More stories

  • in

    Negative frequency-dependent selection and asymmetrical transformation stabilise multi-strain bacterial population structures

    1.
    Russell JB. Discussion on diphtheria. BMJ. 1891;2:631–40.
    Article  Google Scholar 
    2.
    Joseph FH. Notes on some pathogenic bacteria as found in the Transvaal, and the variations from their European prototype. Rep. S Afr Assoc Adv Sci. 1904;2:237–42.
    Google Scholar 

    3.
    Eyre JW, Washbourn JW. Varities and virulence of the pneumococcus. Lancet. 1899;153:19–22.
    Article  Google Scholar 

    4.
    Gladstone RA, Lo SW, Lees JA, Croucher NJ, van Tonder AJ, Corander J, et al. International genomic definition of pneumococcal lineages, to contextualise disease, antibiotic resistance and vaccine impact. EBioMedicine. 2019;43:338–46.
    PubMed  PubMed Central  Article  Google Scholar 

    5.
    Colijn C, Corander J, Croucher NJ. Designing ecologically optimized pneumococcal vaccines using population genomics. Nat Microbiol. 2020;5:473–85.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    6.
    Weinberger DM, Malley R, Lipsitch M. Serotype replacement in disease after pneumococcal vaccination. Lancet. 2011;378:1962–73.
    PubMed  PubMed Central  Article  Google Scholar 

    7.
    Lefevre JC, Faucon G, Sicard AM, Gasc AM. DNA fingerprinting of Streptococcus pneumoniae strains by pulsed-field gel electrophoresis. J Clin Microbiol. 1993;31:2724 LP–8.
    Article  Google Scholar 

    8.
    Selander RK, Levin BR. Genetic diversity and structure in Escherichia coli populations. Science. 1980;210:545–7.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    9.
    Smith JM, Smith NH, O’Rourke M, Spratt BG. How clonal are bacteria? Proc Natl Acad Sci USA. 1993;90:4384–8.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    10.
    Maiden MCJ, Bygraves JA, Feil E, Morelli G, Russell JE, Urwin R, et al. Multilocus sequence typing: a portable approach to the identification of clones within populations of pathogenic microorganisms. Proc Natl Acad Sci USA. 1998;95:3140–5.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    11.
    Smith JM, Feil EJ, Smith NH. Population structure and evolutionary dynamics of pathogenic bacteria. BioEssays. 2000;22:1115–22.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    12.
    Croucher NJ, Coupland PG, Stevenson AE, Callendrello A, Bentley SD, Hanage WP. Diversification of bacterial genome content through distinct mechanisms over different timescales. Nat Commun. 2014;5:5471.
    PubMed  PubMed Central  Article  Google Scholar 

    13.
    Lees JA, Harris SR, Tonkin-Hill G, Gladstone RA, Lo S, Weiser JN, et al. Fast and flexible bacterial genomic epidemiology with PopPUNK. Genome Res. 2019;29:304–16.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    14.
    Alikhan NF, Zhou Z, Sergeant MJ, Achtman M. A genomic overview of the population structure of Salmonella. PLoS Genet. 2018;14:e1007261.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    15.
    Hanage WP, Spratt BG, Turner KME, Fraser C. Modelling bacterial speciation. Philos Trans R Soc B Biol Sci. 2006;361:2039–44.
    Article  Google Scholar 

    16.
    Ford Doolittle W, Papke RT. Genomics and the bacterial species problem. Genome Biol. 2006;7:116.
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    17.
    Shapiro BJ, David LA, Friedman J, Alm EJ. Looking for Darwin’s footprints in the microbial world. Trends Microbiol. 2009;17:196–204.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    18.
    Abu-Raddad LJ, Ferguson NM. The impact of cross-immunity, mutation and stochastic extinction on pathogen diversity. Proc R Soc B Biol Sci. 2004;271:2431–8.
    Article  Google Scholar 

    19.
    Fraser C, Hanage WP, Spratt BG. Neutral microepidemic evolution of bacterial pathogens. Proc Natl Acad Sci USA. 2005;102:1968–73.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    20.
    Corander J, Fraser C, Gutmann MU, Arnold B, Hanage WP, Bentley SD, et al. Frequency-dependent selection in vaccine-associated pneumococcal population dynamics. Nat Ecol Evol. 2017;1:1950–60.
    PubMed  PubMed Central  Article  Google Scholar 

    21.
    Buckee CO, Jolley KA, Recker M, Penman B, Kriz P, Gupta S, et al. Role of selection in the emergence of lineages and the evolution of virulence in Neisseria meningitidis. Proc Natl Acad Sci USA. 2008;105:15082–7.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    22.
    Croucher NJ, Finkelstein JA, Pelton SI, Mitchell PK, Lee GM, Parkhill J, et al. Population genomics of post-vaccine changes in pneumococcal epidemiology. Nat Genet. 2013;45:656–63.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    23.
    Croucher NJ, Hanage WP, Harris SR, McGee L, van der Linden M, de Lencastre H, et al. Variable recombination dynamics during the emergence, transmission and ‘disarming’ of a multidrug-resistant pneumococcal clone. BMC Biol. 2014;12:49.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    24.
    Croucher NJ, Chewapreecha C, Hanage WP, Harris SR, McGee L, van der Linden M, et al. Evidence for soft selective sweeps in the evolution of pneumococcal multidrug resistance and vaccine escape. Genome Biol Evol. 2014;6:1589–602.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    25.
    Vos M. A species concept for bacteria based on adaptive divergence. Trends Microbiol. 2011;19:1–7.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    26.
    Whitaker RJ. Allopatric origins of microbial species. Philos Trans R Soc B Biol Sci. 2006;361:1975–84.
    Article  Google Scholar 

    27.
    Fraser C, Alm EJ, Polz MF, Spratt BG, Hanage WP. The bacterial species challenge: Making sense of genetic and ecological diversity. Science. 2009;323:741–6.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    28.
    Hanage WP, Fraser C, Spratt BG. Fuzzy species among recombinogenic bacteria. BMC Biol. 2005;3:6.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    29.
    Fraser C, Hanage WP, Spratt BG. Recombination and the nature of bacterial speciation. Science. 2007;315:476–80.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    30.
    Sheppard SK, McCarthy ND, Falush D, Maiden MCJ. Convergence of Campylobacter species: implications for bacterial evolution. Science. 2008;320:237–9.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    31.
    Laible G, Spratt BG, Hakenbeck R. Interspecies recombinational events during the evolution of altered PBP 2x genes in penicillin‐resistant clinical isolates of Streptococcus pneumoniae. Mol Microbiol. 1991;5:1993–2002.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    32.
    Dowson CG, Coffey TJ, Kell C, Whiley RA. Evolution of penicillin resistance in Streptococcus pneumoniae; the role of Streptococcus mitis in the formation of a low affinity PBP2B in S. pneumoniae. Mol Microbiol. 1993;9:635–43.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    33.
    Majewski J, Zawadzki P, Pickerill P, Cohan FM, Dowson CG. Barriers to genetic exchange between bacterial species: Streptococcus pneumoniae transformation. J Bacteriol. 2000;182:1016–23.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    34.
    Bobay L-M, Ochman H. Biological Species Are Universal across Life’s Domains. Genome Biol Evol. 2017;9:491–501.
    PubMed Central  Article  Google Scholar 

    35.
    Croucher NJ, Harris SR, Barquist L, Parkhill J, Bentley SD. A high-resolution view of genome-wide pneumococcal transformation. PLoS Pathog. 2012;8:e1002745.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    36.
    Cohan F. Bacterial species and speciation. Syst Biol. 2001;50:513–24.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    37.
    Nosil P, Funk DJ, Ortiz-Barrientos D. Divergent selection and heterogeneous genomic divergence. Mol Ecol. 2009;18:375–402.
    PubMed  Article  PubMed Central  Google Scholar 

    38.
    Cohan FM, Perry EB. A systematics for discovering the fundamental units of bacterial diversity. Curr Biol. 2007;17:R373–R386.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    39.
    Marttinen P, Hanage WP. Speciation trajectories in recombining bacterial species. PLoS Comput Biol. 2017;13:e1005640.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    40.
    Gjini E, Valente C, Sá-Leão R, Gomes MGM. How direct competition shapes coexistence and vaccine effects in multi-strain pathogen systems. J Theor Biol. 2016;388:50–60.
    PubMed  Article  PubMed Central  Google Scholar 

    41.
    Numminen E, Cheng L, Gyllenberg M, Corander J. Estimating the transmission dynamics of Streptococcus pneumoniae from strain prevalence data. Biometrics. 2013;69:748–57.
    PubMed  Article  PubMed Central  Google Scholar 

    42.
    Majewski J, Cohan FM. Adapt globally, act locally: the effect of selective sweeps on bacterial sequence diversity. Genetics. 1999;152:1459–74.
    CAS  PubMed  PubMed Central  Google Scholar 

    43.
    Wiedenbeck J, Cohan FM. Origins of bacterial diversity through horizontal genetic transfer and adaptation to new ecological niches. FEMS Microbiol Rev. 2011;35:957–76.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    44.
    Barraclough TG, Balbi KJ, Ellis RJ. Evolving concepts of bacterial species. Evol Biol. 2012;39:148–57.
    Article  Google Scholar 

    45.
    Chesson P. Mechanisms of maintenance of species diversity. Annu Rev Ecol Syst. 2000;31:343–66.
    Article  Google Scholar 

    46.
    Neher RA, Shraiman BI. Competition between recombination and epistasis can cause a transition from allele to genotype selection. Proc Natl Acad Sci USA. 2009;106:6866–71.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    47.
    Mostowy R, Croucher NJ, Andam CP, Corander J, Hanage WP, Marttinen P. Efficient inference of recent and ancestral recombination within bacterial populations. Mol Biol Evol. 2017;34:1167–82.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    48.
    Croucher NJ, Harris SR, Fraser C, Quail MA, Burton J, van der Linden M, et al. Rapid pneumococcal evolution in response to clinical interventions. Science. 2011;331:430–4.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    49.
    Romero P, Croucher NJ, Hiller NL, Hu FZ, Ehrlich GD, Bentley SD, et al. Comparative genomic analysis of ten Streptococcus pneumoniae temperate bacteriophages. J Bacteriol. 2009;191:4854–62.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    50.
    Ferguson NM, Galvanl AP, Bush RM. Ecological and immunological determinants of influenza evolution. Nature. 2003;422:428–33.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    51.
    Cobey S, Lipsitch M. Niche and neutral effects of acquired immunity permit coexistence of pneumococcal serotypes. Science. 2012;335:1376–80.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    52.
    Cobey S. Pathogen evolution and the immunological niche. Ann N Y Acad Sci. 2014;1320:1–15.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    53.
    Levin BR. Frequency-dependent selection in bacterial populations. Philos Trans R Soc Lond B Biol Sci. 1988;319:459–72.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    54.
    Gomes MGM, Medley GF, Nokes DJ. On the determinants of population structure in antigenically diverse pathogens. Proc R Soc Lond Ser B Biol Sci. 2002;269:227–33.
    Article  Google Scholar 

    55.
    Binsker U, Lees JA, Hammond AJ, Weiser JN. Immune exclusion by naturally acquired secretory IgA against pneumococcal pilus-1. J Clin Investig. 2020;130:927–41.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    56.
    Gupta S, Maiden MCJ, Feavers IM, Nee S, May RM, Anderson RM. The maintenance of strain structure in populations of recombining infectious agents. Nat Med. 1996;2:437–42.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    57.
    Gupta S, Ferguson N, Anderson R. Chaos, persistence, and evolution of strain structure in antigenically diverse infectious agents. Science. 1998;280:912–5.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    58.
    Croucher NJ, Campo JJ, Le TQ, Liang X, Bentley SD, Hanage WP, et al. Diverse evolutionary patterns of pneumococcal antigens identified by pangenome-wide immunological screening. Proc Natl Acad Sci USA. 2017;114:E357–66.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    59.
    Weinberger DM, Dagan R, Givon‐Lavi N, Regev‐Yochay G, Malley R, Lipsitch M. Epidemiologic evidence for serotype‐specific acquired immunity to pneumococcal carriage. J Infect Dis. 2008;197:1511–8.
    PubMed  Article  PubMed Central  Google Scholar 

    60.
    Malley R, Lipsitch M, Bogaert D, Thompson CM, Hermans P, Claire Watkins A, et al. Serum antipneumococcal antibodies and pneumococcal colonization in adults with chronic obstructive pulmonary disease. J Infect Dis. 2007;196:928–35.
    PubMed  Article  Google Scholar 

    61.
    Campo JJ, Le TQ, Pablo JV, Hung C, Teng AA, Tettelin H, et al. Panproteome-wide analysis of antibody responses to whole cell pneumococcal vaccination. Elife. 2018;7:e37015.
    PubMed  PubMed Central  Article  Google Scholar 

    62.
    Buckee CO, Koelle K, Mustard MJ, Gupta S. The effects of host contact network structure on pathogen diversity and strain structure. Proc Natl Acad Sci USA. 2004;101:10839 LP–44.
    Article  Google Scholar 

    63.
    Buckee CO, Recker M, Watkins ER, Gupta S. Role of stochastic processes in maintaining discrete strain structure in antigenically diverse pathogen populations. Proc Natl Acad Sci USA. 2011;108:15504 LP–9.
    Article  Google Scholar 

    64.
    Watkins ER, Penman BS, Lourenço J, Buckee CO, Martin C, Maiden MCJ, et al. Vaccination drives changes in metabolic and virulence profiles of Streptococcus pneumoniae. PLoS Pathog. 2015;11:e1005034.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    65.
    Skwark MJ, Croucher NJ, Puranen S, Chewapreecha C, Pesonen M, Xu YY, et al. Interacting networks of resistance, virulence and core machinery genes identified by genome-wide epistasis analysis. PLoS Genet. 2017;13:e1006508.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    66.
    Pensar J, Puranen S, Arnold B, MacAlasdair N, Kuronen J, Tonkin-Hill G, et al. Genome-wide epistasis and co-selection study using mutual information. Nucleic Acids Res. 2019;47:e112.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    67.
    Rocha EPC. Neutral theory, microbial practice: challenges in bacterial population genetics. Mol Biol Evol. 2018;35:1338–47.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    68.
    Azarian T, Martinez PPP, Arnold BJ, Grant LR, Corander J, Fraser C, et al. Frequency-dependent selection can forecast evolution in Streptococcus pneumoniae. PLoS Biol. 2020;18:e3000878.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    69.
    Johnston C, Campo N, Bergé MJ, Polard P, Claverys JP. Streptococcus pneumoniae, le transformiste. Trends Microbiol. 2014;22:113–9.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    70.
    Marttinen P, Gutmann MU, Croucher NJ, Corander J, Hanage WP. Recombination produces coherent bacterial species clusters in both core and accessory genomes. Micro Genom. 2015;1:e000038.
    Google Scholar 

    71.
    Croucher NJ, Mostowy R, Wymant C, Turner P, Bentley SD, Fraser C. Horizontal DNA Transfer mechanisms of bacteria as weapons of intragenomic conflict. PLoS Biol. 2016;14:e1002394.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    72.
    Apagyi KJ, Fraser C, Croucher NJ. Transformation asymmetry and the evolution of the bacterial accessory genome. Mol Biol Evol. 2018;35:575–81.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    73.
    Spearman C. The proof and measurement of association between two things. Am J Psychol. 1904;100:441–71.
    Article  Google Scholar 

    74.
    Smirnov N. Table for estimating the goodness of fit of empirical distributions. Ann Math Stat. 1948;19:279–81.
    Article  Google Scholar 

    75.
    Kelley T. Statistical method. New York: Macmillan; 1923.
    Google Scholar 

    76.
    Simpson EH. Measurement of diversity. Nature. 1949;163:688.
    Article  Google Scholar 

    77.
    Kimura M. On the probability of fixation of mutant genes in a population. Genetics. 1962;47:713–9.
    CAS  PubMed  PubMed Central  Google Scholar 

    78.
    Hanage WP, Fraser C, Spratt BG. The impact of homologous recombination on the generation of diversity in bacteria. J Theor Biol. 2006;239:210–9.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    79.
    Lawrence JG. Gene transfer in bacteria: speciation without species? Theor Popul Biol. 2002;61:449–60.
    PubMed  Article  PubMed Central  Google Scholar 

    80.
    Pybus OG, Harvey PH. Testing macro-evolutionary models using incomplete molecular phylogenies. Proc R Soc B Biol Sci. 2000;267:2267–72.
    CAS  Article  Google Scholar 

    81.
    Barraclough TG, Birky CW, Burt A. Diversification in sexual and asexual organisms. Evolution. 2003;57:2166–72.
    PubMed  Article  PubMed Central  Google Scholar 

    82.
    Hudson RR. Gene genealogies and the coalescent process. In: Oxford surveys in evolutionary biology. Oxford: Oxford University Press; 1990. p. 1–44.

    83.
    Numminen E, Gutmann M, Shubin M, Marttinen P, Méric G, van Schaik W, et al. The impact of host metapopulation structure on the population genetics of colonizing bacteria. J Theor Biol. 2016;396:53–62.
    PubMed  Article  PubMed Central  Google Scholar 

    84.
    Castillo-Chavez C, Hethcote HW, Andreasen V, Levin SA, Liu WM. Epidemiological models with age structure, proportionate mixing, and cross-immunity. J Math Biol. 1989;27:233–58.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    85.
    McNally A, Kallonen T, Connor C, Abudahab K, Aanensen D, Horner C, et al. Signatures of negative frequency dependent selection in colonisation factors and the evolution of a multi-drug resistant lineage of Escherichia coli. MBio. 2019;10:e00644–19.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    86.
    Kwun MJ, Oggioni MR, Bentley SD, Fraser C, Croucher NJ. Synergistic activity of mobile genetic element defences in Streptococcus pneumoniae. Genes. 2019;10:707.
    CAS  PubMed Central  Article  Google Scholar 

    87.
    Iranzo J, Puigbo P, Lobkovsky AE, Wolf YI, Koonin EV. Inevitability of genetic parasites. Genome Biol Evol. 2016;8:2856–69. https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5631039/.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    88.
    Budroni S, Siena E, Dunning Hotopp JC, Seib KL, Serruto D, Nofroni C, et al. Neisseria meningitidis is structured in clades associated with restriction modification systems that modulate homologous recombination. Proc Natl Acad Sci USA. 2011;108:4494–9.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    89.
    Johnston C, Martin B, Granadel C, Polard P, Claverys JP. Programmed protection of foreign DNA from restriction allows pathogenicity island exchange during pneumococcal transformation. PLoS Pathog. 2013;9:e1003178.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    90.
    Kwun MJ, Oggioni MR, De Ste Croix M, Bentley SD, Croucher NJ. Excision-reintegration at a pneumococcal phase-variable restriction-modification locus drives within- and between-strain epigenetic differentiation and inhibits gene acquisition. Nucleic Acids Res. 2018;46:11438–53.
    CAS  PubMed  PubMed Central  Google Scholar  More

  • in

    Microbial niche differentiation explains nitrite oxidation in marine oxygen minimum zones

    Experimental site
    Seawater samples were collected from two stations (Fig. S1a, offshore OMZ station PS2 and coastal OMZ station PS3) in the ETNP in March and April 2018 on board R/V Sally Ride (Cruise ID: SR 1805). NO2− kinetics and O2 effects experiments (≤1 day incubation) were performed at the oxycline, the oxic/anoxic interface (top) of the ODZ and the core of the ODZ at each station (Fig. S1b and Table S1). Long incubations (6–7 days) were performed at the core of the ODZ at both stations to investigate nitrogen reaction rates at very low O2 concentrations.
    Sampling and incubation experiments for NO2− kinetics and O2 effects
    Twelve 30-L Niskin bottles on a rosette with a conductivity–temperature–depth (CTD) profiler were used to collect seawater while recording in situ O2 concentration, temperature, pressure, salinity, and chlorophyll fluorescence. O2 concentration at selected ODZ depths was measured by STOX sensor (on the CTD profiler) with detection limit of 10 nM [42]. NO2− concentrations were measured by standard spectrophotometric methods onboard. O2 and NO2− concentrations were used to select sampling depths. NO3− and NO2− concentrations in incubation samples were measured on a chemiluminescence NO/NOx Analyzer (Teledyne API, San Diego, CA, USA) in the lab.
    Seawater was collected from Niskin bottles into 60 mL air-tight serum bottles after overflowing three times in order to minimize O2 contamination. Serum bottles were sealed with rubber septa and aluminum seals ensuring the absence of bubbles inside bottles. Septa were deoxygenated in anaerobic chambers with three cycles of vacuum/helium flushing over a period of 1 month prior to the cruise. A helium headspace was created for each sample collected from anoxic depths (the top and the core of the ODZ), and then samples were flushed with helium for at least 15 min to remove O2 that might have been introduced during sampling.
    To determine a single NO2− oxidation rate, a set of five serum bottles amended with 15NO2− tracer was incubated on board at 12 °C in the dark. Incubations were terminated in time series (one bottle at day zero (T0), two at 0.5 day and two at 1 day) by adding 0.2 mL of saturated ZnCl2. The T0 bottles served as killed controls for both tracer contamination and abiotic reactions. The observed temporal changes in isotopic signals that occurred in the live samples over a few days in the time courses would not be detected if abiotic reactions were taking place during the >3 months that elapsed before all samples were measured on the mass spec.
    In the lab, NO2− was removed from samples using sulfamic acid, and NO3− in serum bottles was converted into N2O using the denitrifier method [43, 44]. Both concentration and isotopic composition of N2O were measured on a mass spectrometer (Delta Vplus, Thermo Fisher Scientific, Waltham, MA, USA) for calculating nitrite oxidation rate from the linear regression of the five nitrate concentrations at the three time points as previously described [3]. NO2− kinetics of NO2− oxidation were determined by measuring NO2− oxidation rates under different 15NO2− tracer concentrations (0.5–13.8 μM). For responses of NO2− oxidation to O2, different amounts (0, 0.2, 0.5, 1, 2, and 5 mL) of O2 saturated seawater collected from the same Niskin bottle were added into serum bottles to achieve different O2 concentrations. O2 concentrations in serum bottles were monitored by optical oxygen sensors with a detection limit of 62.5 nM (PyroScience GmbH, Aachen, Germany).
    Half-saturation constant
    Half-saturation constant (Km) is the nitrite concentration at which nitrite oxidation rate (V) equals half of the potential maximum rate (Vm). The curve fitting tool in MATLAB_R2015a was used to fit the Michaelis–Menten model (Eq. (2)) to determine Km.

    $${{V}} ,=, {{V}}_{mathrm{m}} ,times,left[ {{mathrm{NO}}_2^-} right]{mathrm{/}}left( {left[ {{mathrm{NO}}_2^-} right] ,+, {{K}}_{mathrm{m}}} right).$$
    ( 2)

    Long incubation experiments
    Longer (≥6 days) incubations were performed in 12-mL exetainers. Seawater from the core of the ODZ (250 m at station PS2, 160 m at station PS3) was sampled into 320 mL ground glass-stoppered bottles, which were immediately transferred into a N2 flushed glove bag. 15NO2− was added into these bottles to reach final concentrations of 7.24 and 8.01 µM at stations PS2 and PS3, respectively. 15NO2− labeled seawater was aliquoted into exetainers and capped within the glove bag. The septa had been stored under helium for at least 1 month. Exetainers were purged with helium for 5 min. Every 12 h, microbial activity in triplicate exetainers was terminated by adding 0.05 mL of 50% w/v ZnCl2. In the lab, N2 produced in exetainers was measured on a mass spectrometer (Europa Scientific 20–20, Crewe, UK) [45]. Denitrification and anammox rates were computed from linear regression of N2 produced at three time points. 15NO3− was measured (as described above) in the same exetainers, and NO2– oxidation rates were computed from linear regression of NO3− produced at three time points. O2 was monitored throughout the incubations in parallel exetainer vials using LUMOS sensors [12, 13]. Each O2 production or consumption rate was determined by linear regression of O2 concentrations at 32 time points. One sensing spot, glued inside an exetainer vial, allowed the optode to measure O2 concentrations every 5 min from the outside. Detection limit and resolution of LUMOS sensors was ≈1 nM.
    DNA sampling, extraction, sequencing, metagenomics, and metatranscriptomics analysis
    Particulate DNA samples were collected by filtration onto 0.22 µm Sterivex filters from the ODZ core samples (250 m at station PS2, 160 m at station PS3). DNA was extracted using the modified plant tissue protocol (All Prep DNA/RNA Mini Kit, Qiagen, Valencia, CA, USA), and subjected to paired-end sequencing on an Illumina MiSeq to generate over 10 million read pairs for each sample by the Genomics Core Facility of Lewis-Sigler Institute for Integrative Genomics at Princeton University. Quality control of raw reads was performed by BBDuk (DOE Joint Genome Institute, Walnut Creek, CA, USA), and assembled into contigs using metaSPAdes v3.12.0 [46] with specified options (-k 21, 33, 55, 77, 99, 127 -m 500). Metagenome-assembled genomes (MAGs) were constructed using MetaBAT v2.12.1 [47] with default (sensitive) mode and contigs longer than 1500 bp. The quality of MAGs was determined by checkM [48]. Taxonomy of MAGs was predicted by GTDB-tk [49]. ETNP PS2 MAG-11 from the ODZ core at station PS2 and ETNP PS3 MAG-54 from the ODZ core at station PS3 were identified as Nitrospina-like NOB. The taxonomy of these two NOB MAGs (ETNP PS2 MAG-11 and ETNP PS3 MAG-54) was further confirmed by comparing them with known OMZ NOB MAGs (MAG-1 and MAG-2) [8] using ANI. ANI between MAGs was assessed using enveomics [50]. MAG-2 [8], ETNP PS2 MAG-11 and ETNP PS3 MAG-54 belong to the same species (threshold for species: ANI ≥95%) based on their ANI values (Table S3). Considering the low completeness of MAG-11 and the high contamination of MAG-54 constructed from the two new metagenomes here (Table S4), we decided to use MAG-2 as the representative for this NOB species.
    We estimated the relative abundance and the transcriptional activity of the two known OMZ NOB species in different oceanic regions (including the two ETNP stations in this study) by mapping metagenomes and metatranscriptomes from this and other studies to MAG-1 and MAG-2. The relative abundance of MAG-1 or MAG-2 at stations PS2 and PS3 in the ETNP was calculated as RPKG (the number of metagenomic reads obtained in this study mapped to a MAG per MAG length (kb) per genome equivalents) [15]. Genome equivalents were estimated using MicrobeCensus v1.1.1 [15]. The transcriptional activity of MAG-1 or MAG-2 in ETNP and ETSP OMZs was assessed by mapping published metatranscriptomic reads from the ETNP (PRJNA263621) [51] and the ETSP (SRA023632.1) [52] to MAG-1 and MAG-2. The relative abundance of RNA in Fig. S4 was calculated as the number of metatranscriptomic reads mapped to a MAG divided by the number of total reads. Mapping was performed by Bowtie2 [53] using “very-sensitive” mode, and only reads with a mapping quality above 20 were included as mapped reads.
    In order to explore the possibility of the presence of other NOB in the ODZ core at stations PS2 and PS3, we also estimated the relative abundance of other (putative) NOB using their marker gene, nxrB (nitrite oxidoreductase). First, we downloaded previously identified (putative) nxrB sequences from the ETSP OMZ [8]. Then, we color coded the genes in Fig. 4 based on previously defined clusters in a phylogenetic tree (see Fig. 3 in [8]): the Nitrospina cluster (blue) contains nxrB grouped with cultured marine NOB, Nitrospina gracilis. Based on BLASTp search results, amino acid sequence identities between all the OMZ nxrB in this cluster and that of Nitrospina gracilis were 96.71%, 96.71%, and 96.87% for NODE_69234, NODE_114897, and NODE_102582, respectively. Only this cluster is associated with known NOB, and the nitrite oxidation capacity of all the other genes associated with nxrB needs to be confirmed. The anammox cluster (red) contains anammox nxrB sequences. The putative cluster (black) contains nxrB grouped with microbes in which nitrite oxidation capacity has not been proven. This putative cluster fell between known NOB and anammox, and was implied to represent unidentified NOB [54]. The last cluster is called unknown nxrB (gray) because neither their phylogeny nor function can be determined based on their distant relationship with known NOB or anammox nxrB. Finally, we mapped the ETNP metagenomic reads obtained in this study to each nxrB gene. Relative abundance of nxrB genes was also expressed as RPKG: relative abundance of nxrB gene = (number of mapped reads to a certain nxrB gene)/(length of this nxrB in kb)/(genome equivalents).
    To explore the potential metabolisms of NOB in anoxic ODZ cores, we searched for chlorite dismutase and NO dismutase (nod) genes in the NOB MAGs. First, protein-coding sequences in the two new NOB MAGs obtained here (ETNP PS2 MAG-11 and ETNP PS3 MAG-54) were predicted by Prodigal v2.6.3 [55]. The protein-coding sequences were annotated by the best BLASTp hits against the nr protein database. DNA sequences encoding Cld were identified in both ETNP NOB MAGs. Predicted Cld amino acid sequences encoded by ETNP MAGs were too short (120 aa) to be compared to Cld of Candidatus Nitrospira defluvii, but they only had one mismatch with Cld amino acid sequences of MAG-2 from the ETSP OMZ based on MUSCLE alignment using MEGA 7 software (Fig. S5b). Thus, we looked for the arginine173, the Cld activity marker, in longer Cld sequences of MAG-1 and MAG-2 by aligning their Cld with the Cld of Candidatus Nitrospira defluvii using MUSCLE in MEGA 7 software. Since nod was not found in the ETNP MAGs via annotation and was not reported in MAG-1 and MAG-2 in the previous study [8], gene search (i.e., searching nod against NOB MAGs) was performed using Hidden Markov Models by HMMER3 [56]. Reference sequences of the search included a nod sequence retrieved from Candidatus Methylomirabilis oxyfera [32] genome (accession numbers FP565575.1), and three environmental nod sequences (accession numbers: KX364450.1, KX364454.1, and KU933965.1). Search queries were the two ETNP MAGs and two ETSP MAGs in Table S3.
    Estimation of NO2− oxidation through disproportionation using an inverse isotope model
    We simulated the distribution of NO2− oxidation rates via disproportionation using a 1-D inverse isotope model [37] in the anoxic ODZ core from two stations in the ETSP OMZ (Fig. S6), where the complete suite of isotope and rate data have been previously published (Table S7). Briefly, the net biochemical production or consumption rate of each nitrogen compound (R14Ammonium, R14Nitrite, R14Nitrate, R15Nitrite, and R15Nitrate) was balanced by vertical diffusion and advection at steady state, and then five equations were used to balance measured 14NH4+, 14NO2−, 15NO2−, 14NO3−, and 15NO3− concentrations (Eqs. (3–7)). The exclusion of horizontal processes was justified [37] since the model was not run all the way to the surface, and using a constant vertical advection term does not violate continuity of the model. F is the rate of each nitrogen cycle process represented in this model (Fig. 5a, b). We modified the previous model by replacing canonical NO2− oxidation with NO2− disproportionation and using the recently determined stoichiometry of nitrate production by anammox from 0.3 to 0.16 [24]. Since OMZ NOB genomes encode nitrite oxidoreductase (catalyzes NO2−→NO3−) and nitrite reductase (catalyzes NO2−→NO), we assumed that the fractionation factor of nitrate production from nitrite by disproportionation is the same as by canonical nitrite oxidation (αDis = αNxr) based on the close clustering between OMZ NOB nxr and aerobic NOB (Nitrospina gracilis) nxr [8], and the fractionation factor of N2 produced by disproportionation is the same as by denitrification (αDisN2 = αNir). Rates of NO2− disproportionation (FDis), NO3− reduction (FNar), denitrification (FNir), and anammox (FAmx) were solved from the equations by the nonnegative least squares optimization routine (lsqnonneg) in MATLAB_R2015a as described previously [37].

    $$R_{14Ammonium} ,=, 0.11,times {,}^{14}F_{Nir} ,+, 0.07,times {,}^{14}F_{Nar} ,- {,}^{14}F_{Amx},$$
    ( 3)

    $$R_{14Nitrite} = – ,^{14}F_{Nir} ,+, {,}^{14}F_{Nar} ,-, (1 ,+, c) ,times,{,}^{14}F_{Amx} \ -, left( {5/3} right) ,times,{,}^{14}F_{Dis},$$
    ( 4)

    $$R_{14Nitrate} ,=, – ^{14}F_{Nar} ,+, c ,times {,}^{14}F_{Amx} ,+, {,}^{14}F_{Dis},$$
    ( 5)

    $$R_{15Nitrite} = -!! {,}^{14}F_{Nir}/alpha_{Nir} ,times,left( {left[ {{,}^{15}{rm{NO}}_{2}^{-}} right]/left[ {{,}^{14}{rm{NO}}_{2}^{-}} right]} right) \ +, {,}^{14}F_{Nar}/alpha_{Nar},times,left( {left[ {{,}^{15}{rm{NO}}_{3}^{-}} right]/left[ {{,}^{14}{rm{NO}}_{3}^{-}} right]} right)\ -, {,}^{14}F_{Amx}/alpha_{Amx},times,left( {left[ {{,}^{15}{rm{NO}}_{2}^{-}} right]/left[ {{,}^{14}{rm{NO}}_{2}^{-}} right]} right) \ -, c times !{,}^{14}F_{Amx}/alpha_{NxrAmx},times,left( {left[ {{,}^{15}{rm{NO}}_{2}^{-}} right]/left[ {{,}^{14}{rm{NO}}_{2}^{-}} right]} right)\ -, {,}^{14}F_{Dis}/alpha_{Dis}timesleft( {left[ {{,}^{15}{rm{NO}}_{2}^{-}} right]/left[ {{,}^{14}{rm{NO}}_{2}^{-}} right]} right) \ -, left( {2/3} right) !times ! {,}^{14}F_{Dis}/alpha_{DisN2} !times left(! {left[ {{,}^{15}{rm{NO}}_{2}^{-}} right]/left[ !{{,}^{14}{rm{NO}}_{2}^{-}} right]} !right),$$
    (6)

    $$R_{15Nitrate} = -!! {,}^{14}F_{Nar}/alpha_{Nar} ,times,left( {left[ {{,}^{15}{rm{NO}}_{3}^{-}} right]/left[ {{,}^{14}{rm{NO}}_{3}^{-}} right]} right) \ +, c times !{,}^{14}F_{Amx}/alpha_{NxrAmx},times,left( {left[ {{,}^{15}{rm{NO}}_{2}^{-}} right]/left[ {{,}^{14}{rm{NO}}_{2}^{-}} right]} right)\ +, ^{14}F_{Dis}/alpha_{Dis},times,left( {left[ {{,}^{15}{rm{NO}}_{2}^{-}} right]/left[ {{,}^{14}{rm{NO}}_{2}^{-}} right]} right).$$
    (7) More

  • in

    Special issue: CO2: capture of, utilization of, and degradation into

    1.
    Liang J, Ye S, Wang S, Xiao M, Meng Y. Design and structure of catalysts: syntheses of carbon dioxide-based copolymers with cyclic anhydrides and/or cyclic esters. Polym. J. https://doi.org/10.1038/s41428-020-0374-1.
    2.
    Bhat GA, Darensbourg MY, Darensbourg DJ. Copolymerization of propylene oxide and 13CO2 to afford completely alternating regioregular 13C-labelled Poly(propylene carbonate). Polym. J. https://doi.org/10.1038/s41428-020-0391-0.

    3.
    Andrea KA, Kerton FM. Iron-catalyzed reactions of CO2 and epoxides to yield cyclic and polycarbonates. Polym. J. https://doi.org/10.1038/s41428-020-00395-6.

    4.
    Honda M, Ebihara T, Ohkawa T, Sugimoto H. Alternating terpolymerization of carbon dioxide, propylene oxide, and various epoxides with bulky side groups for the tuning of thermal properties. Polym. J. https://doi.org/10.1038/s41428-020-00412-8.

    5.
    Nakabayashi Y, Nakano K. Polycarbonate-block-polycycloalkenes via epoxide/carbon dioxide copolymerization and ring-opening metathesis polymerization. Polym. J. https://doi.org/10.1038/s41428-020-00423-5.

    6.
    Kobayashi K, Pagot G, Vezzù K, Bertasi F, Noto VD, Tominaga Y. Effect of plasticizer on the ion-conductive and dielectric behavior of poly(ethylene carbonate)-based Li electrolytes. Polym. J. https://doi.org/10.1038/s41428-020-00397-4.

    7.
    Taniguchi I, Kinugasa K, Toyoda M, Minezaki K, Tanaka H, Mitsuhara K. Piperazine-immobilized polymeric membranes for CO2 capture: mechanism of preferential CO2 permeation. Polym. J. https://doi.org/10.1038/s41428-020-0389-7.

    8.
    Kamio E, Minakata M, Iida Y, Yasui T, Matsuoka A, Matsuyama H. Inorganic/organic double-network ion gel membrane with a high ionic liquid content for CO2 separation. Polym. J. https://doi.org/10.1038/s41428-020-0393-y.

    9.
    Yamada H. Amine-based capture of CO2 for utilization and storage. Polym. J. https://doi.org/10.1038/s41428-020-00400-y.

    10.
    Honda R, Hamasaki A, Miura Y, Hoshino Y. Thermoresponsive CO2 absorbent for various CO2 concentrations: tuning the pKa of ammonium ions for effective carbon capture. Polym. J. https://doi.org/10.1038/s41428-020-00407-5.

    11.
    Fujikawa S, Selyanchyn R, Kunitake T. A new strategy of membrane-based direct air capture. Polym. J. https://doi.org/10.1038/s41428-020-00429-z.

    12.
    Hairudin NHBM, Ganesan S, Sudesh K. Revalorization of adsorbed residual oil in spent bleaching clay as a sole carbon source for polyhydroxyalkanoate (PHA) accumulation in Cupriavidus necator Re2058/pCB113. Polym. J. https://doi.org/10.1038/s41428-020-00418-2.

    13.
    Thorbecke R, Yamamoto M, Miyahara Y, Oota M, Mizuno S, Tsuge T. The gene dosage effect of carbonic anhydrase on the biosynthesis of poly(3-hydroxybutyrate) under autotrophic and mixotrophic culture conditions. Polym. J. https://doi.org/10.1038/s41428-020-00409-3.

    14.
    Suzuki M, Tachibana Y, Kasuya K. Biodegradability of poly(3-hydroxyalkanoate) and poly(ε-caprolactone) via biological carbon cycles in marine environments. Polym. J. https://doi.org/10.1038/s41428-020-00396-5.

    15.
    Taguchi S, Matsumoto K. Evolution of polyhydroxyalkanoate synthesizing systems toward a sustainable plastic industry. Polym. J. https://doi.org/10.1038/s41428-020-00420-8.

    16.
    Castro LM, Foong CP, Higuchi-Takeuchi M, Morisaki K, Lopes EF, Numata K, Mota AJ. Microbial prospection of an Amazonian blackwater lake and whole-genome sequencing of bacteria capable of polyhydroxyalkanoate synthesis. Polym. J. https://doi.org/10.1038/s41428-020-00424-4.

    17.
    Okeyoshi K, Okajima MK, Kaneko T. The cyanobacterial polysaccharide sacran: characteristics, structures, and preparation of LC gels. Polym. J. https://doi.org/10.1038/s41428-020-00426-2.

    18.
    Fujisawa S. Material design of nanocellulose/polymer composites via Pickering emulsion templating. Polym. J. https://doi.org/10.1038/s41428-020-00408-4.

    19.
    Chen J, Ohta Y, Nakamura H, Masunaga H, Numata K. Aqueous spinning system with a citrate buffer for highly extensible silk fibers. Polym. J. https://doi.org/10.1038/s41428-020-00419-1. More

  • in

    Estimating and explaining the spread of COVID-19 at the county level in the USA

    Data selection and handling: death data
    For mortality due to COVID-19, we used time series provided by the New York Times12. We selected the New York Times dataset because it is rigorously curated. We analyzed separately only counties that had records of 100 or more deaths by 23 May, 2020. The threshold of 100 was a balance between including more counties and obtaining reliable estimates of r(t). Preliminary simulations showed that time series with low numbers of deaths would bias r(t) estimates (Supplementary Fig. 2). However, we did not want to use the maximum daily number of deaths as a selection criterion, because this could lead to selection of counties based on data from a single day. It would also involve some circularity, because the information obtained, r(t), would be directly related to the criterion used to select datasets. Therefore, we used the threshold of 100 cumulative deaths. The District of Columbia was treated as a county. Also, because the New York Times dataset aggregated the five boroughs of New York City, we treated them as a single county. For counties with fewer than 100 deaths, we aggregated mortality to the state level to create a single time series. For thirteen states (AK, DE, HI, ID, ME, MT, ND, NH, SD, UT, VM, WV, and WY), the aggregated time series did not contain 100 or more deaths and were therefore not analyzed.
    Data selection and handling: explanatory county-level variables
    County-level variables were collected from several public data sources36,37,38,39,40,41,42. We selected socio-economic variables a priori in part to represent a broad set of population characteristics.
    Time series analysis: time series model
    We used a time-varying autoregressive model15,56 designed explicitly to estimate the rate of increase of a variable using nonlinear, state-dependent error terms16. We assume in our analyses that the susceptible proportion of the population represented by a time series is close to one, and therefore there is no decrease in the infection rate caused by a pool of individuals who were infected, recovered, and were then immune to further infection.
    The model is

    $$xleft( t right) = rleft( {t-1} right) + xleft( {t-1} right)$$
    (1a)

    $$rleft( t right) = rleft( {t-1} right) + omega_rleft( t right)$$
    (1b)

    $$Dleft( t right) = {mathrm{exp}}(xleft( t right) + phi left( t right))$$
    (1c)

    Here, x(t) is the unobserved, log-transformed value of daily deaths at time t, and D(t) is the observed count that depends on the observation uncertainty described by the random variable ϕ(t). Because a few of the datasets that we analyzed had zeros, we replaced zeros with 0.5 before log-transformation. The model assumes that the death count increases exponentially at rate r(t), where the latent state variable r(t) changes through time as a random walk with ωr(t) ~ N(0, σ2r). We assume that the count data follow a quasi-Poisson distribution. Thus, the expectation of counts at time t is exp(x(t)), and the variance is proportional to this expectation.
    We fit the model using the extended Kalman filter to compute the maximum likelihood57,58. In addition to the parameters σ2r and σ2ϕ, we estimated the initial value of r(t) at the start of the time series, r0, and the initial value of x(t), x0. The estimation also requires terms for the variances in x0 and r0, which we assumed were zero and σ2r, respectively. In the validation using simulated data (Supplementary Methods: Simulation model), we found that the estimation process tended to absorb σ2r to zero too often. To eliminate this absorption to zero, we imposed a minimum of 0.02 on σ2r.
    Time series analysis: parametric bootstrapping
    To generate approximate confidence intervals for the time-varying estimates of r(t) (Eq. 1b), we used a parametric bootstrap designed to simulate datasets with the same characteristics as the real data that are then refit using the autoregressive model. We used bootstrapping to obtain confidence intervals, because an initial simulation study showed that standard methods, such as obtaining the variance of r(t) from the Kalman filter, were too conservative (the confidence intervals too narrow) when the number of counts was small. Furthermore, parametric bootstrapping can reveal bias and other features of a model, such as the lags we found during model fitting (Supplementary Fig. 1a, b).
    Changes in r(t) consist of unbiased day-to-day variation and the biased deviations that lead to longer-term changes in r(t). The bootstrap treats the day-to-day variation as a random variable while preserving the biased deviations that generate longer-term changes in r(t). Specifically, the bootstrap was performed by calculating the differences between successive estimates of r(t), Δr(t) = r(t) – r(t-1), and then standardizing to remove the bias, Δrs(t) = Δr(t) – E[Δr(t)], where E[] denotes the expected value. The sequence Δrs(t) was fit using an autoregressive time-series model with time lag 1, AR(1), to preserve any shorter-term autocorrelation in the data. For the bootstrap, a new time series was simulated from this AR(1) model, Δρ(t), and then standardized, Δρs(t) = Δρ(t) – E[Δρ(t)]. The simulated time series for the spread rate was constructed as ρ(t) = r(t) + Δρs(t)/21/2, where dividing by 21/2 accounts for the fact that Δρs(t) was calculated from the difference between successive values of r(t). A new time series of count data, ξ(t), was then generated using equation 1 with the parameters from fitting the data. Finally, the statistical model was fit to the reconstructed ξ(t). In this refitting, we fixed the variance in r(t), σ2r, to the same value as estimated from the data. Therefore, the bootstrap confidence intervals are conditional of the estimate of σ2r.
    Time series analysis: calculating R0
    We derived estimates of R(t) directly from r(t) using the Dublin-Lotka equation21 from demography. This equation is derived from a convolution of the distribution of births under the assumption of exponential population growth. In our case, the “birth” of COVID-19 is the secondary infection of susceptible hosts leading to death, and the assumption of exponential population growth is equivalent to assuming that the initial rate of spread of the disease is exponential, as is the case in equation 1. Thus,

    $$Rleft( t right) = 1/mathop {sum}nolimits_{_tau} {{mathrm{e}}^{ – r(t)}} tau p(tau)$$
    (2)

    where p(τ) is the distribution of the proportion of secondary infections caused by a primary infection that occurred τ days previously. We used the distribution of p(τ) from Li et al.59 that had an average serial interval of T0 = 7.5 days; smaller or larger values of T0, and greater or lesser variance in p(τ), will decrease or increase R(t) but will not change the pattern in R(t) through time. Note that the uncertainty in the distribution of serial times for COVID-19 is a major reason why we focus on estimating r0, rather than R0: the estimates of r0 are not contingent on time distributions that are poorly known. Computing R(t) from r(t) also does not depend on the mean or variance in time between secondary infection and death. We report values of R(t) at dates that are offset by 18 days, the average length of time between initial infection and death given by Zhou et al.60.
    Time series analysis: Initial date of the time series
    Many time series consisted of initial periods containing zeros that were uninformative. As the initial date for the time series, we chose the day on which the estimated daily death count exceeded 1. To estimate the daily death count, we fit a Generalized Additive Mixed Model (GAMM) to the death data while accounting for autocorrelation and greater measurement error at low counts using the R package mgcv61. We used this procedure, rather than using a threshold of the raw death count, because the raw death count will include variability due to sampling small numbers of deaths. Applying the GAMM to “smooth” over the variation in count data gives a well-justified method for standardizing the initial dates for each time series.
    Time series analysis: validation
    We performed extensive simulations to validate the time-series analysis approach (Supplementary Methods: Simulation model).
    Regression analysis for r 0
    We applied a Generalized Least Squares (GLS) regression model to explain the variation in estimates of r0 from the 160 county and county-aggregate time series:

    $$r_0 = b_0 + b_1start.date + b_2logleft( {pop.size} right) + b_3pop.den^{0.25} + varepsilon$$
    (3)

    where start.date is the Julian date of the start of the time series, log(pop.size) and pop.den0.25 are the log-transformed population size and 0.25 power-transformed population density of the county or county-aggregate, respectively, and ε is a multivariate Gaussian random variable with covariance matrix σ2Σ. We used the transforms log(pop.size) and pop.den0.25 to account for nonlinear relationships with r0; these transforms give the highest maximum likelihood of the overall regression. The covariance matrix contains a spatial correlation matrix of the form C = uI + (1–u)S(g) where u is the nugget and S(g) contains elements exp(−dij/g), where dij is the distance between spatial locations and g is the range62. To incorporate differences in the precision of the estimates of r0 among time series, we weighted by the vector of their standard errors, s, so that Σ = diag(s) * C * diag(s), where * denotes matrix multiplication. With this weighting, the overall scaling term for the variance, σ2, will equal 1 if the residual variance of the regression model matches the square of the standard errors of the estimates of r0 from the time series. We fit the regression model with the function gls() in the R package nlme63.
    To make predictions for new values of r0, we used the relationship

    $$hat e_{mathrm{i}} = bar e + {mathbf{v}}_{mathbf{i}} ast ,{mathbf{V}}^{ – 1}(epsilon_i – bar e)$$
    (4)

    where ει is the GLS residual for data i, (hat e)i is the predicted residual, (bar e) is the mean of the GLS residuals, V is the covariance matrix for data other than i, and vi is a row vector containing the covariances between data i and the other data in the dataset64. This equation was used for three purposes. First, we used it to compute R2pred for the regression model by removing each data point, recomputing (hat e)i, and using these values to compute the predicted residual variance23. Second, we used it to obtain predicted values of r0, and subsequently R0, for the 160 counties and county-aggregates for which r0 was also estimated from time series. Third, we used equation (4) to obtain predicted values of r0, and hence predicted R0, for all other counties. We also calculated the variance of the estimates from64

    $$hat v_{mathrm{i}} = sigma^2-{mathbf{v}}_{mathbf{i}} ast ,{mathbf{V}}^{ – 1} ast v_i^{mathbf{t}}$$
    (5)

    Predicted values of R0 were mapped using the R package usmap65.
    Regression analysis for SARS-CoV-2 effects on r0
    The GISAID metadata27 for SARS-CoV-2 contains the clade and state-level location for strains in the USA; strains G, GH, and GR contain the G614 mutation. For each state, we limited the SARS-CoV-2 genomes to those collected no more than 30 days following the onset of outbreak that we used as the starting point for the time series from which we estimated r0; from these genomes (totaling 5290 from all states), we calculated the proportion that had the G614 mutation. We limited the analyses to the 28 states that had five or more genome samples. For each state, we selected the estimates of r0 from the county or county-aggregate representing the greatest number of deaths. We fit these estimates of r0 with the weighted Least Squares (LS) model as in equation (3) with additional variables for strain. Figure 3 was constructed using the R packages usmap65 and scatterpie66.
    Statistics and reproducibility
    The statistics for this study are summarized in the preceding sections of the “Methods”. No experiments were conducted, so experimental reproducibility is not an issue. Nonetheless, we repeated analyses using alternative datasets giving county-level characteristics, and also an alternative dataset on SARS-CoV-2 strains (Supplementary Methods: Analysis of Nextstrain metadata of SARS-CoV-2 strains), and all of the conclusions were the same.
    Reporting summary
    Further information on research design is available in the Nature Research Reporting Summary linked to this article. More

  • in

    Travails of an intrepid platypus counter

    As a wildlife ecologist, I study many species, but the platypus (Ornithorhynchus anatinus) is my speciality. I’ve worked with these duck-billed monotremes, which are indigenous to eastern Australia, for 13 years. So much remains unknown about them — as solitary animals, they are very hard to capture. The males have a venomous spur on each of their hind feet, so I always hold them by the tail. Luckily, I’ve never been spurred.
    At the environmental consultancy firm Cesar Australia, we’ve focused on whether platypuses’ numbers are rising or falling. Their aquatic habitat has shrunk by 25% in the past 30 years — partly owing to reduced rainfall from climate change and partly to residential and industrial development, leaving less water, which is of poorer quality.
    To monitor the population’s size and health, we do overnight trapping surveys — typically during the spring breeding season and in the autumn, when the babies emerge. For each survey, we set up nets along 5–10 kilometres of waterway and then drive round checking them through the night. It’s a long, sleepless process that usually yields just one or two catches. In the 15 minutes or so for which each animal is out of the water, we weigh and measure it. If it’s carrying an identification chip, we know we’ve caught it before. If not, we implant one in the animal. In this picture, I am releasing a platypus back into a stream outside Melbourne.
    In the past five years, we’ve also been sampling environmental DNA for signatures of platypuses’ presence. It’s been a game-changer in terms of mapping their distribution. We can take a water sample, extract all the DNA and confidently identify whether platypus DNA is present. We’ve surveyed more than 2,000 sites spanning 500,000 square kilometres, enabling us to measure at scales never before possible.
    We’ve also developed a citizen-science platform through which people can submit platypus sightings. This helps us still further. More

  • in

    Empirical support for the biogeochemical niche hypothesis in forest trees

    1.
    Tracy, C. R. & Christian, K. A. Ecological relations among space, time, and thermal niche axes. Ecology 67, 609–615 (1986).
    Article  Google Scholar 
    2.
    Peterson, A. T., Soberon, J. & Sanchez-Cordero, V. Conservatism of ecological niches in evolutionary time. Science 285, 1265–1267 (1999).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    3.
    Hirzel, A. H., Hausser, J., Chessel, D. & Perrin, N. Ecological-niche factor analysis: how to compute habitat-suitability maps without absence data? Ecology 83, 2017–2036 (2002).
    Article  Google Scholar 

    4.
    Sexton, J. P., Montiel, J., Shay, J. E., Stephens, M. R. & Slatyer, R. A. Evolution of ecological niche breadth. Rev. Ecol. Evol. Syst. 48, 183–206 (2017).
    Article  Google Scholar 

    5.
    Wright, J. W., Davies, K. F., Lau, J. A., McCall, A. C. & McKay, J. K. Experimental verification of ecological niche modelling in a heterogeneous environment. Ecology 87, 2433–2439 (2006).
    PubMed  Article  PubMed Central  Google Scholar 

    6.
    Swanson, H. K. et al. A new probabilistic method for quantifying n-dimensional ecological niches and niche overlap. Ecology 96, 318–324 (2015).
    PubMed  Article  PubMed Central  Google Scholar 

    7.
    Grubb, P. J. The maintenance of species-richness in plant communities: the importance of the regeneration niche. Biol. Rev. 52, 107–145 (1977).
    Article  Google Scholar 

    8.
    Herrel, A., Spithoven, L., Van Damme, V. & De Vree, F. Sexual dimorphism of head size in Gallotia galloti: testing the divergence hypothesis by functional analyses. Funct. Ecol. 13, 289–297 (1999).
    Article  Google Scholar 

    9.
    Mouillot, D. et al. Niche overlap estimates based on quantitative functional traits: a new family of non-parametric indices. Oecologia 145, 345–353 (2005).
    PubMed  Article  PubMed Central  Google Scholar 

    10.
    Kraft, N., Valencia, R. & Ackerly, D. D. Functional traits and niche based tree community assembly in an Amazonian forest. Science 322, 580–582 (2008).
    CAS  PubMed  Article  Google Scholar 

    11.
    Peñuelas, J., Sardans, J., Ogaya, R. & Estiarte, M. Nutrient stoichiometric relations and biogeochemical niche in coexisting plant species: effect of simulated climate change. Pol. J. Ecol. 56, 613–622 (2008).
    Google Scholar 

    12.
    Peñuelas, J. et al. Faster returns on “leaf economics” and different biogeochemical niche in invasive compared with native plant species. Glob. Change Biol. 16, 2171–2185 (2010).
    Article  Google Scholar 

    13.
    Peñuelas, J. et al. The bioelements, the elementome and the “biogeochemical niche”. Ecology 100, e02652 (2019).
    PubMed  Article  Google Scholar 

    14.
    Sardans, J. et al. Factors influencing the foliar elemental composition and stoichiometry in forest trees in Spain. Persp. Plant Ecol. Evol. Syst. 18, 52–69 (2016).
    Article  Google Scholar 

    15.
    Sardans, J. et al. Foliar elemental composition of European forest tree species associated with evolutionary traits and present environmental and competitive conditions. Glob. Ecol. Biogeogr. 24, 240–255 (2015).
    Article  Google Scholar 

    16.
    Urbina, I. et al. Shifts in the elemental composition of plants during a very severe drought. Environ. Exp. Bot. 111, 63–73 (2015).
    CAS  PubMed  Article  Google Scholar 

    17.
    Urbina, I. et al. Plant community composition affects the species biogeochemical niche. Ecosphere 8, e01801 (2017).
    Article  Google Scholar 

    18.
    White, P. J. et al. Testing distinctness of shoot ionomes of angiosperm families using the Rothamsted Park grass continuous hay experiment. N. Phytol. 196, 101–109 (2012).
    CAS  Article  Google Scholar 

    19.
    Kerkhoff, A. J., Fagan, W. F., Elser, J. J. & Enquist, B. J. Phylogenetic and growth form variation in the scaling of nitrogen and phosphorus in the seed plants. Am. Nat. 168, E103–E122 (2006).
    PubMed  Article  Google Scholar 

    20.
    Sun, L. K. et al. Leaf elemental stoichiometry of Tamarix Lour. Species in relation to geographic, climatic, soil, and genetic components in China. Ecol. Eng. 106, 448–457 (2017).
    Article  Google Scholar 

    21.
    Neugebauer, K. et al. Variation in the angiosperm ionome. Physiol. Plant. 163, 306–322 (2018).
    CAS  Article  Google Scholar 

    22.
    Gillman, L. N., Keeling, D. J., Gardner, R. C. & Wright, S. D. Faster evolution of highly conserved in tropical plants. J. Evol. Biol. 23, 1327–1330 (2010).
    CAS  PubMed  Article  Google Scholar 

    23.
    Puurtinen, M. et al. Temperature-dependent mutational robustness can explain faster molecular evolution at warm temperatires, affecting speciation rate and global patterns of species diversity. Ecography 39, 1025–1033 (2016).
    Article  Google Scholar 

    24.
    Kellner, A., Ritz, C. M., Schlittenhaedt, P. & Hellwig, F. H. Genetic differentiation in the genus Lithops L. (Ruschoideae, Aizoaceae) reveals a high level of convergent evolution and reflects geographic distribution. Plant Biol. 13, 368–380 (2011).
    CAS  PubMed  Article  Google Scholar 

    25.
    Jwa, N. S. & Hwang, B. K. Convergent evolution of pathogen effectors toward reactive oxygen species signaling networks in plants. Front. Plant Sci. 8, 1687 (2017).
    PubMed  PubMed Central  Article  Google Scholar 

    26.
    Molina-Montenegro, M. A. et al. Is the success of plant invasions the result of rapid adaptive evolution in seed traits? Evidence from a latitudinal rainfall gradient. Front. Plant Sci. 9, 208 (2018).
    PubMed  PubMed Central  Article  Google Scholar 

    27.
    Anacker, B. L. & Strauss, S. Y. Ecological similarity is related to phylogenetic distance between species in a cross-niche field transplant experiment. Ecology 97, 1807–1818 (2016).
    PubMed  Article  Google Scholar 

    28.
    Reich, P. B. & Oleksyn, J. Global patterns of plant leaf N and P in relation to temperature and latitude. Proc. Natl Acad. Sci. USA 101, 11001–11106 (2004).
    CAS  PubMed  Article  Google Scholar 

    29.
    Ordoñez, J. C. et al. A global study of relationships between leaf traits, climate and soil measures of nutrient fertility. Glob. Ecol. Biogeogr. 18, 137–149 (2009).
    Article  Google Scholar 

    30.
    Kerkhoff, A. J., Enquist, B. J., Elser, J. J. & Fagan, W. F. Plantallometry, stoichiometry and the temperature-dependence of primary productivity. Glob. Ecol. Biogeogr. 14, 585–598 (2005).
    Article  Google Scholar 

    31.
    Yuan, Z. Y. & Chen, H. Y. H. Global trends in senesced-leaf nitrogen and phosphorus. Glob. Ecol. Biogeogr. 18, 532–542 (2009).
    Article  Google Scholar 

    32.
    Vitousek, P. M., Porder, S., Houlton, B. Z. & Chadwick, O. A. Terrestrial phosphorus limitation: mechanisms, implications and nitrogen–phosphorus interactions. Ecol. Appl. 20, 5–15 (2010).
    PubMed  Article  PubMed Central  Google Scholar 

    33.
    McGroddy, M. E., Daufresne, T. & Hedin, L. O. Scaling of C/N/P stoichiometry in forest worldwide: implications of terrestrial Redfield-type ratios. Ecology 85, 2390–2401 (2004).
    Article  Google Scholar 

    34.
    Townsend, A. R., Cleveland, C. C., Asner, G. P. & Bustamante, M. M. C. Controls over foliar N:P ratios in tropical rainforest. Ecology 88, 107–118 (2007).
    PubMed  Article  PubMed Central  Google Scholar 

    35.
    Lovelock, C. E., Feller, I. C., Ball, M. C., Ellis, J. & Sorell, B. Testing the growth rate vs. geochemical hypothesis for latitudinal variation in plant nutrients. Ecol. Lett. 10, 1154–1163 (2007).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    36.
    Marschner, H. Mineral Nutrition of Higher Plants (Academic Press, 1995).

    37.
    Zhang, Y. et al. Log-term trends in total inorganic nitrogen and sulfur deposition in US from 1990 to 2010. Atmos. Chem. Phys. 18, 9091–9106 (2018).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    38.
    Horn, K. J. et al. Growth and survival relationships of 71 tree species with nitrogen and sulfur deposition across the conterminous U.S. PLoS ONE 14, e0212984 (2019).
    PubMed  PubMed Central  Article  Google Scholar 

    39.
    Papanikolaou, N., Britton, A. J., Helliwell, R. C. & Johnson, D. Nitrogen deposition, vegetation burning and climate warming act independently on microbial community structure and enzyme activity associated with decomposing litter in low-alpine heath. Glob. Change Biol. 16, 3120–3132 (2010).
    Google Scholar 

    40.
    Marklein, A. R. & Houlton, B. Z. Nitrogen inputs accelerate phosphorus cycling rates across a wide variety of terrestrial ecosystems. N. Phytol. 193, 696–704 (2012).
    CAS  Article  Google Scholar 

    41.
    Sardans, J. et al. Foliar and soil concentrations and stoichiometry of nitrogen and phosphorus across European Pinus sylvestris forests: relationships with climate, N deposition and tree growth. Funct. Ecol. 30, 676–689 (2016).
    Article  Google Scholar 

    42.
    Peñuelas, J. et al. Human-induced nitrogen–phosphorus imbalances alter natural and managed ecosystems across the globe. Nat. Commun. 4, 2934 (2013).
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    43.
    Penuelas, J. et al. Increasing atmospheric CO2 concentrations correlate with declining nutritional status of European forests. Commun. Biol. 3, 125 (2020).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    44.
    Ahmad, N. & Mermut, A. Vertisols and Technologies for their Development 1st edn, Vol. 24 (Elsevier, 1996).

    45.
    Nishiue, A., Nanzyo, M., Kanno, H. & Takahashi, T. Properties and classification of volcanic ash soils around Lake Kuwanuma on the eastern footslope of Mt. Funagata in Miyagi prefecture, northeastern Japan. Soil Sci. Plant Nutr. 60, 848–862 (2014).
    CAS  Article  Google Scholar 

    46.
    De la Riva, E. G. et al. Biogeochemical and ecomorphological niche segregation of Mediterranean woody species along a local gradient. Fron. Plant Sci. 8, 1242 (2017).
    Article  Google Scholar 

    47.
    Yu, Q. et al. Stoichiometry homeostasis of vascular plants in the inner Mongolia grassland. Oecologia 166, 1–10 (2011).
    PubMed  Article  PubMed Central  Google Scholar 

    48.
    Sardans, J., Albert Rivas-Ubach, A. & Peñuelas, J. The elemental stoichiometry of aquatic and terrestrial ecosystems and its relationships with organismic lifestyle and ecosystem structure and function: a review. Biogeochemistry 111, 1–39 (2012).
    Article  Google Scholar 

    49.
    Gracia, C., Burriel, J. A., Ibàñez, J. J., Mata, T. & Vayreda, J. Inventari ecològic i forestal de Catalunya: regió forestal V (CREAF, 2004).

    50.
    Fick, A. E. & Hijmans, R. J. WorldClim 2: new 1-km spatial resolution climate surfaces for global land areas. Int. J. Climatol. 37, 4302–4315 (2017).
    Article  Google Scholar 

    51.
    Lang, R. Verwitterung und Bodenbildung als Einfuehrung in die Bodenkunde (Schweizerbart Science Publishers, 1920).

    52.
    Köppen, W. Klassification der Klimate nach Tempertur, Niederschlag and Jahreslauf. Petermanns Geog. Mitt. 64, 243–248 (1918).
    Google Scholar 

    53.
    De Martonne, E. Nouvelle carte mondiale de l’indece d’aridité. Ann. Géogr. 51, 242–250 (1942).
    Google Scholar 

    54.
    Emberger, L. La vegetation de la región Mèditerranéenne, essai d’une classification des groupements vegetaux. Rev. Gén. Bot. 42, 641–662, 705–721 (1930).

    55.
    Vorösmarty, C. J. et al. Global threats to human water security and river biodiversity. Nature 467, 555–561 (2010).
    PubMed  Article  CAS  Google Scholar 

    56.
    R Development Core Team R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, 2011).

    57.
    Qian, H. & Jin, Y. An updated megaphylogeny of plants, a tool for generating plant phylogenies, and an analysis of phylogenetic community structure. J. Plant Ecol. 9, 233–239 (2016).
    Article  Google Scholar 

    58.
    Paradis, E., Claude, J. & Strimmer, K. APE: analyses of phylogenetics and evolution in R language. Bioinformatics 20, 289–290 (2004).

    59.
    Revell, L. J. Phytools: an R package for phylogenetic comparative biology (and other things). Methods Ecol. Evol. 3, 217–223 (2012).
    Article  Google Scholar 

    60.
    Blomberg, S. P., Garland, T. & Ives, A. R. Testing for phylogenetic signal in comparative data: behavioral traits are more labile. Evolution 57, 717–745 (2003).
    PubMed  Article  Google Scholar 

    61.
    Münkemüller, T. et al. How to measure and test phylogenetic signal. Methods Ecol. Evol. 3, 743–756 (2012).
    Article  Google Scholar 

    62.
    Pagel, M. Inferring the historical patterns of biological evolution. Nature 401, 877–884 (1999).
    CAS  PubMed  Article  Google Scholar 

    63.
    Felsenstein, J. Phylogenies and the comparative method. Am. Nat. 125, 1–15 (1985).
    Article  Google Scholar 

    64.
    Revell, L. J. Two new graphical methods for mapping trait evolution on phylogenies. Methods Ecol. Evol. 4, 754–759 (2013).
    Article  Google Scholar 

    65.
    Raamsdonk, L. M. et al. A functional genomics strategy that uses metabolome data to reveal the phenotype of silent mutations. Nat. Biotechnol. 19, 45–50 (2001).
    CAS  PubMed  Article  Google Scholar 

    66.
    Hadfield, J. D. MCMC methods for multi-response generalised linear mixed models: the MCMCglmm R package. J. Stat. Softw. 33, 2 (2010). More

  • in

    Author Correction: Tree mode of death and mortality risk factors across Amazon forests

    School of Geography, Earth and Enviornmental Sciences, University of Birmingham, Birmingham, UK
    Adriane Esquivel-Muelbert & Thomas A. M. Pugh

    School of Geography, University of Leeds, Leeds, UK
    Adriane Esquivel-Muelbert, Oliver L. Phillips, Roel J. W. Brienen, Martin J. P. Sullivan, Timothy R. Baker, Emanuel Gloor, Aurora Levesley, Simon L. Lewis, Karina Liana Lisboa Melgaço Ladvocat, Gabriela Lopez-Gonzalez, Nadir Pallqui Camacho, Julie Peacock, Georgia Pickavance & David Galbraith

    Birmingham Institute of Forest Research, University of Birmingham, Birmingham, UK
    Adriane Esquivel-Muelbert & Thomas A. M. Pugh

    School of Geography, Earth and Environmental Sciences, University of Plymouth, Plymouth, UK
    Sophie Fauset

    Department of Natural Sciences, Manchester Metropolitan University, Manchester, UK
    Martin J. P. Sullivan

    International Master Program of Agriculture, National Chung Hsing University, Taichung, Taiwan
    Kuo-Jung Chao

    Geography, College of Life and Environmental Sciences, University of Exeter, Exeter, UK
    Ted R. Feldpausch

    Instituto Nacional de Pesquisas da Amazônia, Manaus, Brazil
    Niro Higuchi, Adriano José Nogueira Lima & Carlos Quesada

    School of Mathematics, University of Leeds, Leeds, UK
    Jeanne Houwing-Duistermaat & Haiyan Liu

    Faculty of Natural Sciences, Department of Life, Imperial College London Sciences, London, UK
    Jon Lloyd

    Environmental Change Institute, School of Geography and the Environment, University of Oxford, Oxford, UK
    Yadvinder Malhi & Simone Matias de Almeida Reis

    UNEMAT – Universidade do Estado de Mato Grosso PPG-Ecologia e Conservação, Campus de Nova Xavantina, Nova Xavantina, MT, Brazil
    Beatriz Marimon, Ben Hur Marimon Junior, Paulo Morandi, Edmar Almeida de Oliveira & Simone Matias de Almeida Reis

    Jardín Botánico de Missouri, Oxapampa, Peru
    Abel Monteagudo-Mendoza, Victor Chama Moscoso, Luis Valenzuela Gamarra & Rodolfo Vasquez Martinez

    Forest Ecology and Forest Management Group, Wageningen University and Research, Wageningen, Netherlands
    Lourens Poorter, Frans Bongers, Marielos Peña-Claros & Pieter Zuidema

    Centro de Ciências Biológicas e da Natureza, Universidade Federal do Acre, Rio Branco, AC, Brazil
    Marcos Silveira

    Instituto de Investigaciones para el Desarrollo Forestal (INDEFOR), Universidad de Los Andes, Mérida, Venezuela
    Emilio Vilanova Torre & Julio Serrano

    University of California, Berkeley, CA, USA
    Emilio Vilanova Torre

    Escuela de Ciencias Agropecuarias y Ambientales, Universidad Nacional Abierta y a Distancia, Boyacá, Colombia
    Esteban Alvarez Dávila

    Fundación ConVida, Medellín, Colombia
    Esteban Alvarez Dávila

    Instituto de Investigaciones de la Amazonia Peruana, Iquitos, Peru
    Jhon del Aguila Pasquel, Nallaret Davila Cardozo & Eurídice Honorio Coronado

    Instituto de Biodiversidade e Florestas, Universidade Federal do Oeste do Pará, Santarém, Brazil
    Everton Almeida

    Center for Tropical Conservation, Nicholas School of the Environment, University in Durham, Durham, NC, USA
    Patricia Alvarez Loayza

    Projeto Dinâmica Biológica de Fragmentos, Instituto Nacional de Pesquisas da Amazônia Florestais, Manaus, AM, Brazil
    Ana Andrade & José Luís Camargo

    National Institute for Space Research (INPE), São José dos Campos, SP, Brazil
    Luiz E. O. C. Aragão

    Museo de Historia Natural Noel Kempff Mercado, Universidad Autónoma Gabriel Rene Moreno, Santa Cruz de la Sierra, Bolivia
    Alejandro Araujo-Murakami & Marisol Toledo

    Wageningen Environmental Research, Wageningen University and Research, Wageningen, Netherlands
    Eric Arets

    Dirección de la Carrera de Biología, Universidad Autónoma Gabriel René Moreno, Santa Cruz de la Sierra, Bolivia
    Luzmila Arroyo

    UNELLEZ-Guanare, Herbario Universitario (PORT), Portuguesa, Venezuela Compensation International Progress S.A. Ciprogress–Greenlife, Bogotá, D.C., Colombia
    Gerardo A. Aymard C.

    INRAE, UMR EcoFoG, CNRS, Cirad, AgroParisTech, Université des Antilles, Université de Guyane, Kourou, France
    Michel Baisie, Damien Bonal, Benoit Burban, Aurélie Dourdain, Maxime Rejou-Machain & Clement Stahl

    Department of Biological Sciences, International Center for Tropical Botany, Florida International University, Miami, FL, USA
    Christopher Baraloto

    Centro de Energia Nuclear na Agricultura, Universidade de São Paulo, Piracicaba, Brazil
    Plínio Barbosa Camargo

    Universidade Federal do Acre, Campus Floresta, Cruzeiro do Sul, Brazil
    Jorcely Barroso

    UR Forest & Societies, CIRAD, Montpellier, France
    Lilian Blanc

    Department of Biology, Utrecht, Netherlands
    René Boot

    Woods Hole Research Center, Falmouth, MA, USA
    Foster Brown

    Laboratório de Botânica e Ecologia Vegetal, Universidade Federal do Acre, Rio Branco, AC, Brazil
    Wendeson Castro

    Laboratoire Evolution et Diversite Biologique, CNRS, Toulouse, France
    Jerome Chave

    Inventory and Monitoring Program, National Park Service, Fort Collins, CO, USA
    James Comiskey

    Proyecto Castaña, Madre de Dios, Peru
    Fernando Cornejo Valverde

    Instituto de Geociências, Faculdade de Meteorologia, Universidade Federal do Para, Belém, Brazil
    Antonio Lola da Costa

    Department of Anthropology and Primate Molecular Ecology and Evolution Laboratory, University of Texas, Austin, TX, USA
    Anthony Di Fiore

    National Museum of Natural History, Smithsonian Institute, Washington, DC, USA
    Terry Erwin

    Universidad Nacional Jorge Basadre de Grohmann, Tacna, Peru
    Gerardo Flores Llampazo

    Museu Paraense Emílio Goeldi, Belém, Brazil
    Ima Célia Guimarães Vieira & Rafael Salomão

    Instituto Venezolano de Investigaciones Científicas (IVIC), Caracas, Venezuela
    Rafael Herrera

    IIAMA, Universitat Politécnica de València, València, Spain
    Rafael Herrera

    Universidad Nacional de San Antonio Abad del Cusco, Cusco, Peru
    Isau Huamantupa-Chuquimaco

    Instituto Amazónico de Investigaciones Imani, Universidad Nacional de Colombia Sede Amazonia, Leticia, Colombia
    Eliana Jimenez-Rojas

    Agteca, Santa Cruz, Bolivia
    Timothy Killeen

    College of Science and Engineering, James Cook University, Cairns, QLD, Australia
    Susan Laurance & William Laurance

    Department of Geography, University College London, London, UK
    Simon L. Lewis

    Environmental Science and Policy, George Mason University, Fairfax, VA, USA
    Thomas Lovejoy

    Research School of Biology, Australian National University, Canberra, ACT, Australia
    Patrick Meir

    School of Geosciences, University of Edinburgh, Edinburgh, UK
    Patrick Meir

    Escuela de Ciencias Forestales, Unidad Académica del Trópico, Universidad Mayor de San Simón, Cochabamba, Bolivia
    Casimiro Mendoza

    Facultad de Ingeniería Ambiental, Universidad Estatal Amazónica, Puyo, Ecuador
    David Neill

    Universidad Nacional de San Antonio Abad del Cusco, Cusco, Perú
    Percy Nuñez Vargas, Nadir Pallqui Camacho & Javier Silva Espejo

    Universidad Autónoma del Beni José Ballivián, Trinidad, Bolivia
    Guido Pardo & Vincent Vos

    Universidad Regional Amazónica Ikiam, Ikiam, Ecuador
    Maria Cristina Peñuela-Mora

    Broward County Parks Recreation, Oakland Park, FL, USA
    John Pipoly

    Keller Science Action Center, Field Museum, Chicago, IL, USA
    Nigel Pitman

    Instituto de Ciencias Naturales, Universidad Nacional de Colombia, Bogotá, Colombia
    Adriana Prieto & Agustín Rudas

    Institute of Research for Forestry Development (INDEFOR), Universidad de los Andes, Mérida, Venezuela
    Hirma Ramirez-Angulo

    Socioecosistemas y Cambio Climatico, Fundacion Con Vida, Medellín, Colombia
    Zorayda Restrepo Correa

    Centro de Conservacion, Investigacion y Manejo de Areas Naturales, CIMA Cordillera Azul, Lima, Peru
    Lily Rodriguez Bayona

    Universidade Federal Rural da Amazônia, Belém, Brazil
    Rafael Salomão & Natalino Silva

    Departamento de Biología, Universidad de La Serena, La Serena, Chile
    Javier Silva Espejo

    Guyana Forestry Commission, Georgetown, Guyana
    James Singh

    Federal University of Alagoas, Maceió, Brazil
    Juliana Stropp

    Institute for Conservation Research, Escondido, CA, USA
    Varun Swamy

    Institute for Transport Studies, University of Leeds, Leeds, UK
    Joey Talbot

    Biodiversity Dynamics, Naturalis Biodiversity Center, Leiden, The Netherlands
    Hans ter Steege

    Systems Ecology, Free University, De Boelelaan 1087, Amsterdam, Netherlands
    Hans ter Steege

    Department of Biology, University of Florida, Gainesville, FL, USA
    John Terborgh

    Iwokrama International Centre for Rainforest Conservation and Development, Georgetown, Guyana
    Raquel Thomas

    Universidad de los Andes, Mérida, Venezuela
    Armando Torres-Lezama

    School of Geography, University of Nottingham, Nottingham, UK
    Geertje van der Heijden

    Van Hall Larenstein University of Applied Sciences, Leeuwarden, Netherlands
    Peter van der Meer

    Van der Hoult Forestry Consulting, Leeuwarden, The Netherlands
    Peter van der Hout

    Núcleo de Estudos e Pesquisas Ambientais – Universidade Estadual de Campinas, Campinas, Brazil
    Simone Aparecida Vieira

    Herbario del Sur de Bolivia, Universidad de San Francisco Xavier de Chuquisaca, Sucre, Bolivia
    Jeanneth Villalobos Cayo

    Tropenbos International, Wageningen, Netherlands
    Roderick Zagt More