 and Supplementary Data 4). This revealed 113 core species (0 strict, 9 general and 104 loose). The general core species (Fig. 7) included Nitrospira defluvii and Tetrasphaera midas_s_5, a common nitrifier and PAO, respectively. Arcobacter midas_s_2255, a potential pathogen commonly abundant in the influent wastewater, was also part of the general core34. The loose core contained additional species associated with nitrification (Nitrosomonas midas_s_139 and Nitrospira nitrosa), polyphosphate accumulation (Ca. Accumulibacter phosphatis, Dechloromonas midas_s_173, Tetrasphaera midas_s_45), as well as known filamentous species (Ca. Microthrix parvicella and midas_s_2 (recently named Ca. M. subdominans35), Ca. Villigracilis midas_s_471 and midas_s_9223, Leptothrix midas_s_884). In addition to the core species, we identified 1417 CRAT species. As CRAT taxa are generally found in low abundance and the current study does not include time series or influent data, we cannot say anything conclusive about their general implications for the ecosystem. However, they may be present due to short-term mass immigration25 or specific operational conditions36 and in both cases, potentially affect the plant operation. They should therefore be considered important target for further investigations together with the core taxa.Many core taxa and CRAT can only be identified with MiDAS 4The core taxa and CRAT included a large proportion of MiDAS 4 de novo taxa. At the genus-level, 106/250 (42%) of the core genera and 500/715 (70%) of the CRAT genera had MiDAS placeholder names. At the species-level, the proportion was even higher. Here placeholder names were assigned to 101/113 (89%) of the core species and 1352/1417 (95%) CRAT species. This highlights the importance of a comprehensive taxonomy that includes the uncultured environmental taxa.The core and CRAT taxa cover the majority of the global activated sludge microbiotaAlthough the core taxa and CRAT represent a small fraction of the total diversity observed in the MiDAS 4 reference database, they accounted for the majority of the observed global activated sludge microbiota (Fig. 6c, d). Accumulated read abundance estimates ranged from 57–68% for the core genera and 11–13% for the CRAT, and combined they accounted for 68–79% of total read abundance in the WWTPs depending on process types. The core taxa represented a larger proportion of the activated sludge microbiota for the more advanced process types, which likely reflects the requirement of more versatile bacteria associated with the alternating redox conditions in these types of WWTPs. The remaining fraction, 21–32%, consisted of 6–8% unclassified genera and genera present in very low abundance, presumably with minor importance for the plant performance. The species-level core taxa and CRAT represented 11–24% and 24–33% accumulated read abundance, respectively. Combined, they accounted for almost 50% of the observed microbiota.Global diversity within important functional guildsThe general change from simple to advanced WWTPs with nutrient removal and the transition to water resource recovery facilities (WRRFs) requires increased knowledge about the bacteria responsible for the removal and recovery of nutrients, so we examined the global diversity of well-described nitrifiers, denitrifiers, PAOs and GAOs (Fig. 8). GAOs were included because they may compete with the PAOs for nutrients and thereby interfere with the biological recovery of phosphorus37. Because MiDAS 4 provided species-level resolution for a large proportion of activated sludge microbiota, we also investigated the species-level diversity within genera affiliated with the functional guilds. A complete overview of species in all genera detected in this global study is provided in the MiDAS field guide (https://www.midasfieldguide.org/guide).Fig. 8: Global diversity of genera belonging to major functional groups.The percent relative abundance represents the mean abundance for each country considering only WWTPs with the relevant process types. Countries are grouped based on continent (shifting colour).Full size imageNitrosomonas and potential comammox Nitrospira were the only abundant (≥0.1% average relative abundance) genera found among ammonia-oxidising bacteria (AOBs), whereas both Nitrospira and Nitrotoga were abundant among the nitrite oxidisers (NOBs), with Nitrospira being the most abundant across all countries (Fig. 8). Nitrobacter was not detected, and Nitrosospira was detected in only a few plants in very low abundance (≤0.01% average relative abundance). At the species-level, each genus had 2–5 abundant species (Supplementary Fig. 8). The most abundant and widespread Nitrosomonas species was midas_s_139. However, midas_s_11707 and midas_s_11733 were dominating in a few countries. For Nitrospira, the most abundant species in nearly all countries was N. defluvii. ASVs classified as the comammox N. nitrosa38,39 was also common in many countries across the world. However, because the comammox trait is not phylogenetically conserved at the 16S rRNA gene level38,39, we cannot conclude that these ASVs represent true comammox bacteria. For Nitrotoga, only two species were detected with notable abundance, midas_s_181 and midas_s_9575. Ammonia-oxidising archaea (AOAs) were not detected with MiDAS 4 due to the lack of reference sequences, and because AOAs are not targeted by the V1–V3 primer pair. However, analyses of our V4 amplicon dataset classified with the SILVA database revealed a considerable relative read abundance of AOAs in Malaysia and the Philippines, but absence or low abundance of AOAs in other countries (Supplementary Fig. 9). Other studies have occasionally found AOAs across the world, but generally in lower abundance than AOBs40,41,42. To ensure detection of AOAs with MiDAS 4, we anticipate adding external reference sequences for AOAs in a future release of the database.Denitrifying bacteria are very common in advanced activated sludge plants, but are generally poorly described. Among the known genera, Rhodoferax, Zoogloea and Thauera were most abundant (Fig. 8). Zoogloea and Thauera are well-known floc formers, sometimes causing unwanted slime formation43. Rhodoferax was the most common denitrifier in Europe, whereas Thauera dominated in Asia. Many denitrifiers could not be classified at the species-level (Supplementary Fig. 10), likely due to highly conserved 16S rRNA genes. An exception was Zoogloea, where midas_s_1080 and Z. caeni and were the most abundant species worldwide.EBPR is performed by PAOs, with three genera recognised as important in full-scale WWTPs: Tetrasphaera, Dechloromonas and Ca. Accumulibacter13. According to relative read abundance, all three were found in EBPR plants globally, with Tetrasphaera as the most prevalent (Fig. 8). Dechloromonas was also abundant in nitrifying and denitrifying plants without EBPR, indicating a more diverse ecology. Four recognised GAOs were found globally: Ca. Competibacter, Defluviicoccus, Propionivibrio and Micropruina, with Ca. Competibacter being the most abundant (Fig. 8). Only a few species (2–6 species) in each genus were dominant across the world for both PAOs (Supplementary Fig. 11) and GAOs (Supplementary Fig. 12), except for Ca. Competibacter, which covered ~20 abundant but country-specific species. Among PAOs, the abundant species were Tetrasphaera midas_s_5, Dechloromonas midas_s_173, (recently named D. phosphorivorans) Ca. Accumulibacter midas_s_315, Ca. A. phosphatis and Ca. A. aalborgensis. Interestingly, some of the most abundant PAOs and GAOs were also abundant in the simple process design with C-removal, indicating more versatile metabolisms.Global diversity of filamentous bacteriaFilamentous bacteria are essential for creating strong activated sludge flocs. However, in large numbers, they can also lead to loose flocs and poor settling properties. This is known as bulking, a major operational problem in many WWTPs. Many can also form foam on top of process tanks due to hydrophobic surfaces. Presently, approximately 20 genera are known to contain filamentous species44, and among those, the most abundant are Ca. Microthrix, Leptothrix, Ca. Villigracilis, Trichococcus and Sphaerotilus (Fig. 9). They are all well-known from studies on mitigation of poor settling properties in WWTPs. Interestingly, Leptothrix, Sphaerotilus and Ca. Villigracilis belong to the genera where abundance-estimation depended strongly on primers, with V4 underestimating their abundance (Fig. 3). Ca. Microthrix and Leptothrix were strongly associated with continents, most common in Europe and less in Asia and North America (Fig. 9).Fig. 9: Global diversity of known filamentous organisms.The percent relative abundance represents the mean abundance for each country across all process types. Countries are grouped based on the continent (shifting colour).Full size imageMany of the filamentous bacteria were linked to specific process types (Supplementary Fig. 13), e.g. Ca. Microthrix were not observed in WWTPs with carbon removal only, and Ca. Amarolinea were only abundant in plants with nutrient removal. The number of abundant species within the genera were generally low, with one species in Trichococcus, two in Ca. Microthrix and approximately five in Leptothrix and Ca. Villigracilis (Supplementary Fig. 14). Only five abundant species were observed for Sphaerotilus. However, a substantial fraction of unclassified ASVs was also observed, demonstrating that certain species within this genus are poorly resolved based on the 16S rRNA gene. Ca. Promineofilum was also poorly resolved at the species-level (Supplementary Fig. 15).Conclusion and perspectivesWe present a worldwide collaborative effort to produce MiDAS 4, an ASV-resolved full-length 16S rRNA gene reference database, which covers more than 31,000 species and enables genus- to species-level resolution in microbial community profiling studies. MiDAS 4 covers the vast majority of WWTP bacteria globally and provides a strongly needed common taxonomy for the field, which provides the foundation for comprehensive linking of microbial taxa in the ecosystem with their functional traits. Presently, hundreds of studies are undertaken to combine engineering and microbial aspects of full-scale WWTPs. However, most ASVs or OTUs in these studies are classified at poor taxonomic resolution (family-level or above) due to the use of incomplete universal reference databases. Because many important functional traits are only conserved at high taxonomic resolution (genus- or species-level), this strongly hampers our ability to transfer taxa-specific knowledge from one study to another. This will change with MiDAS 4, and we expect that reprocessing of data from earlier studies may reveal new perspectives into wastewater treatment microbiology. Our online Global MiDAS Field Guide presents the data generated in this study and summarises present knowledge about all taxa. We encourage researchers within the field to contribute new knowledge to MiDAS using the contact link in the MiDAS website (https://www.midasfieldguide.org/guide/contact).The global microbiota of activated sludge plants has been predicted to harbour a massive diversity with up to one billion species2. However, most of these occur at very low abundance and are of little importance for the treatment process. By focusing only on the abundant taxa, we can see that this number is much smaller, i.e., ~1000 genera and 1500 species. We consider these taxa functionally the most important globally, representing a “most wanted list” for future studies. Some taxa are abundant in most WWTPs (core taxa), and others are occasionally abundant in fewer plants (CRAT). The CRAT have received little attention in the field of wastewater treatment, but they can be of profound importance for WWTP performance. Both groups have a high fraction of poorly characterised species. The high taxonomic resolution provided by MiDAS 4 enables us to identify samples where these important core taxa occur in high abundance. This provides an ideal starting point for obtaining high-quality metagenome-assembled genomes (MAGs), isolation of pure cultures, in addition to targeted culture-independent studies to uncover their physiological and ecological roles.Among the known functional guilds, such as nitrifiers or polyphosphate-accumulating organisms, the same genera were found worldwide, with only a few abundant species in each genus. There were differences in the community structure, and the abundance of dominant species was mainly shaped by process type, temperature, and in some cases, continent. This discovery sends an important message to the field: relatively few species are abundant worldwide, so research or operational results can reliably be transferred from one geographical region to another, stimulating the transition from WWTPs to more sustainable WRRFs.The relatively low number of uncharacterised abundant species also shows that it is within our reach to describe them all in terms of identity, physiology, ecology and dynamics, providing the necessary knowledge for informed process optimisation and management. The number of poorly described genera (i.e. those with only a MiDAS placeholder genus name) was 88 among the 250 core genera (35%) and more than 89% at the species-level, so there is still some work to do to link their identities and function. An important step in this direction is the visualisation of the populations. With the comprehensive set of FL-ASVs, it is possible to design highly specific FISH probes, and to critically evaluate the old probes. In the Danish WWTPs, we have successfully done this for groups in the Acidobacteriota42 based on the MiDAS 3 database18. Our recent retrieval of more than 1000 high-quality MAGs from Danish WWTPs with advanced process design is also an important step to link identity to function43. The HQ-MAGs can be linked directly to MiDAS 4 as they contain complete 16S rRNA genes. They cover 62% (156/250) of the core genera and 61% (69/113) of the core species identified in this study. These MAGs may also form the basis for further studies to link identity and function, e.g. by applying metatranscriptomics44 and other in situ techniques such as FISH combined with Raman45,46, guided by the “most wanted” list provided in this study. We expect that MiDAS 4 will have significant implications for future microbial ecology studies in wastewater treatment systems. More

  • in

    System dynamics modeling of lake water management under climate change

    System dynamics methodThe SD method applies systemic processing to simulate complex non-linear dynamics and feedback. Systemic processing resorts to various tools to simulate complex system behavior and performance24. Systems evolve through states, which change with flows. An example of a state variable is water storage in the study of lakes. The SD method simulates changes in system states driven by flows and various feedbacks25.This work employs the SD method to simulate storage change in Lake Urmia in one historical period (1957–2005) and two future periods (2021–2050 and 2051–2080). The lake’s water volume is the state variable, which is governed by inflows (precipitation, surface water inflows, and groundwater inflows) and outflows (evaporation, leakage, and surface water outflows). The lake’s mass balance equation is expressed as:$$S_{t + 1} = intlimits_{t}^{t + 1} {[I_{s} – O_{s} ]ds + S_{t} }$$
    (1)
    where St+1 , St, Is, and Os denote the lake’s storage at time t + 1, the lake’s storage at time t, the inflow rate to the lake at time s (units of volume/time), and the outflow rate from the lake at time s (units of volume/time), respectively.The SD method employs the Euler and Runge Kutta methods for the solution of differential equations. The software STELLA, Vensim, Powersim, and Dynamo feature SD solvers26. This work applies the widely-used Vensim software27.Climate changeThe data sets needed for modeling Lake Urmia’s storage over the two future periods were generated after simulating the lake’s water balance during the historical period. HADCM3, a coupled atmosphere–ocean general circulation model’s (AOGCM) climate projections were used to generate precipitation and surface temperature projections over the future periods. The AOGCM data at coarse spatial scales were downscaled to the regional scale suitable for lake storage simulation. The commonly used downscaling methods are statistic and dynamic in nature28,29. This works applies the delta-change downscaling method, in which monthly temperature and precipitation differences between the future and historical are calculated by29:$$Delta T_{t} = overline{T}_{GCM,fut,t} – overline{T}_{GCM,hist,t}$$
    (2)
    $$Delta P_{t} = overline{P}_{GCM,fut,t} – overline{P}_{GCM,hist,t}$$
    (3)
    where ∆Tt denotes the difference in long-term average temperatures simulated by HADCM3 for the future ((overline{T}_{GCM,fut,t})) and historical ((overline{T}_{GCM,hist,t})) periods in month t (°C); ∆Pt represents the difference in long-term average precipitations simulated by HADCM3 for the future ((overline{P}_{GCM,fut,t})) and historical ((overline{P}_{GCM,hist,t})) periods in month t (mm). Then, ∆Tt and ∆Pt are applied to project the future downscaled data as follows29:$$T_{t} = T_{obs,t} + , Delta T_{t}$$
    (4)
    $$P_{t} = P_{obs,t} { + }Delta P_{t}$$
    (5)
    where Tobs,t, and Pobs,t denote respectively the observed temperature (°C) and precipitation (mm) in month t in the baseline period; and Tt and Pt are the downscaled temperature (°C) and precipitation (mm) in month t of the future period, respectively. Delta-change downscaling is a simple yet efficient option when it comes to spatial downscaling of climate change projections (e.g.30,31,32). The gist of this method is to replicate the changing patterns that are projected by the atmospheric ocean general circulation models (AOGCMs) to generate the climate change patterns of hydro-climatic variables on a regional scale. As such, one would simply compute the relative changes in the long-term variations of the variable that is projected by the models within the baseline and future timeframes. These relative changing patterns would be applied to the historical data to project the impact of climate change on a local scale.Rainfall-runoff modelingThe IHACRES (identification of unit hydrographs and component flows from rainfall, evapotranspiration and streamflow) model is herein applied to simulate runoff from precipitation. Ashofteh et al.33 implemented the IHACRES model to investigate the effects of climate change on reservoir performance in agricultural water supply. Ashofteh et al.34 evaluated the probability of flood occurrence in future periods with IHACRES.The IHACRES model includes a non-linear loss module and a linear unit hydrograph module. The non-linear loss module converts the observed rainfall into the effective rainfall, after which the linear unit hydrograph module converts the effective rainfall into the simulated streamflow35. Here, precipitation rk in time step k is converted to effective precipitation uk through the non-linear loss module employing a catchment wetness index sk:$$u_{k} = , s_{k} times , r_{k}$$
    (6)
    The effective precipitation is converted to the surface runoff in time step k with the linear unit hydrograph module. The parameters of this model can be set through a thorough grid numeric search and trial-and-error. Perhaps, one of the major advantages of the IHACRES model over other commonly-used rainfall-runoff models is its minimal input data requirement (i.e., air temperature and precipitation)31,35.The other alternative for hydrologic simulation is to use data-driven models. Here, the multilayer perceptron (MLP), a variety of the artificial neural network (ANN) method, was also used to simulate runoff. This model consists of an inlet layer, one or several middle (hidden) layer(s), and an output layer. All of the neurons of a layer are connected to the ones in the next layer, forming a network with complete connections. The primary parameters in modeling the neural network of MLP are: (1) the number of neurons in each layer, (2) the number of layers in the network, and (3) the forcing functions. A regular MLP neural network has three layers36. The first and the third layers are respectively the system inputs and outputs. The middle layer consists of neurons that perform calculations on the inputs. Choosing the number of layers in a neural network is made by trial and error37. From a hydrological simulation standpoint the main idea behind this model is to create a suitable artificial neural network that is capable of accurately converting a set of hydro-climatic variables such as precipitation and temperature as input data into streamflow values. It should be noted that, like most data-driven models, the process of opting for a proper neural network architecture (i.e., selecting the number of layers, number of neurons, and the forcing function) is, for the most part, a trial-and-error procedure.One must objectively evaluate the performance of the hydrological models in order to opt for the setting of a suitable parameter. The root mean square error (RMSE), coefficient of determination (R2), and mean absolute error (MAE) are herein employed to assess the performance of the rainfall-runoff model. They are respectively calculated as follows:$$RMSE = sqrt {frac{{sumlimits_{t = 1}^{N} {(x_{t} – y_{t} )^{2} } }}{N}}$$
    (7)
    $$R^{2} = left( {frac{{sumnolimits_{t = 1}^{N} {(x_{t} – overline{x} ).(y_{t} – overline{y} )} }}{{sqrt {sumnolimits_{t = 1}^{N} {(x_{t} – overline{x} )^{2} } } .sqrt {sumnolimits_{t = 1}^{N} {(y_{t} – overline{y} )^{2} } } }}} right)^{2}$$
    (8)
    $$MAE = frac{{sumnolimits_{t = 1}^{N} {left| {x_{t} – y_{t} } right|} }}{N}$$
    (9)
    where xt , yt, and N denote the simulated value in time step t; the observed value in time step t; and the number data values, respectively. Large errors have a disproportionately large effect on RMSE or MAE.Performance criteriaVarious quantitative measures can be used to assess the performance of water resources systems under different strategies. When it comes to water resources planning and management, perhaps, some of the most common performance criteria are the probability-based performance criteria (PBPC) (i.e., reliability, vulnerability, and resiliency)31,38. In this context, reliability represents the probability of successful functioning of a system; resiliency measures the probability of successful functioning following a system failure; lastly, vulnerability is the severity of failure during an operation horizon39,40. The basic idea behind a performance evaluation attribute is to provide a quantitative measure to describe and assess the performance of a system. In the context of water resources planning and management, these measures have proven time and again that they can be reliable options to evaluate a set of strategic management options objectively (see, e.g.40,41,42,43, and44, just to name a few).Operating policyAny water resources system requires something called the “rule curve,” which determines how water is allocated in a given situation45. A common and effective rule curve when it comes to operation of water resource systems is the standard operation policy (SOP). SOP is a simple, and perhaps best-known real-time operation policy in water resources planning and management46. The core principle here is to minimize the water shortage at the current time step with no conservation policy (e.g., hedging rules) in place. The SOP, as a standard rule curve, determines how the operator acts to control a system at any given state of a reservoir47,48. This rule curve is established as an attempt to balance various water demands including but not limited to flood control, hydropower, water supply, and recreation49. A SOP operating system attempts to release water to meet a water demand at the current time, with no regard to the future. Thus, according to the SOP’s principle, the decision-makers, first allocate the available water to meet the demand of the stakeholder with the highest priority. After this first water demand is fully satisfied, the available water can be used for the next demand. Such an allocation process continues until no water is available.Ethics approvalAll authors accept all ethical approvals.Consent to participateAll authors consent to participate.Consent to publishAll authors consent to publish. More

  • in

    Linking transcriptional dynamics of CH4-cycling grassland soil microbiomes to seasonal gas fluxes

    Canadell JG, Monteiro PMS, Costa, MH, Cotrim da Cunha L, Cox PM, Eliseev AV, et al. Global carbon and other biogeochemical cycles and feedbacks. In: Masson-Delmotte V, Zhai P, Pirani A, Connors SL, Péan C, Berger S, et al. editors. Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press; 2021, in press.Rosentreter JA, Borges AV, Deemer BR, Holgerson MA, Liu S, Song C, et al. Half of global methane emissions come from highly variable aquatic ecosystem sources. Nat Geosci. 2021;14:225–30.CAS 

    Google Scholar 
    Saunois M, Stavert AR, Poulter B, Bousquet P, Canadell JG, Jackson RB, et al. The global methane budget 2000 – 2017. Earth Syst. Sci Data. 2020;12:1561–623.
    Google Scholar 
    Lamentowicz M, Gałka M, Pawlyta J, Lamentowicz Ł, Goslar T, Miotk-Szpiganowicz G. Climate change and human impact in the southern Baltic during the last millennium reconstructed from an ombrotrophic bog archive. Stud Quat. 2011;28:3–16.
    Google Scholar 
    Davidson NC. How much wetland has the world lost? Long-term and recent trends in global wetland area. Mar Freshw Res. 2014;65:934–41.
    Google Scholar 
    Oertel C, Matschullat J, Zurba K, Zimmermann F, Erasmi S. Greenhouse gas emissions from soils – a review. Geochemistry. 2016;76:327–52.CAS 

    Google Scholar 
    Liesack W, Schnell S, Revsbech NP. Microbiology of flooded rice paddies. FEMS Microbiol Rev. 2000;24:625–45.CAS 
    PubMed 

    Google Scholar 
    Conrad R. The global methane cycle: recent advances in understanding the microbial processes involved. Environ Microbiol Rep. 2009;1:285–92.CAS 
    PubMed 

    Google Scholar 
    Lyu Z, Shao N, Akinyemi T, Whitman WB. Methanogenesis. Curr Biol. 2018;28:R727–R732.CAS 
    PubMed 

    Google Scholar 
    Kurth JM, Nobu MK, Tamaki H, de Jonge N, Berger S, Jetten MSM, et al. Methanogenic archaea use a bacteria-like methyltransferase system to demethoxylate aromatic compounds. ISME J. 2021;15:3549–65.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Mayumi D, Mochimaru H, Tamaki H, Yamamoto K, Yoshioka H, Suzuki Y, et al. Methane production from coal by a single methanogen. Science. 2016;354:222–6.CAS 
    PubMed 

    Google Scholar 
    Bridgham SD, Cadillo-Quiroz H, Keller JK, Zhuang Q. Methane emissions from wetlands: Biogeochemical, microbial, and modeling perspectives from local to global scales. Glob Chang Biol. 2013;19:1325–46.PubMed 

    Google Scholar 
    Narrowe AB, Borton MA, Hoyt DW, Smith GJ, Daly RA, Angle JC, et al. Uncovering the diversity and activity of methylotrophic methanogens in freshwater wetland soils. mSystems. 2019;4:e00320–19.PubMed 
    PubMed Central 

    Google Scholar 
    Zalman CA, Meade N, Chanton J, Kostka JE, Bridgham SD, Keller JK. Methylotrophic methanogenesis in Sphagnum-dominated peatland soils. Soil Biol Biochem. 2018;118:156–60.CAS 

    Google Scholar 
    Knief C. Diversity and habitat preferences of cultivated and uncultivated aerobic methanotrophic bacteria evaluated based on pmoA as molecular marker. Front Microbiol. 2015;6:1346.PubMed 
    PubMed Central 

    Google Scholar 
    Le Mer J, Roger P. Production, oxidation, emission and consumption of methane by soils: a review. Eur J Soil Biol. 2001;37:25–50.
    Google Scholar 
    Wieczorek AS, Drake HL, Kolb S. Organic acids and ethanol inhibit the oxidation of methane by mire methanotrophs. FEMS Microbiol Ecol. 2011;77:28–39.CAS 
    PubMed 

    Google Scholar 
    Welte CU, Rasigraf O, Vaksmaa A, Versantvoort W, Arshad A, Op den Camp HJM, et al. Nitrate- and nitrite-dependent anaerobic oxidation of methane. Environ Microbiol Rep. 2016;8:941–55.CAS 
    PubMed 

    Google Scholar 
    Cui M, Ma A, Qi H, Zhuang X, Zhuang G. Anaerobic oxidation of methane: An ‘active’ microbial process. Microbiol Open. 2015;4:1–11.
    Google Scholar 
    Ettwig KF, Zhu B, Speth D, Keltjens JT, Jetten MSM, Kartal B. Archaea catalyze iron-dependent anaerobic oxidation of methane. Proc Natl Acad Sci USA. 2016;113:12792–6.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Stiehl-Braun PA, Hartmann AA, Kandeler E, Buchmann N, Niklaus PA. Interactive effects of drought and N fertilization on the spatial distribution of methane assimilation in grassland soils. Glob Chang Biol 2011;17:2629–39.
    Google Scholar 
    Bodelier PLE, Meima-Franke M, Hordijk CA, Steenbergh AK, Hefting MM, Bodrossy L, et al. Microbial minorities modulate methane consumption through niche partitioning. ISME J. 2013;7:2214–28.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Karbin S, Hagedorn F, Dawes MA, Niklaus PA. Treeline soil warming does not affect soil methane fluxes and the spatial micro-distribution of methanotrophic bacteria. Soil Biol Biochem. 2015;86:164–71.CAS 

    Google Scholar 
    Stiehl-Braun PA, Powlson DS, Poulton PR, Niklaus PA. Effects of N fertilizers and liming on the micro-scale distribution of soil methane assimilation in the long-term Park Grass experiment at Rothamsted. Soil Biol Biochem. 2011;43:1034–41.CAS 

    Google Scholar 
    Menyailo OV, Hungate BA, Abraham WR, Conrad R. Changing land use reduces soil CH4 uptake by altering biomass and activity but not composition of high-affinity methanotrophs. Glob Chang Biol. 2008;14:2405–19.
    Google Scholar 
    Ciais P, Sabine C, Bala G, Bopp L, Brovkin V, Canadell J, et al. Carbon and Other Biogeochemical Cycles. In: Stocker TF, Qin D, Plattner G-K, Tignor M, Allen SK, Boschung J, et al. editors. Climate Change 2013 the Physical Science Basis: Working Group I Contribution to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. New York, NY: Cambridge University Press; 2013, 465–570.Täumer J, Kolb S, Boeddinghaus RS, Wang H, Schöning I, Schrumpf M, et al. Divergent drivers of the microbial methane sink in temperate forest and grassland soils. Glob Chang Biol. 2021;27:929–40.PubMed 

    Google Scholar 
    Kolb S. The quest for atmospheric methane oxidizers in forest soils. Environ Microbiol Rep. 2009;1:336–46.CAS 
    PubMed 

    Google Scholar 
    Kolb S, Horn MA. Microbial CH4 and N2O consumption in acidic wetlands. Front Microbiol. 2012;3:78.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Cai Y, Zheng Y, Bodelier PLE, Conrad R, Jia Z. Conventional methanotrophs are responsible for atmospheric methane oxidation in paddy soils. Nat Commun. 2016;7:11728.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Dean JF, Middelburg JJ, Röckmann T, Aerts R, Blauw LG, Egger M, et al. Methane feedbacks to the global climate system in a warmer world. Rev Geophys. 2018;56:207–50.
    Google Scholar 
    Levy-Booth DJ, Giesbrecht IJW, Kellogg CTE, Heger TJ, D’Amore DV, Keeling PJ, et al. Seasonal and ecohydrological regulation of active microbial populations involved in DOC, CO2, and CH4 fluxes in temperate rainforest soil. ISME J. 2019;13:950–63.CAS 
    PubMed 

    Google Scholar 
    Lombard N, Prestat E, van Elsas JD, Simonet P. Soil-specific limitations for access and analysis of soil microbial communities by metagenomics. FEMS Microbiol Ecol. 2011;78:31–49.CAS 
    PubMed 

    Google Scholar 
    Carini P, Marsden PJ, Leff JW, Morgan EE, Strickland MS, Fierer N. Relic DNA is abundant in soil and obscures estimates of soil microbial diversity. Nat Microbiol. 2016;2:16242.PubMed 

    Google Scholar 
    Blazewicz SJ, Barnard RL, Daly RA, Firestone MK. Evaluating rRNA as an indicator of microbial activity in environmental communities: limitations and uses. ISME J. 2013;7:2061–8.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Sukenik A, Kaplan-Levy RN, Welch JM, Post AF. Massive multiplication of genome and ribosomes in dormant cells (akinetes) of Aphanizomenon ovalisporum (Cyanobacteria). ISME J. 2012;6:670–9.CAS 
    PubMed 

    Google Scholar 
    Schwartz E, Hayer M, Hungate BA, Koch BJ, McHugh TA, Mercurio W, et al. Stable isotope probing with 18O-water to investigate microbial growth and death in environmental samples. Curr Opin Biotechnol. 2016;41:14–18.CAS 
    PubMed 

    Google Scholar 
    Angel R, Conrad R. Elucidating the microbial resuscitation cascade in biological soil crusts following a simulated rain event. Environ Microbiol. 2013;15:2799–815.CAS 
    PubMed 

    Google Scholar 
    Papp K, Mau RL, Hayer M, Koch BJ, Hungate BA, Schwartz E. Quantitative stable isotope probing with H218O reveals that most bacterial taxa in soil synthesize new ribosomal RNA. ISME J. 2018;12:3043–5.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Urich T, Lanzén A, Qi J, Huson DH, Schleper C, Schuster SC. Simultaneous assessment of soil microbial community structure and function through analysis of the meta-transcriptome. PLoS One. 2008;3:e2527.PubMed 
    PubMed Central 

    Google Scholar 
    Peng J, Wegner CE, Liesack W. Short-term exposure of paddy soil microbial communities to salt stress triggers different transcriptional responses of key taxonomic groups. Front Microbiol. 2017;8:400.PubMed 
    PubMed Central 

    Google Scholar 
    Peng J, Wegner CE, Bei Q, Liu P, Liesack W. Metatranscriptomics reveals a differential temperature effect on the structural and functional organization of the anaerobic food web in rice field soil. Microbiome. 2018;6:169.PubMed 
    PubMed Central 

    Google Scholar 
    Abdallah RZ, Wegner CE, Liesack W. Community transcriptomics reveals drainage effects on paddy soil microbiome across all three domains of life. Soil Biol Biochem. 2019;132:131–42.CAS 

    Google Scholar 
    Gloor GB, Macklaim JM, Pawlowsky-Glahn V, Egozcue JJ. Microbiome datasets are compositional: and this is not optional. Front Microbiol. 2017;8:2224.PubMed 
    PubMed Central 

    Google Scholar 
    Moran MA, Satinsky B, Gifford SM, Luo H, Rivers A, Chan LK, et al. Sizing up metatranscriptomics. ISME J. 2013;7:237–43.CAS 
    PubMed 

    Google Scholar 
    Gifford SM, Sharma S, Rinta-Kanto JM, Moran MA. Quantitative analysis of a deeply sequenced marine microbial metatranscriptome. ISME J. 2011;5:461–72.PubMed 

    Google Scholar 
    Söllinger A, Tveit AT, Poulsen M, Noel SJ, Bengtsson M, Bernhardt J, et al. Holistic assessment of rumen microbiome dynamics through quantitative metatranscriptomics reveals multifunctional redundancy during key steps of anaerobic feed degradation. mSystems 2018;3:e00038–18.PubMed 
    PubMed Central 

    Google Scholar 
    Fischer M, Bossdorf O, Gockel S, Hänsel F, Hemp A, Hessenmöller D, et al. Implementing large-scale and long-term functional biodiversity research: the biodiversity exploratories. Basic Appl Ecol. 2010;11:473–85.
    Google Scholar 
    IUSS Working Group WRB. World reference base for soil resources 2014, update 2015 international soil classification system for naming soils and creating legends for soil maps. World Soil Resour Reports No 106. Rome: FAO; 2015.Vance ED, Brookes PC, Jenkinson DS. An extraction method for measuring soil microbial biomass C. Soil Biol Biochem. 1987;19:703–7.CAS 

    Google Scholar 
    Joergensen RG, Mueller T. The fumigation-extraction method to estimate soil microbial biomass: calibaration of the kEN value. Soil Biol Biochem. 1996;28:33–37.CAS 

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

    Google Scholar 
    Bamminger C, Zaiser N, Zinsser P, Lamers M, Kammann C, Marhan S. Effects of biochar, earthworms, and litter addition on soil microbial activity and abundance in a temperate agricultural soil. Biol Fertil Soils. 2014;50:1189–1200.CAS 

    Google Scholar 
    Koch O, Tscherko D, Kandeler E. Seasonal and diurnal net methane emissions from organic soils of the Eastern Alps, Austria: Effects of soil temperature, water balance, and plant biomass. Arct Antarct Alp Res. 2007;39:438–48.
    Google Scholar 
    Tveit AT, Urich T, Svenning MM. Metatranscriptomic analysis of arctic peat soil microbiota. Appl Environ Microbiol. 2014;80:5761–72.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Magoč T, Salzberg SL. FLASH: fast length adjustment of short reads to improve genome assemblies. Bioinformatics. 2011;27:2957–63.PubMed 
    PubMed Central 

    Google Scholar 
    Schmieder R, Edwards R. Quality control and preprocessing of metagenomic datasets. Bioinformatics. 2011;27:863–4.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kopylova E, Noé L, Touzet H. SortMeRNA: Fast and accurate filtering of ribosomal RNAs in metatranscriptomic data. Bioinformatics. 2012;28:3211–7.CAS 
    PubMed 

    Google Scholar 
    Edgar RC. Search and clustering orders of magnitude faster than BLAST. Bioinformatics. 2010;26:2460–1.CAS 
    PubMed 

    Google Scholar 
    Lanzén A, Jørgensen SL, Huson DH, Gorfer M, Grindhaug SH, Jonassen I, et al. CREST – Classification resources for environmental sequence tags. PLoS One. 2012;7:e49334.Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ. Basic local alignment search tool. J Mol Biol. 1990;215:403–10.CAS 
    PubMed 

    Google Scholar 
    Huson DH, Beier S, Flade I, Górska A, El-Hadidi M, Mitra S, et al. MEGAN community edition – interactive exploration and analysis of large-scale microbiome sequencing data. PLoS Comput Biol. 2016;12:e1004957.PubMed 
    PubMed Central 

    Google Scholar 
    Buchfink B, Xie C, Huson DH. Fast and sensitive protein alignment using DIAMOND. Nat Methods. 2014;12:59–60.PubMed 

    Google Scholar 
    Dumont MG, Lüke C, Deng Y, Frenzel P. Classification of pmoA amplicon pyrosequences using BLAST and the lowest common ancestor method in MEGAN. Front Microbiol. 2014;5:34.PubMed Central 

    Google Scholar 
    R Core Team. R: a language and environment for statistical computing. Vienna, Austria: R Foundation for Statistical Computing; 2018.Oksanen J, Blanchet f. G, Friendly M, Kindt R, Legendre P, McGlinn D, et al. vegan: Community ecology package. 2020. R package version 2.5-7. https://CRAN.R-project.org/package=vegan.Graves S, Piepho H-P, Selzer L. multcompView: Visualizations of paired comparisons. 2019. R package version 0.1-8. https://CRAN.R-project.org/package=multcompView.Günther A, Barthelmes A, Huth V, Joosten H, Jurasinski G, Koebsch F, et al. Prompt rewetting of drained peatlands reduces climate warming despite methane emissions. Nat Commun. 2020;11:1644.PubMed 
    PubMed Central 

    Google Scholar 
    IPCC Task Force on National Greenhouse Gas Inventories. Methodological guidance on lands with wet and drained soilds, and constructed wetlands for wastewater treatment. 2013 Supplement to the 2006 IPCC Guidelines for National Greenhouse Gas Inventories: Wetlands. 2014.Tiemeyer B, Albiac Borraz E, Augustin J, Bechtold M, Beetz S, Beyer C, et al. High emissions of greenhouse gases from grasslands on peat and other organic soils. Glob Chang Biol. 2016;22:4134–49.PubMed 

    Google Scholar 
    Kirschbaum MUF. The temperature dependence of soil organic matter decomposition, and the effect of global warming on soil organic C storage. Soil Biol Biochem. 1995;27:753–60.CAS 

    Google Scholar 
    Knorr W, Prentice IC, House JI, Holland EA. Long-term sensitivity of soil carbon turnover to warming. Nature 2005;433:298–301.CAS 
    PubMed 

    Google Scholar 
    Dutaur L, Verchot LV. A global inventory of the soil CH4 sink. Glob Biogeochem Cycles. 2007;21:GB4013.
    Google Scholar 
    McDaniel MD, Saha D, Dumont MG, Hernández M, Adams MA. The effect of land-use change on soil CH4 and N2O fluxes: A global meta-analysis. Ecosystems. 2019;22:1424–43.CAS 

    Google Scholar 
    Gulledge J, Schimel JP. Moisture control over atmospheric CH4 consumption and CO2 production in diverse Alaskan soils. Soil Biol Biochem. 1998;30:1127–32.CAS 

    Google Scholar 
    Tveit AT, Urich T, Frenzel P, Svenning MM. Metabolic and trophic interactions modulate methane production by Arctic peat microbiota in response to warming. Proc Natl Acad Sci USA. 2015;112:E2507–E2516.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Conrad R. Methane production in soil environments – anaerobic biogeochemistry and microbial life between flooding and desiccation. Microorganisms 2020;8:881.CAS 
    PubMed Central 

    Google Scholar 
    Lyu Z, Lu Y. Metabolic shift at the class level sheds light on adaptation of methanogens to oxidative environments. ISME J. 2018;12:411–23.PubMed 

    Google Scholar 
    Smith KS, Ingram-Smith C. Methanosaeta, the forgotten methanogen? Trends Microbiol. 2007;15:150–5.CAS 
    PubMed 

    Google Scholar 
    Whitman WB, Bowen TL, Boone DR. The methanogenic bacteria. In: Rosenberg E, DeLong EF, Lory S, Stackebrandt E, Thompson F editors. The prokaryotes: other major lineages of bacteria and the archaea. Berlin, Heidelberg: Springer; 2014, pp 123–63.Conrad R. Importance of hydrogenotrophic, aceticlastic and methylotrophic methanogenesis for methane production in terrestrial, aquatic and other anoxic environments: a mini review. Pedosphere. 2020;30:25–39.
    Google Scholar 
    Söllinger A, Urich T. Methylotrophic methanogens everywhere – physiology and ecology of novel players in global methane cycling. Biochem Soc Trans. 2019;47:1895–907.PubMed 

    Google Scholar 
    Yang S, Liebner S, Winkel M, Alawi M, Horn F, Dörfer C, et al. In-depth analysis of core methanogenic communities from high elevation permafrost-affected wetlands. Soil Biol Biochem. 2017;111:66–77.CAS 

    Google Scholar 
    Weil M, Wang H, Bengtsson M, Köhn D, Günther A, Jurasinski G, et al. Long-term rewetting of three formerly drained peatlands drives congruent compositional changes in pro- and eukaryotic soil microbiomes through environmental filtering. Microorganisms. 2020;8:550.CAS 
    PubMed Central 

    Google Scholar 
    Söllinger A, Seneca J, Dahl MB, Motleleng LL, Prommer J, Verbruggen E, et al. Down-regulation of the microbial protein biosynthesis machinery in response to weeks, years, and decades of soil warming. Sci Adv. 2022;8:eabm3230.PubMed 
    PubMed Central 

    Google Scholar 
    Luesken FA, Wu ML, Op den Camp HJM, Keltjens JT, Stunnenberg H, Francoijs KJ, et al. Effect of oxygen on the anaerobic methanotroph ‘Candidatus Methylomirabilis oxyfera’: Kinetic and transcriptional analysis. Environ Microbiol. 2012;14:1024–34.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Baani M, Liesack W. Two isozymes of particulate methane monooxygenase with different methane oxidation kinetics are found in Methylocystis sp. strain SC2. Proc Natl Acad Sci. 2008;105:10203–8.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Yimga MT, Dunfield PF, Ricke P, Heyer J, Liesack W. Wide distribution of a novel pmoA-like gene copy among type II methanotrophs, and its expression in Methylocystis strain SC2. Appl Environ Microbiol. 2003;69:5593–602.CAS 

    Google Scholar 
    Tveit AT, Hestnes AG, Robinson SL, Schintlmeister A, Dedysh SN, Jehmlich N, et al. Widespread soil bacterium that oxidizes atmospheric methane. Proc Natl Acad Sci USA. 2019;116:8515–24.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Freitag TE, Toet S, Ineson P, Prosser JI. Links between methane flux and transcriptional activities of methanogens and methane oxidizers in a blanket peat bog. FEMS Microbiol Ecol. 2010;73:157–65.CAS 
    PubMed 

    Google Scholar 
    Qin H, Tang Y, Shen J, Wang C, Chen C, Yang J, et al. Abundance of transcripts of functional gene reflects the inverse relationship between CH4 and N2O emissions during mid-season drainage in acidic paddy soil. Biol Fertil Soils. 2018;54:885–95.
    Google Scholar  More

  • in

    Impact of rice paddy agriculture on habitat usage of migratory shorebirds at the rice paddy scale in Korea

    Boere, G. C., Galbraith, C. A., Stroud, D. & Thompson, D. B. A. The conservation of waterbirds around the world in Waterbirds Around the World (ed. Boere, G. C., Galbraith, C. A. & Stroud, D. A.) 32–45 (The Stationery Office, Edinburgh, UK, 2007).Butchart, S. H. M. et al. Global biodiversity: Indicators of recent declines. Science 328, 1164–1168. https://doi.org/10.1126/science.1187512,Pubmed:20430971 (2010).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    Nebel, S., Porter, J. L. & Kingsford, R. T. Long-term trends of shorebird populations in eastern Australia and impacts of freshwater extraction. Biol. Conserv. 141, 971–980. https://doi.org/10.1016/j.biocon.2008.01.017 (2008).Article 

    Google Scholar 
    MacKinnon, J., Verkuil, Y. I. & Murray, N. IUCN situation analysis on east and Southeast Asian intertidal habitats, with particular reference to the Yellow Sea (including the Bohai Sea). Occas. Pap. IUCN Species Surviv. Comm. 047, 70.Li, X. et al. Assessing changes of habitat quality for shorebirds in stopover sites: A case study in Yellow River Delta, China. Wetlands 39, 67–77. https://doi.org/10.1007/s13157-018-1075-9 (2019).Article 

    Google Scholar 
    Murray, N. J., Clemens, R. S., Phinn, S. R., Possingham, H. P. & Fuller, R. A. Tracking the rapid loss of tidal wetlands in the Yellow Sea. Front. Ecol. Environ. 12, 267–272. https://doi.org/10.1890/130260 (2014).Article 

    Google Scholar 
    Green, J. M. H., Sripanomyom, S., Giam, X. & Wilcove, D. S. The ecology and economics of shorebird conservation in a tropical human-modified landscape. J. Appl. Ecol. 52, 1483–1491. https://doi.org/10.1111/1365-2664.12508 (2015).Article 

    Google Scholar 
    Toral, G. M. & Figuerola, J. Unraveling the importance of rice fields for waterbird populations in Europe. Biodivers. Conserv. 19, 3459–3469. https://doi.org/10.1007/s10531-010-9907-9 (2010).Article 

    Google Scholar 
    Masero, J. A. Assessing alternative anthropogenic habitats for conserving waterbirds: salinas as buffer areas against the impact of natural habitat loss for shorebirds. Biodivers. Conserv. 12, 1157–1173. https://doi.org/10.1023/A:1023021320448 (2003).Article 

    Google Scholar 
    Athearn, N. D. et al. Variability in habitat value of commercial salt production ponds: implications for waterbird management and tidal marsh restoration planning. Hydrobiologia 697, 139–155. https://doi.org/10.1007/s10750-012-1177-y (2012).CAS 
    Article 

    Google Scholar 
    Navedo, J. G. et al. Agroecosystems and conservation of migratory waterbirds: Importance of coastal pastures and factors influencing their use by wintering shorebirds. Biodivers. Conserv. 22, 1895–1907. https://doi.org/10.1007/s10531-013-0516-2 (2013).Article 

    Google Scholar 
    Navedo, J. G., Fernández, G., Valdivia, N., Drever, M. C. & Masero, J. A. Identifying management actions to increase foraging opportunities for shorebirds at semi-intensive shrimp farms. J. Appl. Ecol. 54, 567–576. https://doi.org/10.1111/1365-2664.12735 (2017).Article 

    Google Scholar 
    Lawler, S. P. Rice fields as temporary wetlands: A review. Isr. J. Zool. 47, 513–528. https://doi.org/10.1560/X7K3-9JG8-MH2J-XGX1 (2001).Article 

    Google Scholar 
    Fasola, M. & Ruiz, X. The value of rice fields as substitutes for natural wetlands for waterbirds in the Mediterranean region. Waterbirds 19, 122–128. https://doi.org/10.2307/1521955 (1996).Article 

    Google Scholar 
    Elphick, C. S. Why study birds in rice fields?. Waterbirds 33, 1–7. https://doi.org/10.1675/063.033.s101 (2010).Article 

    Google Scholar 
    Lourenço, P. M. & Piersma, T. Stopover ecology of Black-tailed Godwits Limosa limosa limosa in Portuguese rice fields: A guide on where to feed in winter. Bird Study 55, 194–202. https://doi.org/10.1080/00063650809461522 (2008).Article 

    Google Scholar 
    Elphick, C. S. & Oring, L. W. Conservation implications of flooding rice fields on winter waterbird communities. Agric. Ecosyst. Environ. 94, 17–29. https://doi.org/10.1016/S0167-8809(02)00022-1 (2003).Article 

    Google Scholar 
    Maeda, T. Patterns of bird abundance and habitat use in rice fields of the Kanto Plain, central Japan. Ecol. Res. 16, 569–585. https://doi.org/10.1046/j.1440-1703.2001.00418.x (2001).Article 

    Google Scholar 
    Choi, S. H., Nam, H. K. & Yoo, J. C. Characteristics of population dynamics and habitat use of shorebirds in rice fields during spring migration. Korean J. Environ. Agric. 33, 334–343. https://doi.org/10.5338/KJEA.2014.33.4.334 (2014).Article 

    Google Scholar 
    Shuford, W. D., Humphrey, J. M. & Nur, N. Breeding status of the Black tern in California. West. Birds 32, 189–217 (2001).
    Google Scholar 
    Sánchez-Guzmán, J. M. et al. Identifying new buffer areas for conserving waterbirds in the Mediterranean basin: The importance of the rice fields in Extremadura, Spain. Biodivers. Conserv. 16, 3333–3344. https://doi.org/10.1007/s10531-006-9018-9 (2007).Article 

    Google Scholar 
    Rendón, M. A., Green, A. J., Aguilera, E. & Almaraz, P. Status, distribution and long-term changes in the waterbird community wintering in Doñana, south–west Spain. Biol. Conserv. 141, 1371–1388. https://doi.org/10.1016/j.biocon.2008.03.006 (2008).Article 

    Google Scholar 
    Ibáñez, C., Curcó, A., Riera, X., Ripoll, I. & Sánchez, C. Influence on birds of rice field management practices during the growing season: A review and an experiment. Waterbirds 33, 167–180. https://doi.org/10.1675/063.033.s113 (2010).Article 

    Google Scholar 
    Pierluissi, S. Breeding waterbirds in rice fields: A global review. Waterbirds 33, 123–132. https://doi.org/10.1675/063.033.0117 (2010).Article 

    Google Scholar 
    Day, J. H. & Colwell, M. A. Waterbird communities in rice fields subjected to different post-harvest treatments. Waterbirds 21, 185–197. https://doi.org/10.2307/1521905 (1998).Article 

    Google Scholar 
    Elphick, C. S. & Oring, L. W. Winter management of Californian rice fields for waterbirds. J. Appl. Ecol. 35, 95–108. https://doi.org/10.1046/j.1365-2664.1998.00274.x (1998).Article 

    Google Scholar 
    Manley, S. W., Kaminski, R. M., Reinecke, K. J. & Gerard, P. D. Waterbird foods in winter-managed ricefields in Mississippi. J. Wildl. Manag. 68, 74–83. https://doi.org/10.2193/0022-541X(2004)068[0074:WFIWRI]2.0.CO;2 (2004).Article 

    Google Scholar 
    Pernollet, C. A., Cavallo, F., Simpson, D., Gauthier-Clerc, M. & Guillemain, M. Seed density and waterfowl use of rice fields in Camargue, France. Jour. Wild. Mgmt 81, 96–111. https://doi.org/10.1002/jwmg.21167 (2017).Article 

    Google Scholar 
    Nam, H. K., Choi, S. H. & Yoo, J. C. Influence of foraging behaviors of shorebirds on habitat use in rice fields during spring migration. Korean J. Environ. Agric. 34, 178–185. https://doi.org/10.5338/KJEA.2015.34.3.35 (2015).Article 

    Google Scholar 
    Nam, H. K., Choi, S. H., Choi, Y. S. & Yoo, J. C. Patterns of waterbirds abundance and habitat use in rice fields. Korean J. Environ. Agr. 31, 359–367. https://doi.org/10.5338/KJEA.2012.31.4.359 (2012).Article 

    Google Scholar 
    Nam, H. K., Choi, Y. S., Choi, S. H. & Yoo, J. C. Distribution of waterbirds in rice fields and their use of foraging habitats. Waterbirds 38, 173–183. https://doi.org/10.1675/063.038.0206 (2015).Article 

    Google Scholar 
    Hua, N., Tan, K., Chen, Y. & Ma, Z. Key research issues concerning the conservation of migratory shorebirds in the Yellow Sea region. Bird Conserv. Int. 25, 38–52. https://doi.org/10.1017/S0959270914000380 (2015).Article 

    Google Scholar 
    Stroud, D. A. et al. The conservation and population status of the world’s waders at the turn of the millennium in. Waterbirds Around the World Conference 643–648 (The Stationery Office, Edinburgh, UK, 2006).Taft, O. W. & Haig, S. M. The value of agricultural wetlands as invertebrate resources for wintering shorebirds. Agric. Ecosyst. Environ. 110, 249–256. https://doi.org/10.1016/j.agee.2005.04.012 (2005).Article 

    Google Scholar 
    Strum, K. M. et al. Winter management of California’s rice fields to maximize waterbird habitat and minimize water use. Agric. Ecosyst. Environ. 179, 116–124. https://doi.org/10.1016/j.agee.2013.08.003 (2013).Article 

    Google Scholar 
    Dias, R. A., Blanco, D. E., Goijman, A. P. & Zaccagnini, M. E. Density, habitat use, and opportunities for conservation of shorebirds in rice fields in southeastern South America. Condor Ornithol. Appl. 116, 384–393. https://doi.org/10.1650/CONDOR-13-160.1 (2014).Article 

    Google Scholar 
    Golet, G. H. et al. Using ricelands to provide temporary shorebird habitat during migration. Ecol. Appl. 28, 409–426. https://doi.org/10.1002/eap.1658,Pubmed:29205645 (2018).Article 
    PubMed 

    Google Scholar 
    Choi, G., Nam, H. K., Son, S. J., Do, M. S. & Yoo, J. C. The impact of agricultural activities on habitat use by the Wood sandpiper and Common greenshank in rice fields. Ornithol. Sci. 20, 27–37. https://doi.org/10.2326/osj.20.27 (2021).Article 

    Google Scholar 
    Choi, S. H. & Nam, H. K. Flexible behavior of the Black-tailed godwit Limosa limosa is key to successful refueling during staging at rice paddy fields in Midwestern Korea. Zool. Sci. 37, 255–262. https://doi.org/10.2108/zs190120,Pubmed:32549539 (2020).Article 

    Google Scholar 
    Cole, M. L., Leslie, D. M. Jr. & Fisher, W. L. Habitat use by shorebirds at a stopover site in the southern Great Plains. Southwest. Nat. 47, 372–378. https://doi.org/10.2307/3672495 (2002).Article 

    Google Scholar 
    Rundle, W. D. & Fredrickson, L. H. Managing seasonally flooded impoundments for migrant rails and shorebirds. Wildl. Soc. Bull., 80–87 (1981).Taylor, D. M. & Trost, C. H. Use of lakes and reservoirs by migrating shorebirds in Idaho. Gr Basin Nat. 52, 179–184 (1992).
    Google Scholar 
    Hands, H. M., Ryan, M. R. & Smith, J. W. Migrant shorebird use of marsh, Moist-Soil, and flooded agricultural habitats. Wildl. Soc. Bull. 1973–2006(19), 457–464 (1991).
    Google Scholar 
    Skagen, S. K. & Knopf, F. L. Migrating shorebirds and habitat dynamics at a prairie wetland complex. Wilson Bull., 91–105 (1994).Weber, L. M. & Haig, S. M. Shorebird use of South Carolina managed and natural coastal wetlands. J. Wildl. Manag. 60, 73–82. https://doi.org/10.2307/3802042 (1996).Article 

    Google Scholar 
    Collazo, J. A., O’Harra, D. A. & Kelly, C. A. Accessible habitat for shorebirds: factors influencing its availability and conservation implications. Waterbirds, 13–24 (2002).Helmers, D.L. Shorebird Management Manual. Western Hemisphere Shorebird Reserve Network (Manomet Cntr for Conservation Sciences, Manomet, MA, 1992).Verkuil, Y., Koolhaas, A. & Van Der Winden, J. Wind effects on prey availability: how northward migrating waders use brackish and hypersaline lagoons in the Sivash, Ukraine. Neth. J. Sea Res. 31, 359–374. https://doi.org/10.1016/0077-7579(93)90053-U (1993).Article 

    Google Scholar 
    Davis, C. A. & Smith, L. M. Ecology and management of migrant shorebirds in the Playa Lakes Region of Texas. Wildl. Monogr., 3–45 (1998).Barter, M. Shorebirds of the Yellow Sea: Importance, threats and conservation status in Wetlands Int. Global Ser. Oceania 9, 5–13.Katayama, N., Baba, Y. G., Kusumoto, Y. & Tanaka, K. A review of post-war changes in rice farming and biodiversity in Japan. Agric. Syst. 132, 73–84. https://doi.org/10.1016/j.agsy.2014.09.001 (2015).Article 

    Google Scholar 
    Mukherjee, A. Adaptiveness of cattle egret’s (Bubulcus ibis) foraging. Zoos Print J. 15, 331–333 (2000). https://doi.org/10.11609/JoTT.ZPJ.15.10.331-3.Choi, Y. S., Kim, S. S. & Yoo, J. C. Feeding activity of cattle egrets and intermediate egrets at different stages of rice culture in Korea. J. Ecol. Environ. 33, 149–155. https://doi.org/10.5141/JEFB.2010.33.2.149 (2010).Article 

    Google Scholar 
    Katayama, N., Amano, T., Fujita, G. & Higuchi, H. Spatial overlap between the intermediate egret Egretta intermedia and its aquatic prey at two spatiotemporal scales in a rice paddy landscape. Zool. Stud. 51, 1105–1112 (2013).
    Google Scholar 
    Sebastián-González, E. & Green, A. J. Reduction of avian diversity in created versus natural and restored wetlands. Ecography 39, 1176–1184. https://doi.org/10.1111/ecog.01736 (2016).Article 

    Google Scholar 
    Walton, M. E. M. et al. A model for the future: ecosystem services provided by the aquaculture activities of Veta La Palma, Southern Spain. Aquaculture 448, 382–390. https://doi.org/10.1016/j.aquaculture.2015.06.017 (2015).Article 

    Google Scholar 
    R Development Core Team. R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, Vienna, Austria, 2014).Bates, D., Maechler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1–48 (2015).Article 

    Google Scholar 
    Lüdecke, D. sjPlot: data visualization for statistics in social science. R package version 1.6.9. org/Package=sjPlot > (2016). http://CRAN.Rproject. More

  • in

    Myzorhynchus series of Anopheles mosquitoes as potential vectors of Plasmodium bubalis in Thailand

    Templeton, T. J., Martinsen, E., Kaewthamasorn, M. & Kaneko, O. The rediscovery of malaria parasites of ungulates. Parasitology 143, 1501–1508. https://doi.org/10.1017/S0031182016001141 (2016).Article 
    PubMed 

    Google Scholar 
    Sheather, A. A malaria parasite in the blood of a buffalo. J. Comp. Pathol. Ther. 32, 80026–80027 (1919).Article 

    Google Scholar 
    Templeton, T. J. et al. Ungulate malaria parasites. Sci. Rep. 6, 23230. https://doi.org/10.1038/srep23230 (2016).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kandel, R. C. et al. First report of malaria parasites in water buffalo in Nepal. Vet. Parasitol. Reg. Stud. Rep. 18, 100348. https://doi.org/10.1016/j.vprsr.2019.100348 (2019).Article 

    Google Scholar 
    Garnham, P. & Edeson, J. Two new malaria parasites of the Malayan mousedeer. Riv. Malariol. 41, 1–8 (1962).CAS 
    PubMed 

    Google Scholar 
    Hoo, C. & Sandosham, A. The early forms of Hepatocystis fieldi and Plasmodium traguli in the Malayan mouse-deer Tragulus javanicus. Med. J. Malays. 22, 299–301 (1968).
    Google Scholar 
    de Mello, F.d., Paes, S. Sur une plasmodiae du sang des chèvres. Cr. Séanc. Soc. Biol. 88, 829–830 (1923).Kaewthamasorn, M. et al. Genetic homogeneity of goat malaria parasites in Asia and Africa suggests their expansion with domestic goat host. Sci. Rep. 8, 5827. https://doi.org/10.1038/s41598-018-24048-0 (2018).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Rattanarithikul, R. H., Bruce, A., Harbach, R. E., Panthusiri, P. & Coleman, R. E. Illustrated keys to the mosquitoes of Thailand IV. Anopheles. Southeast Asian J Trop Med Public Health. 37 (2006).Walter Reed Biosystematics Unit. 2021. Systematic catalogue of Culicidae. http://mosquitocatalog.org. Last accessed on 20/09/2021.Sinka, M. E. et al. The dominant Anopheles vectors of human malaria in the Asia-Pacific region: occurrence data, distribution maps and bionomic précis. Parasit. Vectors. 4, 89. https://doi.org/10.1186/1756-3305-4-89 (2011).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Syafruddin, D. et al. Malaria prevalence in Nias District, North Sumatra Province, Indonesia. Malar. J. 6, 116. https://doi.org/10.1186/1475-2875-6-116 (2007).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Vantaux, A. et al. Anopheles ecology, genetics and malaria transmission in northern Cambodia. Sci. Rep. 11, 6458. https://doi.org/10.1038/s41598-021-85628-1 (2021).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Manguin, S., Garros, C., Dusfour, I., Harbach, R. & Coosemans, M. Bionomics, taxonomy, and distribution of the major malaria vector taxa of Anopheles subgenus Cellia in Southeast Asia: An updated review. Infect. Genet. Evol. 8, 489–503. https://doi.org/10.1016/j.meegid.2007.11.004 (2008).CAS 
    Article 
    PubMed 

    Google Scholar 
    Paredes-Esquivel, C., Donnelly, M. J., Harbach, R. E. & Townson, H. A molecular phylogeny of mosquitoes in the Anopheles barbirostris Subgroup reveals cryptic species: implications for identification of disease vectors. Mol. Phylogenet. Evol. 50, 141–151. https://doi.org/10.1016/j.ympev.2008.10.011 (2009).CAS 
    Article 
    PubMed 

    Google Scholar 
    Sungvornyothin, S., Garros, C., Chareonviriyaphap, T. & Manguin, S. How reliable is the humeral pale spot for identification of cryptic species of the Minimus Complex?. J. Am. Mosq. Control. Assoc. 22, 185–191. https://doi.org/10.2987/8756-971X(2006)22[185:HRITHP]2.0.CO;2 (2006).Article 
    PubMed 

    Google Scholar 
    Brosseau, L. et al. A multiplex PCR assay for the identification of five species of the Anopheles barbirostris complex in Thailand. Parasite. Vectors. 12, 223. https://doi.org/10.1186/s13071-019-3494-8 (2019).Article 

    Google Scholar 
    Taai, K. & Harbach, R. E. Systematics of the Anopheles barbirostris species complex (Diptera: Culicidae: Anophelinae) in Thailand. Zool. J. Linn. Soc. 174, 244–264. https://doi.org/10.1111/zoj.12236 (2015).Article 

    Google Scholar 
    Dahan-Moss, Y. et al. Member species of the Anopheles gambiae complex can be misidentified as Anopheles leesoni. Malar. J. 19, 1–9. https://doi.org/10.1186/s12936-020-03168-x (2020).CAS 
    Article 

    Google Scholar 
    De Ang, J. X., Yaman, K., Kadir, K. A., Matusop, A. & Singh, B. New vectors that are early feeders for Plasmodium knowlesi and other simian malaria parasites in Sarawak, Malaysian Borneo. Sci. Rep. https://doi.org/10.1038/s41598-021-86107-3 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Van Bortel, W. et al. Confirmation of Anopheles varuna in Vietnam, previously misidentified and mistargeted as the malaria vector Anopheles minimus. Am. J. Trop. Med. Hyg. 65, 729–732. https://doi.org/10.4269/ajtmh.2001.65.729 (2001).Article 
    PubMed 

    Google Scholar 
    Wharton, R., Eyles, D. E., Warren, M., Moorhouse, D. & Sandosham, A. Investigations leading to the identification of members of the Anopheles umbrosus group as the probable vectors of mouse deer malaria. Bull. World. Health. Organ. 29, 357 (1963).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Boundenga, L. et al. Haemosporidian parasites of antelopes and other vertebrates from Gabon, Central Africa. PLoS ONE 11, e0148958. https://doi.org/10.1371/journal.pone.0148958 (2016).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Martinsen, E. S. et al. Hidden in plain sight: Cryptic and endemic malaria parasites in North American white-tailed deer (Odocoileus virginianus). Sci. Adv. 2, e1501486. https://doi.org/10.1126/sciadv.1501486 (2016).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Garnham, P. C. C. Malaria Parasites and Other Haemosporidia (Blackwell Sci. Pub, 1966).
    Google Scholar 
    Nguyen, A. H. L., Tiawsirisup, S. & Kaewthamasorn, M. Low level of genetic diversity and high occurrence of vector-borne protozoa in water buffaloes in Thailand based on 18S ribosomal RNA and mitochondrial cytochrome b genes. Infect. Genet. Evol. 82, 104304. https://doi.org/10.1016/j.meegid.2020.104304 (2020).CAS 
    Article 
    PubMed 

    Google Scholar 
    Hebert, P. D., Cywinska, A. & Ball, S. L. Biological identifications through DNA barcodes. Proc. R. Soc. Lond. B Biol. Sci. 270, 313–321. https://doi.org/10.1098/rspb.2002.2218 (2003).CAS 
    Article 

    Google Scholar 
    Cywinska, A., Hunter, F. & Hebert, P. D. Identifying Canadian mosquito species through DNA barcodes. Med. Vet. Entomol. 20, 413–424. https://doi.org/10.1111/j.1365-2915.2006.00653.x (2006).CAS 
    Article 
    PubMed 

    Google Scholar 
    Ogola, E. O., Chepkorir, E., Sang, R. & Tchouassi, D. P. A previously unreported potential malaria vector in a dry ecology of Kenya. Parasites Vectors 12, 80. https://doi.org/10.1186/s13071-019-3332-z (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Maquart, P.-O., Fontenille, D., Rahola, N., Yean, S. & Boyer, S. Checklist of the mosquito fauna (Diptera, Culicidae) of Cambodia. Parasite 28, 60. https://doi.org/10.1051/parasite/2021056 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Saeung, A. et al. Geographic distribution and genetic compatibility among six karyotypic forms of Anopheles peditaeniatus (Diptera: Culicidae) in Thailand. Trop. Biomed. 29, 613–625 (2012).CAS 
    PubMed 

    Google Scholar 
    Tainchum, K., Kongmee, M., Manguin, S., Bangs, M. J. & Chareonviriyaphap, T. Anopheles species diversity and distribution of the malaria vectors of Thailand. Trends Parasitol. 31, 109–119. https://doi.org/10.1016/j.pt.2015.01.004 (2015).Article 
    PubMed 

    Google Scholar 
    Chookaew, S. et al. Anopheles species composition in malaria high-risk areas in Ranong Province. Dis. Cont. J. 46, 483–493. https://doi.org/10.14456/dcj.2020.45 (2020).Article 

    Google Scholar 
    Reid, J. A. The Anopheles barbirostris group (Diptera, Culicidae). Bull. Entomol. Res. 53, 1–57 (1962).Article 

    Google Scholar 
    Harrison, B. A. & Scanlon, J. E. Medical entomology studies–II. The subgenus Anopheles in Thailand (Diptera: Culicidae). Contributions of the American Entomological Institute (Ann Arbor) 12 (1): iv + 1–iv 307 (1975).Wang, Y., Xu, J. & Ma, Y. Molecular characterization of cryptic species of Anopheles barbirostris van der Wulp in China. Parasite Vectors 7, 592. https://doi.org/10.1186/s13071-014-0592-5 (2014).CAS 
    Article 

    Google Scholar 
    Wang, G. et al. An evaluation of the suitability of COI and COII gene variation for reconstructing the phylogeny of, and identifying cryptic species in, anopheline mosquitoes (Diptera Culicidae). Mitochondrial DNA Part A. 28, 769–777. https://doi.org/10.1080/24701394.2016.1186665 (2017).CAS 
    Article 

    Google Scholar 
    Davidson, J. R. et al. Molecular analysis reveals a high diversity of Anopheles species in Karama, West Sulawesi, Indonesia. Parasite Vectors https://doi.org/10.1186/s13071-020-04252-6 (2020).Article 

    Google Scholar 
    Beebe, N. W. DNA barcoding mosquitoes: advice for potential prospectors. Parasitology 145(5), 622–633. https://doi.org/10.1017/S0031182018000343 (2018).CAS 
    Article 
    PubMed 

    Google Scholar 
    Gunathilaka, N. Illustrated key to the adult female Anopheles (Diptera: Culicidae) mosquitoes of Sri Lanka. Appl. Entomol. Zool. 52, 69–77. https://doi.org/10.1007/s13355-016-0455-y (2017).Article 
    PubMed 

    Google Scholar 
    WHO. Pictorial identification key of important disease vectors in the WHO South-East Asia Region. https://apps.who.int/iris/handle/10665/332202 (accessed 20 August 2021).Rigg, C. A., Hurtado, L. A., Calzada, J. E. & Chaves, L. F. Malaria infection rates in Anopheles albimanus (Diptera: Culicidae) at Ipetí-Guna, a village within a region targeted for malaria elimination in Panamá. Infect. Genet. Evol. 69, 216–223. https://doi.org/10.1016/j.meegid.2019.02.003 (2019).Article 
    PubMed 

    Google Scholar 
    Torres-Cosme, R. et al. Natural malaria infection in anophelines vectors and their incrimination in local malaria transmission in Darién, Panama. PLoS ONE 16, e0250059. https://doi.org/10.1371/journal.pone.0250059 (2021).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Beebe, N. W. & Saul, A. Discrimination of all Members of the Anopheles punctulatus complex by polymerase chain reaction-restriction fragment length polymorphism analysis. Am. J. Trop. Med. Hyg. 53, 478–481. https://doi.org/10.4269/ajtmh.1995.53.478 (1995).CAS 
    Article 
    PubMed 

    Google Scholar 
    Perkins, S. L. & Schall, J. J. A molecular phylogeny of malaria parasites recovered from cytochrome b gene sequences. J. Parasitol. 88, 972–978. https://doi.org/10.1645/0022-3395(2002)088[0972:AMPOMP]2.0.CO;2 (2002).CAS 
    Article 
    PubMed 

    Google Scholar 
    Snounou, G. et al. High sensitivity of detection of human malaria parasites by the use of nested polymerase chain reaction. Mol. Biochem. Parasitol. 61, 315–320. https://doi.org/10.1016/0166-6851(93)90077-b (1993).CAS 
    Article 
    PubMed 

    Google Scholar 
    Hall, T. A. BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic. Acids. Symp. Ser. 41, 95–98 (1999).CAS 

    Google Scholar 
    Schoener, E. et al. Avian Plasmodium in Eastern Austrian mosquitoes. Malar. J. 16, 389. https://doi.org/10.1186/s12936-017-2035-1 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Ventim, R. et al. Avian malaria infections in western European mosquitoes. Parasitol. Res. 111, 637–645. https://doi.org/10.1007/s00436-012-2880-3 (2012).Article 
    PubMed 

    Google Scholar 
    Huelsenbeck, J. P. & Ronquist, F. MRBAYES: Bayesian inference of phylogeny. Bioinformatics 17, 754–755 (2001).CAS 
    Article 

    Google Scholar 
    Rambaut, A., Drummond, A. J., Xie, D., Baele, G. & Suchard, M. A. Posterior summarization in Bayesian phylogenetics using tracer 1.7. Syst. Biol. 67, 901–904 (2018).CAS 
    Article 

    Google Scholar  More

  • in

    Spatial cover and carbon fluxes of urbanized Sonoran Desert biological soil crusts

    Bethany, J., Giraldo-Silva, A., Nelson, C., Barger, N. N. & Garcia-Pichel, F. Optimizing the production of nursery-based biological soil crusts for restoration of arid land soils. Appl. Environ. Microbiol. 85(15), e00735-e819 (2019).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Belnap, J. & Gardner, J. S. Soil microstructure in soils of the colorado plateau—The role of the cyanobacterium Microcoleus-vaginatus. Gt. Basin Nat. 53(1), 40–47 (1993).
    Google Scholar 
    Belnap, J. Factors influencing nitrogen fixation and nitrogen release in biological soil crusts. Ecol. Stud. Biol. Soil Crusts Struct. Funct. Manag. 150, 241–261 (2001).
    Google Scholar 
    Cameron, R. E. & Blank, G. B. Desert algae: Soil crusts and diaphanous substrata as algal habitats. Tech. Rep. Jet Propul. Lab. Calif. Technol. 32–971, 1–41 (1966).
    Google Scholar 
    Friedmann EI, Galun M. Desert algae lichens and fungi. in Desert Biology (Brown Jr, G.W. eds). Vol. 2. 165–212. (Illus Academic Press, Inc., 1974).Maier, S., Tamm, A., Wu, D.A.-O., Caesar, J., Grube, M., & Weber, B.A.-O. Photoautotrophic Organisms Control Microbial Abundance, Diversity, and Physiology in Different Types of Biological Soil Crusts. (1751–7370 (electronic)).Cable, J. M. & Huxman, T. E. Precipitation pulse size effects on Sonoran Desert soil microbial crusts. Oecologia 141(2), 317–324 (2004).ADS 
    PubMed 

    Google Scholar 
    Evans, R. D. & Johansen, J. R. Microbiotic crusts and ecosystem processes. Crit. Rev. Plant Sci. 18(2), 183–225 (1999).
    Google Scholar 
    Thompson, J. N. et al. Frontiers of ecology. Bioscience 51(1), 15–24 (2001).
    Google Scholar 
    Warren, S. D., Rosentreter, R. & Pietrasiak, N. Biological soil crusts of the Great Plains: A review. Rangel Ecol. Manag. 1(78), 213–219 (2021).
    Google Scholar 
    Warren, S. D. et al. Biological soil crust response to late season prescribed fire in a Great Basin Juniper Woodland. Rangel. Ecol. Manag. 68(3), 241–247 (2015).
    Google Scholar 
    Thomas, A. D., Hoon, S. R. & Linton, P. E. Carbon dioxide fluxes from cyanobacteria crusted soils in the Kalahari. Appl. Soil Ecol. 39, 254–263 (2008).
    Google Scholar 
    Williams, A. J., Buck, B. J. & Beyene, M. A. Biological soil crusts in the Mojave Desert, USA: Micromorphology and pedogenesis. Soil Sci. Soc. Am. J. 76(5), 1685–1695 (2012).ADS 
    CAS 

    Google Scholar 
    Belnap, J. & Lange, O. L. Ecological studies: Biological soil crusts: Structure, function, and management. Ecol. Stud. Biol. Soil Crusts Struct. Funct. Manag. 150, 1–503 (2001).
    Google Scholar 
    Jordan, W. R. I. Restoration ecology: A synthetic approach to ecological research. Rehabil. Damaged Ecosyst. 2, 373–384 (1995).
    Google Scholar 
    Nash, T. H. et al. Photosynthetic patterns of Sonoran desert lichens.1. Environmental considerations and preliminary field-measurements. Flora 172(4), 335–345 (1982).
    Google Scholar 
    St. Clair, L. L., Johansen, J. R. & Rushforth, S. R. Lichens of soil crust communities in the Intermountain Area of the western United States. Gt Basin Nat. 53(1), 5 (1993).
    Google Scholar 
    Bowker, M. A., Belnap, J. & Miller, M. E. Spatial modeling of biological soil crusts to support rangeland assessment and monitoring. Rangel. Ecol. Manag. 59(5), 519–529 (2006).
    Google Scholar 
    Mayland, H. F., McIntosh, T. H. & Fuller, W. H. Fixation of isotopic nitrogen on a semiarid soil by algal crust organisms. Soil Sci. Soc. Am. Proc. 30(1), 56 (1966).ADS 
    CAS 

    Google Scholar 
    McIlvanie, S. K. Grass seedling establishment, and productivity—Overgrazed vs. protected range soils. Ecology 23(2), 228–231 (1942).
    Google Scholar 
    Webb, R. H. & Wilshire, H. G. Environmental Effects of Off-Road Vehicles : Impacts and Management in Arid Regions (Springer, 1983).
    Google Scholar 
    Zobel, D. & Antos, J. A decade of recovery of understory vegetation buried by volcanic tephra from Mount St. Helens. Ecol. Monogr. 1, 67 (1997).
    Google Scholar 
    Condon, L. & Pyke, D. Resiliency of biological soil crusts and vascular plants varies among morphogroups with disturbance intensity. Plant Soil. 12, 433 (2020).
    Google Scholar 
    Harper, K., & Marble, J. A role for nonvascular plants in management of arid and semiarid rangelands. in Vegetation Science Applications for Rangeland Analysis and Management [Internet] (Tueller, P.T., ed.). Handbook of Vegetation Science. Vol. 14. 135–169. https://doi.org/10.1007/978-94-009-3085-8_7. (Springer, 1988). Evans, R. D. & Belnap, J. Long-term consequences of disturbance on nitrogen dynamics in an arid ecosystem. Ecology 80(1), 150–160 (1999).
    Google Scholar 
    Sheridan, R. P. Impact of emissions from coal-fired electricity generating facilities on N2-fixing lichens. Bryologist 82(1), 54–58 (1979).CAS 

    Google Scholar 
    Henriksson, L. E. & Dasilva, E. J. Effects of some inorganic elements on nitrogen-fixation in blue-green-algae and some ecological aspects of pollution. Z. Allg. Mikrobiol. 18(7), 487–494 (1978).CAS 
    PubMed 

    Google Scholar 
    Freebury, C. Lichens and lichenicolous fungi of Grasslands National Park (Saskatchewan, Canada). Opusc Philolichenum 13, 102–121 (2009).
    Google Scholar 
    Szyja, M. et al. Neglected but potent dry forest players: ecological role and ecosystem service provision of biological soil crusts in the human-modified Caatinga. Front. Ecol. Evol. (Internet). https://doi.org/10.3389/fevo.2019.00482 (2019).Article 

    Google Scholar 
    Rosentreter, R. Biological soil of crusts of North American drylands: Cryptic diversity at risk. in Reference Module in Earth Systems and Environmental Sciences [Internet]. https://www.sciencedirect.com/science/article/pii/B9780128211397000738 (Elsevier, 2021). Kranz, C. N., McLaughlin, R. A., Johnson, A., Miller, G. & Heitman, J. L. The effects of compost incorporation on soil physical properties in urban soils—A concise review. J. Environ. Manag. 261, 110209 (2020).
    Google Scholar 
    Barberán, A. et al. Continental-scale distributions of dust-associated bacteria and fungi. Proc. Natl. Acad. Sci. 112(18), 5756 (2015).ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kaye, J. P., Groffman, P. M., Grimm, N. B., Baker, L. A. & Pouyat, R. V. A distinct urban biogeochemistry?. Trends Ecol. Evol. 21(4), 192–199 (2006).PubMed 

    Google Scholar 
    Pavao-Zuckerman, M. A. The nature of urban soils and their role in ecological restoration in cities. Restor. Ecol. 16(4), 642–649 (2008).
    Google Scholar 
    Pouyat, R., Groffman, P., Yesilonis, I. & Hernandez, L. Soil carbon pools and fluxes in urban ecosystems. Environ. Pollut. 116, S107–S118 (2002).CAS 
    PubMed 

    Google Scholar 
    Behzad, H., Mineta, K., & Gojobori, T. Global Ramifications of Dust and Sandstorm Microbiota. (1759–6653 (electronic)).Warren, S., Clair, L. & Leavitt, S. Aerobiology and passive restoration of biological soil crusts. Aerobiologia 3, 35 (2021).
    Google Scholar 
    Hall, S. J. et al. Urbanization alters soil microbial functioning in the Sonoran Desert. Ecosystems 12(4), 654–671 (2009).CAS 

    Google Scholar 
    Ball, B. A. & Guevara, J. A. The nutrient plasticity of moss-dominated crust in the urbanized Sonoran Desert. Plant Soil. 389(1–2), 225–235 (2015).CAS 

    Google Scholar 
    Allen, C. D. Monitoring environmental impact in the Upper Sonoran lifestyle: A new tool for rapid ecological assessment. Environ. Manag. 43(2), 346–356 (2009).ADS 

    Google Scholar 
    Evans, R. A. & Love, R. M. The step-point method of sampling: A practical tool in range research. J. Range Manag. 10(5), 208–212 (1957).
    Google Scholar 
    Coulloudon, B., & National Applied Resource Sciences C. Sampling Vegetation Attributes Interagency Technical Reference [Internet]. http://www.blm.gov/nstc/library/pdf/samplveg.pdf. (Bureau of Land Management : National Business Center, 1999). Faithfull, N. T. Methods in agricultural chemical analysis: A practical handbook. Methods Agric. Chem. Anal. Pract. Handb. 1–22, 1–266 (2002).
    Google Scholar 
    Kuske, C. R., Yeager, C. M., Johnson, S., Ticknor, L. O. & Belnap, J. Response and resilience of soil biocrust bacterial communities to chronic physical disturbance in arid shrublands. ISME J. 6(4), 886–897 (2012).CAS 
    PubMed 

    Google Scholar 
    Lorenz, K. & Lal, R. Biogeochemical C and N cycles in urban soils. Environ. Int. 35(1), 1–8 (2009).CAS 
    PubMed 

    Google Scholar 
    Chamizo, S., Canton, Y., Lazaro, R., Sole-Benet, A. & Domingo, F. Crust composition and disturbance drive infiltration through biological soil crusts in semiarid ecosystems. Ecosystems 15(1), 148–161 (2012).
    Google Scholar 
    Kidron, G. J. & Gutschick, V. P. Soil moisture correlates with shrub-grass association in the Chihuahuan Desert. CATENA 107, 71–79 (2013).
    Google Scholar 
    Kidron, G. J., Monger, H. C., Vonshak, A. & Conrod, W. Contrasting effects of microbiotic crusts on runoff in desert surfaces. Geomorphology 15(139), 484–494 (2012).ADS 

    Google Scholar 
    Berdugo, M., Soliveres, S. & Maestre, F. T. Vascular plants and biocrusts modulate how abiotic factors affect wetting and drying events in drylands. Ecosystems 17, 1242 (2014).CAS 

    Google Scholar 
    Maestre, F. T. et al. Changes in biocrust cover drive carbon cycle responses to climate change in drylands. Glob Change Biol. 19, 3835 (2013).ADS 

    Google Scholar 
    Valenzuela, A. et al. Aerosol radiative forcing during African desert dust events (2005–2010) over southeastern Spain. Atmos. Chem. Phys. 12(21), 10331–10351 (2012).ADS 
    CAS 

    Google Scholar 
    Kaya, S., Basar, U. G., Karaca, M. & Seker, D. Z. Assessment of urban heat islands using remotely sensed data. Ekoloji 21(84), 107–113 (2012).
    Google Scholar 
    Demmigadams, B. et al. Effect of high light on the efficiency of photochemical energy-conversion in a variety of lichen species with green and blue-green phycobionts. Planta 180(3), 400–409 (1990).CAS 

    Google Scholar 
    Gauslaa, Y. & Rikkinen, J. What’s behind the pretty colours? A study on the photobiology of lichens. Nord. J. Bot. 17(5), 556–556 (1995).
    Google Scholar 
    Garciapichel, F. & Castenholz, R. W. Characterization and biological implications of scytonemin, a cyanobacterial sheath pigment. J. Phycol. 27(3), 395–409 (1991).CAS 

    Google Scholar 
    Garcia-Pichel, F. & Castenholz, R. W. The role of sheath pigments in the adaptation of terrestrial cyanobacteria to near UV radiation. J. Phycol. 27(3 SUPPL), 24–24 (1991).
    Google Scholar 
    McDonnell, M. J. et al. Ecosystem processes along an urban-to-rural gradient. Urban Ecosyst. 1(1), 21–36 (1997).
    Google Scholar 
    Pavao-Zuckerman, M. A. & Byrne, L. B. Scratching the surface and digging deeper: Exploring ecological theories in urban soils. Urban Ecosyst. 12(1), 9–20 (2009).
    Google Scholar 
    Pavao-Zuckerman, M. A. Urban greenscape, soils, and ecosystem functioning in a semi-arid urban ecosystem. J. Nematol. 41(4), 369–370 (2009).
    Google Scholar 
    Collins, S. L. et al. Pulse dynamics and microbial processes in aridland ecosystems. J. Ecol. 96(3), 413–420 (2008).
    Google Scholar 
    Noy-Meir, I. Desert ecosystems environment and producers. In Annual Review on Ecology System (Johnston Richard, F. ed.). Vol. 4. 25–51. (Illus Map Annu Rev Inc, 1973). More

  • in

    The great acceleration of plant phenological shifts

    Kaplan, J. O., Krumhardt, K. M. & Zimmermann, N. Quat. Sci. Rev. 28, 3016–3034 (2009).Article 

    Google Scholar 
    Zheng, Z. et al. Proc. Natl Acad. Sci. USA 118, e2022210118 (2021).CAS 
    Article 

    Google Scholar 
    Lewis, S. L. & Maslin, M. A. Nature 519, 171–180 (2015).CAS 
    Article 

    Google Scholar 
    Steffen, W. et al. Anthr. Rev. 2, 81–98 (2015).
    Google Scholar 
    Ripple, W. J. et al. BioScience 70, 8–12 (2020).
    Google Scholar 
    Cohen, J. M., Lajeunesse, M. J. & Rohr, J. R. Nat. Clim. Change 8, 224–228 (2018).Article 

    Google Scholar 
    Vitasse, Y. et al. Biol. Rev. 96, 1816–1835 (2021).Article 

    Google Scholar 
    Aono, Y. & Kazui, K. Int. J. Climatol. 28, 905–914 (2008).Article 

    Google Scholar 
    Sparks, T. H. & Carey, P. D. J. Ecol. 83, 321–329 (1995).Article 

    Google Scholar 
    Ge, Q. et al. J. Geophys. Res. Biogeosci. 119, 301–311 (2014).Article 

    Google Scholar 
    IPCC Climate Change 2021: The Physical Science Basis (eds Masson-Delmotte, V. et al.) (Cambridge Univ. Press, 2021).Post, E., Steinman, B. A. & Mann, M. E. Sci. Rep. 8, 3927 (2018).Article 

    Google Scholar 
    Primack, R. B. & Miller-Rushing, A. J. Bioscience 62, 170–181 (2012).Article 

    Google Scholar 
    Kharouba, H. M. et al. Proc. Natl Acad. Sci. USA 115, 5211–5216 (2018).CAS 
    Article 

    Google Scholar 
    Richardson, A. D. et al. Agric. For. Meteorol. 169, 156–173 (2013).Article 

    Google Scholar 
    Parker, D. E., Legg, T. P. & Folland, C. K. Int. J. Climatol. 12, 317–342 (1992).Article 

    Google Scholar  More