More stories

  • in

    Direct quantification of ecological drift at the population level in synthetic bacterial communities

    Assembling and monitoring the synthetic bacterial communities
    We selected three bacterial strains, a Chryseobacterium sp., a Staphylococcus sp., and a Bacillus sp., from a large collection of soil isolates, and we screened them with fluorescence-independent flow cytometry as we did previously [17]. We measured at an acquisition speed of 14 μl min−1 for 2 min per sample and setting a threshold of 10,000 regarding the height signal of the front scatter (FSC-H). Differently than in our previous work, here we acquired all scattering profiles based on growth assays at 30 °C that was the temperature at which we performed all the experiments. In addition, we screened the growing cultures with a temporal resolution of 20, rather than 30, min. We recorded significant interactions among the strains by comparing their single and mixed growth profiles at 30 °C (Dataset 1). These interactions were mainly positive, similarly to what we found previously at other temperatures [17]. We performed all the related growth assays in biological triplicates. All flow cytometry data are available in .fcs format online (http://flowrepository.org) under the “FR-FCM-Z25Q” identifier. Henceforth, when referring to the experiments we use the term “population” to describe the cells of a given strain within a flask at a given assay and the term “community” to describe the total bacterial cells within a flask at a given assay.
    Quantification of “background noise”
    Our flow cytometry method for screening the mixed bacterial cultures has an accuracy of 97% for sample densities above 105 cells ml−1 [17]. However, at lower densities sampling errors and instrument inconsistencies become increasingly important because the signal-to-noise ratio drops. This can result in substantially different counts among identical samples and can artificially inflate the observed variability. Thus, it was essential to quantify this “noise” before performing the main experiments and subtract it from the observed variability when quantifying drift.
    To that end, we made a series of separate experiments to quantify “noise”. In these experiments, we mixed overnight cultures of the three strains in all the seven possible combinations and in final cell densities ranging from 1.6 × 104 to 6.3 × 107 strain−1 ml−1 (corresponding to the expected range of cell densities in the main experiment, see below), and we measured repeated aliquots from the same flask to determine the coefficient of variation (CV—Fig. 1a). We treated the samples in exactly the same way as in the main experiments to include the effect of sampling errors in our calculations. We acquired in total of 99 triplicate measurements of 1–3 populations for a total of 148 observations (Fig. 1a, Dataset 2). We hypothesized that the level of “noise” should be inversely related to the cell density of the sample, because the signal-to-noise ratio decreases at low cell densities in the flow cytometer. Accordingly, we fit different functions for the dependency of CV to cell density (Supplementary Table S1). Finally, we calculated the 99.5% confidence intervals of the best-fitting function (i.e., Michaelis-Menten) using the confint function of the MASS package [18] in R [19], and we defined the false discovery rate based on the number of observations that were above the upper 99.5% confidence interval (Supplementary Fig. S1). Finally, we verified that the levels of noise determined in this study are similar to the variability recorded from technical replicate samples taken during our previous experiment where we used the same bacterial system and instrument with identical settings [17] (Supplementary Fig. S2).
    Main experiments
    To quantify drift, we monitored the changes in population densities across identical starting communities incubated under the same environmental conditions (Fig. 1b). To that end, we mixed the three strains in all seven possible combinations, i.e., three monocultures, three mixed cultures of two strains and a mixed culture of all three strains, and in three different starting total cell densities (5 × 104 cells ml−1, 105 cells ml−1, and 106 cells ml−1). To perform each growth assay, we first inoculated the respective strains from overnight pure cultures in a single flask containing 300 ml of Luria–Bertani medium (Sigma). To reach the desired starting total cell density, we estimated the cell density of the overnight pure cultures with flow cytometry [17] and we inoculated the respective volume. We then mixed the culture thoroughly and we immediately split the volume equally into three flasks. We next sampled 500 μl from each flask and we compared the variability in the bacterial populations across the three flasks to the expected “background noise” for the same cell density. In specific, we examined the CVs of the bacterial populations and their z-scores compared to the “background noise”, i.e., how many standard deviations an observed CV differs from the expected “background noise” CV at a given cell density. If the observed z-scores were larger than 2 (95% CI), we aborted the given experiment because it indicated that we introduced variability when we mixed and split the cultures and thus the starting cultures could not be considered identical; this happened in ~50% of the cases. If the observed z-scores were lower than 2, indicating that the recorded variability was not statistically different or was less than the expected variability based on the “background noise”, we proceeded with the experiment, incubating the three flasks in the same chamber (New Brunswick Innova 42R, Eppendorf) at 30 °C and with shaking at 80 rpm.
    We recorded the starting densities (Dataset 3, z-scores between −6.68 and 1.17) and the densities every 20 min until the end of the fourth hour of incubation starting from the 60th minute. To detect and quantify drift, our main assumption was that any larger-than-expected deviations in the population densities of identical starting communities incubated under the same environmental conditions could only be because of drift. Thus, we compared the observed CVs to the expected CVs based on the “background noise” by deploying the z-score. We quantified drift using two different thresholds:
    1.
    The “upper threshold” that focused on excluding false-positive observations. In this quantification, we used a cutoff significance level of z  > 3 (99.5% CI) and we ignored the lowest 17.57% of positive observations (i.e., 15 observations, corresponding to the FDR level of the “background noise”) to minimize the detection of false positives.

    2.
    The “mean threshold” that focused on excluding false negatives and increasing detectability. In this quantification, we used a cutoff significance level of z  > 0, meaning that we scored any observation greater than the mean noise function as positive.

    The “upper threshold” quantification probably overestimates drift by taking into account only the highest among the recorded CV values while the “mean threshold” quantification underestimates drift by taking into account some low CV values that are very close to the noise levels. Thus, the “upper threshold” and “mean threshold” quantifications do not represent the true levels of drift (which are hard to define whatsoever in the presence of noise) but they rather represent the upper and lower boundaries within which the true levels of drift lie.
    Estimating potential growth variability due to temperature differences within the incubation chamber
    To ensure that the recorded variability in the population counts was not due to slight differences in temperature within the incubation chamber, we estimated the potential variability that could have resulted if each strain grew within the extremes of the recorded temperatures in the chamber. For that, we first measured the temperature within each flask at each experimental time point, five times per flask, using a digital immersion thermometer with an accuracy of 0.1 °C. The temperature varied by 0.15 °C ± 0.08 °C on average and by 0.28 °C at maximum. We then calculated the growth rates of each of the strains under the recorded temperature extremes at each time point by interpolating from previously recorded growth rates [17]. We interpolated both with respect to time and with respect to temperature because the previous data were recorded at intervals of 30 min and 0.5 °C (the latter at a range of 25–42 °C). Finally, for each strain, we calculated how the CV in the hypothetical population densities would increase if the strains were constantly growing within the recorded temperature extremes for the duration of the experiment and if the CV was calculated from three observations (like in the real experiments) of the resulting population density distributions. We note that with this analysis, we probably overestimated the hypothetical increase in population densities because we used growth rates from mixed cultures that were generally higher than those in monocultures because of the positive interactions among the strains (Dataset 1).
    Simulations
    To simulate drift in complex bacterial communities, we used in silico communities with diversity and abundance distributions similar to nature [20] where drift acts with magnitude according to our experimental data. A conceptual flowchart of the simulations can be found in Supplementary Information (Supplementary Fig. S3). Each simulation involved a metacommunity of 100 communities that were connected with dispersal and that initially contained 2000 species each.
    We simulated dispersal occurring in a unidirectional way within a closed system; individuals from community n disperse to the community (n + 1) and individuals from community 100 disperse back to community 1. The strength of dispersal equaled to the percentage of individuals that disperse to the respective community and it varied between 2 and 20%. Our aim in modeling dispersal in this way was to create a setting where habitat fragmentation was high and therefore drift’s importance is expected to be more pronounced [21, 22], and where there was no gain or loss of individuals from outside the metacommunity.
    We simulated selection as differences in the growth rates among species within a community. The growth rates were distributed normally with a mean of 1 (resembling systems at their carrying capacity) and with a standard deviation between 0.071 and 0.167. Growth rates changed at every generation by being re-drawn from the same distribution in an effort to represent fluctuating habitats where a given species is not always favored or disfavored. Therefore, in our simulations, the standard deviation in the growth rates represents the strength of selection, because the higher it is the bigger are the differences in the growth rates in a community and the changes in the growth of a species from generation to generation. The distribution of the abundances in a community at time zero was log-normal (mean = 4, sd = 1.1) and the distribution of the abundances of a given species across all communities was normal with a standard deviation equal to the strength of selection.
    The metacommunity grew for 1000 generations under given dispersal and selection conditions with drift, where drift changed the assigned growth rates at every generation according to a distribution based on the defined threshold from the experimental data (“upper” or “mean” threshold). In parallel, an initially identical metacommunity grew under the same dispersal and selection conditions but without drift, meaning that the assigned growth rates at every generation did not change further. More details and an example on how we modeled changes in growth rates due to drift are presented in the Supplementary Text in Supplementary Information.
    For a given generation, we calculated the effect of drift by comparing a given community in the metacommunity that grew under drift to the same community in the metacommunity that grew without drift. In specific, we examined the β-diversity by means of the Bray–Curtis (BC) community similarity and the differences in species richness and in Pielou’s evenness among drift-impacted and drift-free communities, calculating the metacommunity-wise mean and standard deviation on all these properties. Moreover, we kept track of the extinct species at the end of each simulation and we mapped their initial relative abundances, but here we report the metacommunity-wise median because the distribution of the relative abundances is skewed (Supplementary Fig. S4). We ran simulations under 50 different scenarios resulting from five levels of selection strength over ten levels of dispersal rate. To estimate the effect of drift on Bray–Curtis similarity in metacommunities with increasing number of species, we ran the same simulation at the highest selection and lowest dispersal levels, at intermediate selection and dispersal and at the lowest selection and highest dispersal, but we changed the number of species; we ran the simulation three times in metacommunities of 500, 1000, 2000, 4000, 6000, 8000, and 10,000 species. We performed all simulations in R. All code is available on GitHub (https://github.com/sfodel/Drift).
    Reported β-diversity in stochastically assembled communities in nature
    To compare our simulation results with the results from environmental surveys regarding the β-diversity in stochastically assembled communities, we searched for related studies using the following two criteria: (1) the study cites the works of Stegen and colleagues [12, 13], where the term “undominated” community assembly is presented formally for microbial ecology, (2) the study reports data on the range of the observed β-diversity in terms of Bray–Curtis dissimilarity (or similarity) in stochastically assembled communities, or this range can be inferred from the data presented in that study. More

  • in

    Global allele polymorphism indicates a high rate of allele genesis at a locus under balancing selection

    Amos W (2009) Heterozygosity and mutation rate: evidence for an interaction and its implications. Bioessays 32:82–90
    Google Scholar 

    Amos W, Kosanović D, Eriksson A (2015) Inter-allelic interactions play a major role in microsatellite evolution. Proc R Soc B 282:20152125
    PubMed  Google Scholar 

    Amos W (2016) Heterozygosity increases microsatellite mutation rate. Biol Lett 12:20150929
    PubMed  PubMed Central  Google Scholar 

    Azevedo L, Serrano C, Amorim A, Cooper DN (2015) Trans-species polymorphism in humans and the great apes is generally maintained by balancing selection that modulates the host immune response. Hum Genomics 9:21
    PubMed  PubMed Central  Google Scholar 

    Beye M, Hasselmann M, Fondrk MK, Page RE, Omholt SW (2003) The gene csd is the primary signal for sexual development in the honeybee and encodes an SR-type protein. Cell 114:419–429
    CAS  PubMed  Google Scholar 

    Beye M, Seelmann C, Gempe T, Hasselmann M, Vekemans X, Fondrk MK et al. (2013) Gradual molecular evolution of a sex determination switch through incomplete penetrance of femaleness. Curr Biol 23:2559–2564
    CAS  PubMed  Google Scholar 

    Bezabih G, Adgaba N, Hepburn HR, Pirk CWW (2014) The territorial invasion of Apis florea in Africa. Afr Entomol 22:888–890
    Google Scholar 

    Biewer M, Lechner S, Hasselmann M (2016) Similar but not the same: insights into the evolutionary history of paralogous sex-determining genes of the dwarf honey bee Apis florea. Heredity 116:12–22
    CAS  PubMed  Google Scholar 

    Chapuis MP, Plantamp C, Streiff R, Blondin L, Piou C (2015) Microsatellite evolutionary rate and pattern in Schistocerca gregaria inferred from direct observation of germline mutations. Mol Ecol 24:6107–6119
    CAS  PubMed  Google Scholar 

    Chookajorn T, Kachroo A, Ripoll DR, Clark AG, Nasrallah JB (2004) Specificity determinants and diversification of the Brassica self-incompatibility pollen ligand. Proc Natl Acad Sci USA 101:911–917
    CAS  PubMed  Google Scholar 

    Cho S, Huang ZY, Green DR, Smith DR, Zhang J (2006) Evolution of the complementary sex-determination gene of honey bees: balancing selection and trans-species polymorphisms. Genome Res 16:1366–1375
    CAS  PubMed  PubMed Central  Google Scholar 

    Clarke BC (1979) The evolution of genetic diversity. Proc R Soc B 205:453–474
    CAS  Google Scholar 

    Ding G, Xu H, Oldroyd BP, Gloag RS (2017) Extreme polyandry aids the establishment of invasive populations of a social insect. Heredity 119:381–387
    CAS  PubMed  PubMed Central  Google Scholar 

    Ellegren H (2004) Microsatellites: simple sequences with complex evolution. Nat Rev Genet 5:435–445
    CAS  PubMed  Google Scholar 

    Fay JC, Wu CI (2000) Hitchhiking under positive Darwinian selection. Genetics 155:1405–1413
    CAS  PubMed  PubMed Central  Google Scholar 

    Fu YX (1997) Statistical tests of neutrality of mutations against population growth, hitchhiking and background selection. Genetics 147:915–925
    CAS  PubMed  PubMed Central  Google Scholar 

    Glémin S, Gaude T, Guillemin ML, Lourmas M, Olivieri I, Mignot A (2005) Balancing selection in the wild: testing population genetics theory of self-incompatibility in the rare species Brassica insularis. Genetics 171:279–289
    PubMed  PubMed Central  Google Scholar 

    Gervais CE, Castric V, Ressayre A, Billiard S (2011) Origin and diversification dynamics of self-incompatibility haplotypes. Genetics 188:625–636
    CAS  PubMed  PubMed Central  Google Scholar 

    Gibbs MJ, Armstrong JS, Gibbs AJ (2000) Sister-scanning: a Monte Carlo procedure for assessing signals in recombinant sequences. Bioinformatics 16:573–582
    CAS  PubMed  Google Scholar 

    Gloag R, Christie JR, Ding G, Stephens R, Buchmann G, Oldroyd BP (2019) Workers’ sons rescue genetic diversity at the sex locus in an invasive honey bee population. Mol Ecol 28:1585–1592
    PubMed  Google Scholar 

    Gloag R, Ding G, Christie JR, Buchmann G, Beekman M, Oldroyd BP (2017) An invasive social insect overcomes genetic load at the sex locus. Nat Ecol Evol 1:1–6
    Google Scholar 

    Hasselmann M, Beye M (2004) Signatures of selection among sex-determining alleles of the honey bee. Proc Natl Acad Sci USA 101:4888–4893
    CAS  PubMed  Google Scholar 

    Hasselmann M, Vekemans X, Pflugfelder J, Koeniger N, Koeniger G, Tingek S et al. (2008) Evidence for convergent nucleotide evolution and high allelic turnover rates at the complementary sex determiner gene of Western and Asian honeybees. Mol Biol Evol 25:696–708
    CAS  PubMed  Google Scholar 

    Hedrick PW (2005) A standardized genetic differentiation measure. Evolution 59:1633–1638
    CAS  PubMed  Google Scholar 

    Koch V, Nissen I, Schmitt BD, Beye M (2014) Independent evolutionary origin of fem paralogous genes and complementary sex determination in hymenopteran insects. PLoS ONE 9:e91883
    PubMed  PubMed Central  Google Scholar 

    Koetz AH (2013) Ecology, behaviour and control of Apis cerana with a focus on relevance to the Australian incursion. Insects 4:558–592
    PubMed  PubMed Central  Google Scholar 

    Kosakovsky Pond SL, Posada D, Gravenor MB, Woelk CH, Frost SD (2006) Automated phylogenetic detection of recombination using a genetic algorithm. Mol Biol Evol 23:1891–1901
    PubMed  Google Scholar 

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

    Kurbalija Novičić Z, Sayadi A, Jelić M, Arnqvist G (2020) Negative frequency dependent selection contributes to the maintenance of a global polymorphism in mitochondrial DNA. BMC Evol Biol 20:20. https://doi.org/10.1186/s12862-020-1581-2
    CAS  Article  PubMed  PubMed Central  Google Scholar 

    Lechner S, Ferretti L, Schöning C, Kinuthia W, Willemsen D, Hasselmann M (2014) Nucleotide variability at its limit? Insights into the number and evolutionary dynamics of the sex-determining specificities of the honey bee Apis mellifera. Mol Biol Evol 31:272–287
    CAS  PubMed  Google Scholar 

    Levin BR (1988) Frequency dependent selection in bacterial populations. Philos Trans R Soc Lond B 319:459–472
    CAS  Google Scholar 

    Li Y-C, Korol AB, Fahima T, Nevo E (2004) Microsatellites within genes: structure, function, and evolution. Mol Biol Evol 21:991–1007
    CAS  PubMed  Google Scholar 

    Lynch M (2015) Feedforward loop for diversity. Nature 523:414–416
    CAS  PubMed  PubMed Central  Google Scholar 

    Mackensen O (1951) Viability and sex determination in the honeybee (Apis mellifera L.). Genetics 36:500–509
    CAS  PubMed  PubMed Central  Google Scholar 

    Martin DP, Murrell B, Golden M, Khoosal A, Muhire B (2015) RDP4: Detection and analysis of recombination patterns in virus genomes. Virus Evol 1:vev003
    PubMed  PubMed Central  Google Scholar 

    Martin DP, Posada D, Crandall KA, Williamson C (2005) A modified bootscan algorithm for automated identification of recombinant sequences and recombination breakpoints. AIDS Res Hum Retroviruses 21:98–102
    CAS  PubMed  Google Scholar 

    May G, Matzke E (1995) Recombination and variation at the A mating-type of Coprinus cinereus. Mol Biol Evol 12:794–802
    CAS  Google Scholar 

    Moritz RFA, Haddad N, Bataieneh A, Shalmon B, Hefetz A (2010) Invasion of the dwarf honeybee Apis florea into the near East. Biol Invasions 12:1093–1099
    Google Scholar 

    Moritz RFA, Härtel S, Neumann P (2005) Global invasions of the western honeybee (Apis mellifera) and the consequences for biodiversity. Écoscience 12:289–301
    Google Scholar 

    Muirhead CA (2001) Consequences of population structure on genes under balancing selection. Evolution 55:1532–1541
    CAS  PubMed  Google Scholar 

    Muirhead CA, Glass NL, Slatkin M (2002) Multilocus self-recognition systems in fungi as a cause of trans-species polymorphism. Genetics 161:633–641
    CAS  PubMed  PubMed Central  Google Scholar 

    Nasrallah JB (1997) Evolution of the Brassica self-incompatibility locus: a look into S-locus gene polymorphisms. Proc Natl Acad Sci USA 94:9516–9519
    CAS  PubMed  Google Scholar 

    Nasrallah JB, Kao TH, Chen CH, Goldberg ML, Nasrallah ME (1987) Amino-acid sequence of glycoproteins encoded by three alleles of the S locus of Brassica oleracea. Nature 326:617–619
    CAS  Google Scholar 

    Oldroyd BP, Wongsiri S (2006) Asian honey bees. Biology, conservation and human interactions. Harvard University Press, Cambridge
    Google Scholar 

    Posada D, Crandall KA (2001) Evaluation of methods for detecting recombination from DNA sequences: computer simulations. Proc Natl Acad Sci USA 98:13757–13762
    CAS  PubMed  Google Scholar 

    Radloff SE, Hepburn C, Hepburn HR, Fuchs S, Hadisoesilo S, Tan K et al. (2010) Population structure and classification of Apis cerana. Apidologie 41:589–601
    Google Scholar 

    Raper JR (1966) Sexuality in fungi. (Book reviews: genetics of sexuality in higher fungi). Science 154:758
    Google Scholar 

    Richman AD, Kohn JR (1999) Self-incompatibility alleles from Physalis: implications for historical inference from balanced genetic polymorphisms. Proc Natl Acad Sci USA 96:168–172
    CAS  PubMed  Google Scholar 

    Rozas J, Ferrer-Mata A, Sánchez-DelBarrio JC, Guirao-Rico S, Librado P, Ramos-Onsins SE et al. (2017) DnaSP 6: DNA sequence polymorphism analysis of large data sets. Mol Biol Evol 34:3299–3302
    CAS  PubMed  Google Scholar 

    Schierup MH, Bechsgaard JS, Christiansen FB (2008) Selection at work in self-incompatible Arabidopsis lyrata. II. Spatial distribution of S haplotypes in Iceland. Genetics 180:1051–1059
    PubMed  PubMed Central  Google Scholar 

    Schierup MH, Vekemans X, Charlesworth D (2000) The effect of subdivision on variation at multi-allelic loci under balancing selection. Genet Res 76:51–62
    CAS  PubMed  Google Scholar 

    Smith DR (2011) Asian Honeybees and Mitochondrial DNA. In: Hepburn HR, Radloff SE (eds) Honeybees of Asia. Springer-Verlag Berlin, Heidelberg, p 69–93
    Google Scholar 

    Solignac M, Vautrin D, Loiseau A, Mougel F, Baudry E, Estoup A et al. (2003) Five hundred and fifty microsatellite markers for the study of the honeybee (Apis mellifera L.) genome. Mol Ecol Notes 3:307–311
    CAS  Google Scholar 

    Tajima F (1989) The effect of change in population size on DNA polymorphism. Genetics 123:597–601
    CAS  PubMed  PubMed Central  Google Scholar 

    Takahashi J, Shimizu S, Koyama S, Kimura K, Shimizu I, Yoshida T (2009) Variable microsatellite loci isolated from the Asian honeybee, Apis cerana (Hymenoptera; Apidae). Mol Ecol Resour 9:819–821
    CAS  PubMed  Google Scholar 

    Takahata N, Nei M (1990) Allelic genealogy under overdominant and frequency-dependent selection and polymorphism of major histocompatibility complex loci. Genetics 124:967–978
    CAS  PubMed  PubMed Central  Google Scholar 

    Wallberg A, Han F, Wellhagen G, Dahle B, Kawata M, Haddad N et al. (2014) A worldwide survey of genome sequence variation provides insight into the evolutionary history of the honeybee Apis mellifera. Nat Genet 46:1081–1088
    CAS  PubMed  Google Scholar 

    Walsh PS, Metzger DA, Higuchi R (1991) Chelex 100 as amedium for simple extraction of DNA for PCR-based typing from forensic material. Biotechniques 10:506–513
    CAS  PubMed  Google Scholar 

    Wang J (2015) Does GST underestimate genetic differentiation from marker data? Mol Ecol 24:3546–3558
    CAS  PubMed  Google Scholar 

    Weaver S, Shank SD, Spielman SJ, Li M, Muse SV, Kosakovsky Pond SL (2018) Datamonkey 2.0: a modern web application for characterizing selective and other evolutionary processes. Mol Biol Evol 35:773–777
    CAS  PubMed  PubMed Central  Google Scholar 

    Woyke J (1963) What happens to diploid drone larvae in honeybee colony? J Apic Res 2:73–75
    Google Scholar 

    Wright S (1939) The distribution of self-sterility alleles in populations. Genetics 24:538–552
    CAS  PubMed  PubMed Central  Google Scholar 

    Wu Q, Han TS, Chen X, Chen JF, Zou YP, Li ZW et al. (2017) Long-term balancing selection contributes to adaptation in Arabidopsis and its relatives. Genome Biol 18:217
    PubMed  PubMed Central  Google Scholar 

    Yang S, Wang L, Huang J, Zang X, Yuan Y, Chen JQ et al. (2015) Parent-progeny sequencing indicates higher mutation rates in heterozygotes. Nature 523:463–467
    CAS  PubMed  Google Scholar 

    Yokoyama S, Nei M (1979) Population dynamics of sex-determining alleles in honey bees and self-incompatibility alleles in plants. Genetics 91:609–626
    CAS  PubMed  PubMed Central  Google Scholar 

    Zareba J, Blazej P, Laszkiewicz A, Sniezewski L, Majkowski M, Janik S et al. (2017) Uneven distribution of complementary sex determiner (csd) alleles in Apis mellifera population. Sci Rep. 7:2317
    PubMed  PubMed Central  Google Scholar  More

  • in

    Owls’ hoards rot

    For predators, climate change-induced shifts in prey numbers, behaviours and spatial or temporal locations can be a major threat to food security. For predators that hoard prey to ensure survival through harsh winters, climate variation can have a doubled effect — influencing both food capture and store stability. Although northern latitude autumn and winter temperatures have increased dramatically in recent years, the effects of climate on foraging and storing throughout winter remain understudied.

    Credit: Szymon Bartosz / Alamy Stock Photo

    Giulia Masoero at the University of Turku, Finland, and colleagues analysed the impact of climate on Eurasian pygmy owl (Glaucidium passerinum) food-hoarding behaviour across 16 years. They found increased freeze–thaw frequency, lower winter precipitation and deeper snow cover were linked to greater hoard consumption. Higher autumn precipitation and an early hoarding start led to food rot, which reduced female owl recapture (indicating death or emigration).

    Although owls delayed hoarding in autumns with fewer freeze–thaw events, suggesting some potential for climate change adaptation, the study indicates that altered climates can decrease owl overwinter survival, which may in turn have vast impacts on the boreal food web.

    Author information

    Affiliations

    Nature Climate Change
    Tegan Armarego-Marriott

    Authors
    Tegan Armarego-Marriott

    Corresponding author
    Correspondence to Tegan Armarego-Marriott.

    Rights and permissions

    About this article

    Cite this article
    Armarego-Marriott, T. Owls’ hoards rot. Nat. Clim. Chang. (2020). https://doi.org/10.1038/s41558-020-0903-0
    Download citation More

  • in

    Interpreting ancient food practices: stable isotope and molecular analyses of visible and absorbed residues from a year-long cooking experiment

    1.
    Hastorf, C. A. The Social Archaeology of Food: Thinking About Eating From Prehistory to the Present (Cambridge University Press, Cambridge, 2017).
    Google Scholar 
    2.
    Lévi-Strauss, C. The culinary triangle. Partisan Review33, 586–595 (1966).
    Google Scholar 

    3.
    Wrangham, R. W. Catching Fire: How Cooking Made Us Human (Basic Books Inc, New York, 2009).
    Google Scholar 

    4.
    Roffet-Salque, M. et al. From the inside out: upscaling organic residue analyses of archaeological ceramics. J. Archaeol. Sci.16, 627–640 (2017).
    Google Scholar 

    5.
    Appadurai, A. Gastro-Politics in Hindu South Asia. Am. Ethnol.8, 494–511 (1981).
    Google Scholar 

    6.
    Douglas, M. Deciphering a meal. In Food and Culture: A Reader (eds Counihan, C. & Van Esterik, P.) 44–53 (Routledge, New York, 2008).
    Google Scholar 

    7.
    Twiss, K. The Archaeology of Food and Identity (Center for Archaeological Investigations, Southern Illinois University at Carbondale, Carbondale, 2007).
    Google Scholar 

    8.
    Rozin, P. Psychobiological perspectives on food preferences and avoidances. In Food and Evolution (eds Harris, M. & Ross, E. B.) 181–205 (Temple University Press, Philadelphia, 1987).
    Google Scholar 

    9.
    Condamin, J., Formenti, F., Metais, M. O., Michel, M. & Blond, P. The application of gas chromatography to the tracing of oil in ancient amphorae. Archaeometry18, 195–201 (1976).
    Google Scholar 

    10.
    Passi, S., Rothschild-Boros, M. C., Fasella, P., Nazzaro-Porro, M. & Whitehouse, D. An application of high performance liquid chromatography to analysis of lipids in archaeological samples. J. Lipid Res.22, 778–784 (1981).
    PubMed  CAS  Google Scholar 

    11.
    Hastorf, C. A. & DeNiro, M. J. Reconstruction of prehistoric plant production and cooking practices by a new isotopic method. Nature315, 489–491 (1985).
    ADS  CAS  Google Scholar 

    12.
    Patrick, M., de Koning, A. J. & Smith, A. B. Gas liquid chromatographic analysis of fatty acids in food residues from ceramics found in the Southwestern Cape, South Africa. Archaeometry27, 231–236 (1985).
    CAS  Google Scholar 

    13.
    Evershed, R. P., Heron, C. & Goad, L. J. Epicuticular wax components preserved in potsherds as chemical indicators of leafy vegetables in ancient diets. Antiquity65, 540–544 (1991).
    Google Scholar 

    14.
    Morton, J. D. & Schwarcz, H. P. Palaeodietary implications from stable isotopic analysis of residues on prehistoric Ontario ceramics. J. Archaeol. Sci.31, 503–517 (2004).
    Google Scholar 

    15.
    Heron, C., Evershed, R. P. & Goad, L. J. Effects of migration of soil lipids on organic residues associated with buried potsherds. J. Archaeol. Sci.18, 641–659 (1991).
    Google Scholar 

    16.
    Dudd, S. N., Evershed, R. P. & Gibson, A. M. Evidence for varying patterns of exploitation of animal products in different prehistoric pottery traditions based on lipids preserved in surface and absorbed residues. J. Archaeol. Sci.26, 1473–1482 (1999).
    Google Scholar 

    17.
    Hart, J. P., Lovis, W. A., Schulenberg, J. K. & Urquhart, G. R. Paleodietary implications from stable carbon isotope analysis of experimental cooking residues. J. Archaeol. Sci.34, 804–813 (2007).
    Google Scholar 

    18.
    Hansel, F. A., Copley, M. S., Madureira, L. A. S. & Evershed, R. P. Thermally produced ω-(o-alkylphenyl)alkanoic acids provide evidence for the processing of marine products in archaeological pottery vessels. Tetrahedron Lett.45, 2999–3002 (2004).
    CAS  Google Scholar 

    19.
    Hansel, F. A. & Evershed, R. P. Formation of dihydroxy acids from Z-monounsaturated alkenoic acids and their use as biomarkers for the processing of marine commodities in archaeological pottery vessels. Tetrahedron Lett.50, 5562–5564 (2009).
    CAS  Google Scholar 

    20.
    Evershed, R. P. Organic residue analysis in archaeology: the archaeological biomarker revolution. Archaeometry50, 895–924 (2008).
    CAS  Google Scholar 

    21.
    Copley, M. S., Hansel, F. A., Sadr, K. & Evershed, R. P. Organic residue evidence for the processing of marine animal products in pottery vessels from the pre-colonial archaeological site of Kasteelberg D east, South Africa : research article. S. Afr. J. Sci.100, 279–283 (2004).
    CAS  Google Scholar 

    22.
    DeNiro, M. J. & Epstein, S. Influence of diet on the distribution of carbon isotopes in animals. Geochim. Cosmochim. Acta42, 495–506 (1978).
    ADS  CAS  Google Scholar 

    23.
    DeNiro, M. J. & Epstein, S. Influence of diet on the distribution of nitrogen isotopes in animals. Geochim. Cosmochim. Acta45, 341–351 (1981).
    ADS  CAS  Google Scholar 

    24.
    Heron, C. & Craig, O. E. Aquatic resources in foodcrusts: identification and implication. Radiocarbon57, 707–719 (2015).
    Google Scholar 

    25.
    Cramp, L. J. E. et al. Immediate replacement of fishing with dairying by the earliest farmers of the northeast Atlantic archipelagos. Proc. R. Soc. B281, 20132372 (2014).
    PubMed  Google Scholar 

    26.
    Evershed, R. P. Biomolecular archaeology and lipids. World Archaeol.25, 74–93 (1993).
    PubMed  CAS  Google Scholar 

    27.
    Evershed, R. P., Heron, C., Charters, S. & Goad, L. J. The survival of food residues: new methods of analysis, interpretation and application. Proc. Br. Acad.77, 187–208 (1992).
    Google Scholar 

    28.
    Copley, M. S. et al. Direct chemical evidence for widespread dairying in prehistoric Britain. PNAS100, 1524–1529 (2003).
    ADS  PubMed  CAS  Google Scholar 

    29.
    Evershed, R. P., Arnot, K. I., Collister, J., Eglinton, G. & Charters, S. Application of isotope ratio monitoring gas chromatography–mass spectrometry to the analysis of organic residues of archaeological origin. Analyst119, 909–914 (1994).
    ADS  CAS  Google Scholar 

    30.
    Evershed, R. P. et al. Lipids as carriers of anthropogenic signals from prehistory. Philos. Trans. R. Soc. Lond. B354, 19–31 (1999).
    CAS  Google Scholar 

    31.
    Dunne, J., Mercuri, A. M., Evershed, R. P., Bruni, S. & di Lernia, S. Earliest direct evidence of plant processing in prehistoric Saharan pottery. Nat. Plants3, 1–6 (2016).
    Google Scholar 

    32.
    Roffet-Salque, M. et al. Widespread exploitation of the honeybee by early Neolithic farmers. Nature527, 226–230 (2015).
    ADS  PubMed  CAS  Google Scholar 

    33.
    Hammann, S. & Cramp, L. J. E. Towards the detection of dietary cereal processing through absorbed lipid biomarkers in archaeological pottery. J. Archaeol. Sci.93, 74–81 (2018).
    CAS  Google Scholar 

    34.
    Evershed, R. P., Copley, M. S., Dickson, L. & Hansel, F. A. Experimental evidence for the processing of marine animal products and other commodities containing polyunsaturated fatty acids in pottery vessels. Archaeometry50, 101–113 (2008).
    CAS  Google Scholar 

    35.
    Craig, O. E. et al. Molecular and isotopic demonstration of the processing of aquatic products in northern European prehistoric pottery. Archaeometry49, 135–152 (2007).
    CAS  Google Scholar 

    36.
    Craig, O. E. et al. Earliest evidence for the use of pottery. Nature496, 351–354 (2013).
    ADS  PubMed  CAS  Google Scholar 

    37.
    Robson, H. K. et al. Diet, cuisine and consumption practices of the first farmers in the southeastern Baltic. Archaeol. Anthropol. Sci.11, 4011–4024 (2019).
    PubMed  PubMed Central  Google Scholar 

    38.
    Oras, E. et al. The adoption of pottery by north-east European hunter-gatherers: evidence from lipid residue analysis. J. Archaeol. Sci.78, 112–119 (2017).
    CAS  Google Scholar 

    39.
    Admiraal, M., Lucquin, A., von Tersch, M., Jordan, P. D. & Craig, O. E. Investigating the function of prehistoric stone bowls and griddle stones in the Aleutian Islands by lipid residue analysis. Quatern. Res.91, 1003–1015 (2019).
    ADS  Google Scholar 

    40.
    Spangenberg, J. E., Jacomet, S. & Schibler, J. Chemical analyses of organic residues in archaeological pottery from Arbon Bleiche 3, Switzerland—evidence for dairying in the late Neolithic. J. Archaeol. Sci.33, 1–13 (2006).
    Google Scholar 

    41.
    Guiry, E. et al. Differentiating salmonid migratory ecotypes through stable isotope analysis of collagen: archaeological and ecological applications. PLoS ONE15, e0232180 (2020).
    PubMed  PubMed Central  CAS  Google Scholar 

    42.
    Mukherjee, A. J., Gibson, A. M. & Evershed, R. P. Trends in pig product processing at British Neolithic Grooved Ware sites traced through organic residues in potsherds. J. Archaeol. Sci.35, 2059–2073 (2008).
    Google Scholar 

    43.
    Shoda, S., Lucquin, A., Ahn, J., Hwang, C. & Craig, O. E. Pottery use by early Holocene hunter-gatherers of the Korean peninsula closely linked with the exploitation of marine resources. Quatern. Sci. Rev.170, 164–173 (2017).
    ADS  Google Scholar 

    44.
    Evershed, R. P. Experimental approaches to the interpretation of absorbed organic residues in archaeological ceramics. World Archaeol.40, 26–47 (2008).
    Google Scholar 

    45.
    Grace, R. Review article use-wear analysis: the state of the art. Archaeometry38, 209–229 (1996).
    Google Scholar 

    46.
    Oudemans, T. F. Applying organic residue analysis in ceramic studies in archaeology: a functional approach. Leiden J. Pottery Stud.23, 5–20 (2007).
    Google Scholar 

    47.
    Skibo, J. M. Pottery use-alteration analysis. In Use-Wear and Residue Analysis in Archaeology (eds Marreiros, J. M. et al.) 189–198 (Springer International Publishing, New York, 2015).
    Google Scholar 

    48.
    Marino, B. D. & Deniro, M. J. Isotopic analysis of archaeobotanicals to reconstruct past climates: effects of activities associated with food preparation on carbon, hydrogen and oxygen isotope ratios of plant cellulose. J. Archaeol. Sci.14, 537–548 (1987).
    Google Scholar 

    49.
    Lovis, W. A., Urquhart, G. R., Raviele, M. E. & Hart, J. P. Hardwood ash nixtamalization may lead to false negatives for the presence of maize by depleting bulk δ13C in carbonized residues. J. Archaeol. Sci.38, 2726–2730 (2011).
    Google Scholar 

    50.
    Hart, J. P., Urquhart, G. R., Feranec, R. S. & Lovis, W. A. Non-linear relationship between bulk δ13C and percent maize in carbonized cooking residues and the potential of false-negatives in detecting maize. J. Archaeol. Sci.36, 2206–2212 (2009).
    Google Scholar 

    51.
    Warinner, C. & Tuross, N. Alkaline cooking and stable isotope tissue-diet spacing in swine: archaeological implications. J. Archaeol. Sci.36, 1690–1697 (2009).
    Google Scholar 

    52.
    Warinner, C. Life and Death at Teposcolula Yucundaa: Mortuary, Archaeogenetic, and Isotopic Investigations of the Early Colonial Period in Mexico (Harvard University, New York, 2010).
    Google Scholar 

    53.
    Renson, V. et al. Origin and diet of inhabitants of the Pacific Coast of Southern Mexico during the Classic Period—Sr, C and N isotopes. J. Archaeol. Sci. Rep.27, 101981 (2019).
    Google Scholar 

    54.
    Szpak, P., Millaire, J.-F., White, C. D. & Longstaffe, F. J. Influence of seabird guano and camelid dung fertilization on the nitrogen isotopic composition of field-grown maize (Zea mays). J. Archaeol. Sci.39, 3721–3740 (2012).
    CAS  Google Scholar 

    55.
    Wang, C. et al. Aridity threshold in controlling ecosystem nitrogen cycling in arid and semi-arid grasslands. Nat. Commun.5, 1–8 (2014).
    ADS  Google Scholar 

    56.
    Yoneyama, T., Ito, O. & Engelaar, W. M. H. G. Uptake, metabolism and distribution of nitrogen in crop plants traced by enriched and natural 15N: Progress over the last 30 years. Phytochem. Rev.2, 121–132 (2003).
    CAS  Google Scholar 

    57.
    Bocherens, H. & Drucker, D. Trophic level isotopic enrichment of carbon and nitrogen in bone collagen: case studies from recent and ancient terrestrial ecosystems. Int. J. Osteoarchaeol.13, 46–53 (2003).
    Google Scholar 

    58.
    Minagawa, M. & Wada, E. Stepwise enrichment of 15N along food chains: further evidence and the relation between d15N and animal age. Geochim. Cosmochim. Acta48, 1135–1140 (1984).
    ADS  CAS  Google Scholar 

    59.
    DeNiro, M. J. & Hastorf, C. A. Alteration of 15N14N and 13C12C ratios of plant matter during the initial stages of diagenesis: studies utilizing archaeological specimens from Peru. Geochim. Cosmochim. Acta49, 97–115 (1985).
    ADS  CAS  Google Scholar 

    60.
    Fraser, R. A. et al. Assessing natural variation and the effects of charring, burial and pre-treatment on the stable carbon and nitrogen isotope values of archaeobotanical cereals and pulses. J. Archaeol. Sci.40, 4754–4766 (2013).
    CAS  Google Scholar 

    61.
    Phillips, D. L., & Gregg, J. W. Source partitioning using stable isotopes: coping with too many sources. Oecologia136, 261–269 (2003).
    ADS  PubMed  Google Scholar 

    62.
    Drieu, L. et al. Influence of porosity on lipid preservation in the wall of archaeological pottery. Archaeometry61, 1081–1096 (2019).
    CAS  Google Scholar 

    63.
    Charters, S. Chemical Analysis of Absorbed Lipids and Laboratory Simulation Experiments to Interpret Archaeological Pottery Vessel Contents and Use (University of Bristol, Bristol, 1996).
    Google Scholar 

    64.
    Bogaard, A., Heaton, T. H. E., Poulton, P. & Merbach, I. The impact of manuring on nitrogen isotope ratios in cereals: archaeological implications for reconstruction of diet and crop management practices. J. Archaeol. Sci.34, 335–343 (2007).
    Google Scholar 

    65.
    Styring, A. K. et al. The effect of charring and burial on the biochemical composition of cereal grains: investigating the integrity of archaeological plant material. J. Archaeol. Sci.40, 4767–4779 (2013).
    CAS  Google Scholar 

    66.
    Szpak, P. & Chiou, K. L. A comparison of nitrogen isotope compositions of charred and desiccated botanical remains from northern Peru. Veget Hist Archaeobot https://doi.org/10.1007/s00334-019-00761-2 (2019).
    Article  Google Scholar 

    67.
    Casanova, E. et al. Accurate compound-specific 14 C dating of archaeological pottery vessels. Nature580, 506–510 (2020).
    ADS  PubMed  CAS  Google Scholar 

    68.
    Boivin, N. L. et al. Ecological consequences of human niche construction: examining long-term anthropogenic shaping of global species distributions. Proc. Natl. Acad. Sci. USA113, 6388–6396 (2016).
    PubMed  CAS  Google Scholar 

    69.
    Ceballos, G. et al. Accelerated modern human–induced species losses: Entering the sixth mass extinction. Sci. Adv.1, e1400253 (2015).
    ADS  PubMed  PubMed Central  Google Scholar 

    70.
    Foley, S. F. et al. The Palaeoanthropocene: the beginnings of anthropogenic environmental change. Anthropocene3, 83–88 (2013).
    Google Scholar 

    71.
    Tilman, D. & Lehman, C. Human-caused environmental change: Impacts on plant diversity and evolution. PNAS98, 5433–5440 (2001).
    ADS  PubMed  CAS  Google Scholar 

    72.
    Wilk, R. Consumption, human needs, and global environmental change. Glob. Environ. Change12, 5–13 (2002).
    Google Scholar 

    73.
    Correa-Ascencio, M. & Evershed, R. P. High throughput screening of organic residues in archaeological potsherds using direct acidified methanol extraction. Anal. Methods6, 1330–1340 (2014).
    CAS  Google Scholar  More

  • in

    Quantifying net loss of global mangrove carbon stocks from 20 years of land cover change

    Overall simulation approach
    We estimated net changes in soil organic, above-ground and below-ground living biomass carbon stocks due to mangrove-related LCC that occurred globally between 1996 and 2016, for the year 2016. We did not estimate carbon sequestration because while sequestration rates in mangrove forests are higher than in many other ecosystems, there is only limited capacity for this process to substantially impact global carbon fluxes due to the small area of mangrove forest7,32,39. In addition, global high-resolution maps of mangrove carbon sequestration rates are not available.
    Uncertainty in estimates of the area of mangrove loss and gain, carbon stock density, date of deforestation or forestation, proportional carbon stock degradation due to deforestation, and proportional carbon stock accumulation due to forestation, was carried forward using a bootstrap simulation method, through which 1000 replications were used to generate median estimates and 95% confidence intervals. A summary of the sources of data, simulation parameters, and modelling of parameter variability is provided in Supplementary Table 1. For each bootstrap iteration, the area of mangrove gain and loss, carbon stock density, rt, and at values were simulated within each patch of mangrove gain or loss. The D, F, Drt, and Fat values for each patch were then calculated. The median value of the 1000 replicates was used as the estimate. The 2.5 and 97.5 percentiles of the 1000 simulation estimates of change in carbon stock were calculated, thus corresponding to bootstrap 95% confidence intervals for each of the estimated losses or gains in carbon stock within each patch. A comparable bootstrap method was applied to simulate carbon stocks in 1996 for each patch of mangrove present at that time.
    In addition to the sources of uncertainty incorporated in the bootstrap simulation, we made several key methodological decisions that could be expected to impact the conclusions of the study. We conducted four sensitivity analyses to quantify the impacts of such decisions. The sensitivity analyses were conducted only for the region of Southeast Asia, because one sensitivity analysis required detailed and spatially explicit data on the replacement land uses following mangrove deforestation, which are only available with the necessary categorisation for Southeast Asia5. Methodological details and results of the sensitivity analyses are included in Supplementary Methods 1.
    Mangrove areal extent
    We used the Global Mangrove Watch (GMW) datasets of mangrove cover to quantify deforestation and forestation14,15,16,17. While other global mangrove map products exist for specific years42,43, GMW provides maps of extent from multiple years, allowing temporal comparison. The mangrove extent in 1996 and 2016 was mapped using the GMW data products, which are derived from ALOS PALSAR and Landsat satellite-borne sensor data14,15,16. Mangrove-related LCC was defined either as a conversion from mangrove to another form of land or water cover between the 1996 and 2016 datasets (deforestation), or vice versa (forestation). Any change in mangrove cover that was reversed before the end of the study period was therefore not captured. Areas of overlapping and nonoverlapping mangrove extent were compared between these dates to quantify mangrove present in 1996 that was not present in 2016, mangrove present in 2016 that was not previously present in 1996, and areas of no change in mangrove cover between 1996 and 2016. For each patch of mangrove cover in 1996, gain, and loss, the area was calculated under the Eckert VI equal-area projection. This projection was used to calculate the area of mangrove and mangrove change polygons only, and all other analyses were conducted using the World Geodetic System (WGS) 1984 projection17.
    The GMW mapping of mangrove forest has an error rate15, leading to quantifiable uncertainty over the presence of mangroves at each location in 1996 and 2016. Accuracy statistics have not been published for each year in the GMW dataset17, so we assumed that all years had an identical accuracy to the best-documented year, which is 201015. Published error rates correspond to individual pixels in the original GMW dataset, so we modelled uncertainty at this spatial scale. For each pixel of recorded mangrove gain or loss, there is a probability that it is an erroneous, false positive example of gain or loss. For each pixel of mangrove or non-mangrove that is recorded in GMW as being the same in 2016 as in 1996, there is a similar probability of error—a false-negative case of gain or loss. While false positive gain and loss errors can be quantified fully based on the available information, we were not able to incorporate false negative errors in the simulation (Supplementary Methods 2). However, false-negative errors can be expected to impact estimates of forestation and deforestation area almost equally, while false-positive errors are biased toward a greater effect on estimates of deforestation (Supplementary Methods 2). For these reasons we incorporated only false positive classification uncertainty into the bootstrap simulation and generation of carbon stock change confidence intervals. Uncertainty in the areal extent of mangroves was not incorporated into the bootstrap simulation for the estimate of carbon stocks in 1996. As the uncertainty estimation for areal extent change does not include false negative classification errors, we do not report confidence intervals for area change statistics, and report only the median estimates from the bootstrap replicates.
    For areas of mangrove loss and gain, we incorporated the probability of false positive recording of loss and gain into the simulation. For each pixel within each patch of mangrove gain, we simulated whether it was actually not mangrove in 1996 according to the misclassification error rate for non-mangrove (Supplementary Table 2), and whether it was truly mangrove in 2016 according to the misclassification error rate for mangrove (Supplementary Table 2). The simulated number of gain pixels in each patch was thus calculated as the number of pixels that were simulated to have been both not mangrove in 1996, and mangrove in 2016. Similarly, for each pixel recorded as mangrove deforestation, there is a probability that it was a false positive example of loss. For each pixel within each patch of mangrove loss, we simulated whether it was truly mangrove in 1996 according to the misclassification error rate for mangrove (Supplementary Table 2), and whether mangrove was truly absent in 2016 according to the misclassification error rate for non-mangrove (Supplementary Table 2). The simulated number of loss pixels in each mangrove patch was thus calculated as the number of pixels that were simulated to have been mangrove in 1996, and not mangrove in 2016.
    Carbon stock density
    Spatial patterns in mangrove carbon densities were quantified using previously published datasets of soil carbon to 1 m depth36, and above- and belowground tree biomass carbon37. Both datasets are derived from systematic reviews of the literature, so may be biased towards relatively high-quality mangrove forests, rather than those that have experienced some natural or anthropogenic disturbance37. The resulting maps of carbon stock density are therefore likely to represent an upper estimate for the potential carbon stock density at each location37. As the dates and resolutions of these mangrove carbon datasets differed from the GMW mangrove extent, per hectare carbon densities were extracted for each patch of mangrove extent in 1996, gain, and loss of mangrove. Where possible, we extracted mean carbon densities for the 0.05 degree grid cell (approximately 5 km) in which the centre of the mangrove patch coincided (Supplementary Table 3). Where data did not coincide at this resolution, 0.5° grid cells (approximately 50 km) were used (Supplementary Table 3). Any remaining data gaps were filled using the global mean carbon stock density (Supplementary Table 3).
    Uncertainty in the estimate of carbon stock was modelled as a normally distributed random variable, with the mean value taken as the reported carbon stock density extracted from the published map layers36,37, and the standard deviation of the distribution taken as the reported root mean squared error (RMSE) between the model predictions and validation data36,37. The RMSE for soil carbon is reported in the study as 109 Mg per hectare36. The RMSE for aboveground biomass carbon was calculated as 104.1 Mg per hectare, based on a plot of observed versus predicted values digitised from the original study37 (Supplementary Fig. 2). The soil and biomass carbon stock densities for each patch of mangrove in 1996, patch of mangrove loss, and patch of mangrove gain were simulated, and values of less than zero were replaced with zero, to avoid negative carbon densities.
    Loss and accumulation of mangrove carbon
    After mangrove deforestation, carbon is lost gradually over a period of time, with biomass carbon typically depleting more rapidly than carbon stored in soils13. We modelled temporal losses of soil carbon stocks according to a previously-published meta-analysis of the proportion of the reference carbon stock (rt) lost over time13. For losses of biomass carbon stocks, we used a meta-analysis of temporal changes in the proportion of the reference tree diameter as a proxy for biomass carbon stock, because a meta-analysis of temporal changes in biomass carbon stock was not available13. The shape of these temporal rt relationships can be observed in Supplementary Fig. 3a, b. The approximate date of mangrove deforestation was quantified by cross-referencing several dates from the GMW dataset to establish the dates of presence and absence17. We cross-referenced the dates of 1996, 2007, 2010, and 2016 to identify the dates of mangrove presence and absence at each location.
    Uncertainty in proportional losses of carbon due to mangrove deforestation was incorporated in two ways. First, there is uncertainty in the date of mangrove deforestation since the most recent observed date of presence. To model uncertainty in the date of deforestation we used a uniform distribution to select a date between the most recent date of observed mangrove presence and the oldest date of observed mangrove absence. Second, there is uncertainty in the relationship between the time since deforestation and proportion of mangrove carbon lost, quantified as the error present in the regression models (Supplementary Fig. 3a, b). We simulated the projected proportions of mangrove carbon remaining as a function of the length of time since deforestation (2016—date of deforestation), accounting for the error inherent in each linear model (Supplementary Methods 3).
    As mangrove forests grow, they typically accumulate carbon in soil and tree biomass stocks, until reaching the value held by the reference community12,44. This process can be slow, taking from 20 to more than 50 years12,44. At a given point in time before the climax community is reached, the mangrove ecosystem contains a proportion of the value held in the climax community (at). The whole-ecosystem carbon accumulation curve for afforesting mangroves was estimated using data taken from a meta-analysis of blue carbon ecosystem restoration18, that we used to estimate the proportion of the reference ecosystem carbon accumulated following restoration (Supplementary Methods 4). The shape of the temporal relationship can be observed in Supplementary Fig. 3c. We used this general relationship describing restoration of all blue carbon ecosystems, because a mangrove forestation-specific meta-analysis is not currently available. To assess the impacts of this selection on the study findings, we also conducted a sensitivity analysis using data from two case studies of mangrove soil carbon and biomass accumulation in foresting mangroves (Supplementary Methods 2). The approximate date of mangrove forestation was quantified by cross-referencing several dates from the GMW dataset to establish the dates of presence and absence14,15,16. We cross-referenced the dates of 1996, 2007, 2010, and 2016 to identify the dates of mangrove presence and absence at each location.
    Uncertainty in gains of carbon due to mangrove forestation was incorporated in two ways. First, there is uncertainty in the date of mangrove forestation since the most recent observed date of presence. To model uncertainty in the date of forestation we used a uniform distribution to select a date between the most recent date of observed mangrove absence and the oldest date of observed mangrove presence. Second, there is uncertainty in the relationship between the time since deforestation and proportion of mangrove carbon lost, quantified as the error present in the meta-analytic regression model (Supplementary Fig. 3c). We simulated the projected proportions of mangrove carbon remaining as a function of the length of time since deforestation (2016—date of deforestation), accounting for the error inherent in each linear model45 (Supplementary Methods 3).
    We estimated D, F, rt, and at for each patch of mangrove gain and loss between 1996 and 2016. These data were then used to quantify four indicators of net change in mangrove carbon stocks, to evaluate the sensitivity of estimation to the inclusion or exclusion of afforestation and remnant carbon processes. The first indicator estimated the maximum carbon stock at risk of loss due to deforestation (D), following the approach used in the most recent global estimate of potential mangrove carbon emissions9. The second indicator estimated net loss of carbon assuming 100% carbon loss and gain rates (D − F). For the third indicator, we estimated the carbon stock loss due to deforestation but accounting for remnant carbon (Drt)8. Finally, the fourth indicator estimated net changes in mangrove carbon stock between 1996 and 2016 accounting for both forestation and proportional accumulation and loss rates of carbon following LCC (Drt − Fat). For mapping of spatial variability in net gains and losses of mangrove carbon stocks, we quantified the net change in mangrove carbon stock (Fat − Drt) by summarising all patches of mangrove gain and loss with their centroids located in cells across a global grid (Figs. 2 and 3). More

  • in

    Downscaling global ocean climate models improves estimates of exposure regimes in coastal environments

    Study region
    Monterey Bay is the largest bay on the California coast, and has a total area of approximately 1,162 km2 (Fig. 1). Due to upwelling, the bay is highly productive and supports dense kelp forests, dominated by Macrocystis pyrifera, on rocky reefs that extend to approximately 20 m depth27. These forests have the capacity to regulate pH28,29,30 and sustain over 200 different species, from phytoplankton to marine mammals31.
    Temperatures are the coolest in Monterey Bay during spring and early summer due to wind-driven upwelling32, driven by equatorward winds blowing over the California coast causing offshore Ekman transport33. During the same months, filaments of water originating from Point Año Nuevo to the north are trapped within the Bay, where they warm due to solar radiation, forming a lens of warm water close to shore34,35 (Fig. 1). Furthermore, a weak cyclonic eddy is observed within the bay due to the coastal geometry32,34. Inside the bay, sea breezes and tides drive diurnal and semi-diurnal currents that can lead to significant variability in environmental conditions17,36. During the upwelling period, salinity is approximately 34, temperature ranges from 9 to 13 °C at 17 m depth7, DO varies from as low as 100 μmol kg−1 (3.2 mg L−1 for T = 13 °C, S = 34 at 17 m) to as high as 300 μmol kg−1 (9.62 mg L−1 for T = 13 °C, S = 34 at 17 m), and pH varies from 7.7 to 8.13 (Supplementary Fig. 1).
    The three primary barotropic tidal constituents in the region (M2, K1, and S2) are responsible for over 80% of the tidal amplitude observed18. In the southern region of the bay, tides are mainly responsible for the cross-shelf velocity36 and the interaction of these surface tidal currents with the steep topography create internal tides comprised of internal waves and bores occurring at tidal frequencies37,38. Internal tides are formed when currents move over steep slopes, dense waters are forced into shallower regions, and, as these waters sink back to depth, an internal wave is generated39. These internal waves have speeds8 on the order of 0.05–0.2 m s−1 and, as the coast steepens, can break forming internal bores that move upslope and bring cold, low DO, low pH waters into nearshore kelp forest ecosystems40. During upwelling, these processes drive variability in temperature, pH and DO over semi-diurnal and diurnal periods that can exceed the predicated changes in mean conditions predicted by global climate models for year 210041.
    Model description
    To better understand how tides and winds affect exposure of nearshore organisms to variability in temperature, DO, and pH under current and RCP climate scenarios, we used dynamic downscaling and developed a 2D coupled biogeochemical hydrodynamic model using ROMS42. The model domain was created based on the Monterey Bay continental shelf described in Walter et al.6 with a maximum offshore depth of approximately 80 m (Fig. 2). The biogeochemical model used is described in Fennel et al.43,44. We forced the model with representative wind (diurnal sea breeze accounting for regional winds), solar radiation, and tidal currents for the Monterey Bay region (Supplementary Figs. 2–6; Supplementary Table 1). Detail of the full model structure, including boundary and initial conditions, are included in the “Supplementary Information”.
    Figure 2

    Snapshot of model crosshore velocity and temperature (contour lines) on 8 July 2013. Positive values are onshore, negative values are offshore. Contour lines (black) show isotherms with temperature labels.

    Full size image

    We estimated depth profiles for temperature, DO, and pH for the downscaled coupled hydrodynamic-biogeochemical model. We considered a homogeneous salinity (S = 34) for the entire domain since the variation in salinity is less than 0.4 over an entire year and less than 0.05 during the upwelling season in the Monterey Bay region27. Thus, initial and boundary stratification are assumed to be controlled by temperature alone. We represent initial (IC) and boundary conditions (BC) for temperature, DO, and pH using:

    $$ {text{var}} left( {text{z}} right) = left{ {begin{array}{*{20}l} {Delta {text{var}} ;{text{D}}_{{{text{pyc}}}}^{alpha } } hfill & {{text{if}};{text{z}} ge {text{D}}_{{{text{pyc}}}} } hfill \ {Delta {text{var}} ;{text{z}}^{alpha } } hfill & {{text{otherwise}}} hfill \ end{array} } right. $$
    (1)

    where var is the identified variable (e.g. temperature), z is the depth, α is a fit coefficient for each variable determined from a least-squares fit to observational data, and Dpyc is the depth of the pycnocline. We used this method to estimate profiles of temperature, dissolved inorganic carbon (DIC), and DO for present and all downscaled future scenarios setting Dpyc to 17.5 m. The data used for the initial and boundary conditions (BC) of phytoplankton and chlorophyll profiles were taken from Schuckmann et al.45. The initial chlorophyll concentration was converted to zooplankton concentration (zoop = 0.34 × 10–3 mmol m−3) using Eq. 3 from Wiebe46, a method that has been used in other studies47,48. Detritus was initially set to zero in the entire domain. All biogeochemical variables were forced hourly at the southern boundary.
    For each scenario, we estimated depth profiles for temperature, dissolved oxygen, and DIC using Eq. (1). In order to fit Eq. (1), we obtained our estimated values of each variable near the surface, at 80 m depth on the shelf during upwelling for Present and 200 m depth for Future, respectively. We used 200 m depth from the future data set as this is the most common depth of source waters for upwelling in the region49. Present surface and bottom values of temperature, DO, total alkalinity (TA), and DIC for present scenario were based on Koweek et al.27. The mean of the 3-month period of strong upwelling (May, June, and July) was used to obtain values of temperature (surface and depth), oxygen (surface), and DIC (surface) from Representative Concentration Pathway (RCP) for the year of 2100 from the 4th report of the IPCC50. Since only surface values for DO and DIC and no values for TA were available, we estimated values for these parameters at depth.
    For DIC, we assumed that the ratio between surface and bottom values (80 m for Present and 200 m for Future) in present conditions will not change for future scenarios:

    $$frac{{DIC}_{Surf}^{Present}}{{DIC}_{Bottom}^{Present}}=frac{2073}{2280}= 0.909$$
    (2)

    Therefore, to find the bottom values for future scenarios we divided RCP surface values by this ratio:

    $$frac{{DIC}_{Surf}^{RCP8.5} }{0.909}= frac{2167 }{0.909} = 2384;mathrm{mmol;C};{mathrm{m}}^{-3}$$
    (3)

    Surface and bottom TA values were kept the same as present conditions, following Feely et al.3. For DO, bottom values for RCP 2.6 and 8.5 were approximated from Figs. 5 and 6 of Bopp et al.51. We calculated the ratios between surface and bottom DO for Present, RCP 2.6, and RCP 8.5. The ratio of surface:bottom DO was not constant across scenarios, so we approximated the ratios for RCPs 4.5 and 6.0 using linear least squares fit (Table 1; example calculation in the “Supplementary Information”). We, then applied the values from Table 1 in order to calculate α assuming Dpyc = 17.5 m. We then used Eq. (1) to generate the initial and boundary conditions (Fig. 3).
    Table 1 Values for present (empirical data) and future (global ocean models) surface and estimated bottom conditions used to fit Eq. 3 and used as boundary conditions for the downscaled model runs.
    Full size table

    Figure 3

    Initial and Boundary Conditions profiles for present and future scenarios: (a) temperature, (b) O2, (c) DIC, (d) pH.

    Full size image

    We calculated pH and Ωar using the CO2SYS52 package in MATLAB using temperature, salinity, DIC, and Total Alkalinity (TA) from the simulations at the offshore location where the bottom depth was 15 m. We assumed concentrations of phosphate and silica based on Koweek et al.27. We used dissociation constants for H2CO3 and HCO3 from Dickson and Millero53 and hydrogen sulfate ion constant (HSO-4) from Dickson54. All surface oxygen values were shifted positively 60 mmol m−3 (1.87 mg L−1 for T = 13 °C and S = 34) in order to simulate the high primary production due to kelp forests29, which is not specifically accounted for in the model. Overall, temperature at the bottom remained constant across all scenarios, with exception of RCP6.0 where temperature increased 0.2 °C (Supplementary Table 2).
    Integrated exposure
    Field observations24,55 and laboratory experiments5,56 have shown that below sub-lethal thresholds marine organisms inhabiting nearshore marine habitats in upwelling systems exhibit signs of physiological stress when exposed to elevated temperatures, low oxygen levels, or low pH waters, which is especially detrimental especially for calcifying species. Exposure of organisms to stressful temperature, DO, and pH conditions (φth where φ refers to temperature, DO, or pH) was done by subtracting the threshold value for a given organism and life stage from the model or observational data at 15 m water depth, then setting all positive values to zero for pH and O2, and all negative values to zero for temperature. Next, we estimated integrated exposure (Eint) by integrating absolute exposure over a period of a week with a window interval of 1 h:

    $$ emptyset^{prime} = emptyset – emptyset_{th} left{ {begin{array}{*{20}c} {emptyset^{prime} > 0 to emptyset^{prime} = 0;for;pH;and;O_{2} } \ {emptyset^{prime} < 0 to emptyset^{prime} = 0;for;temperature} \ end{array} } right. $$ $$ E_{int} = mathop int limits_{0}^{t} left| {emptyset^{prime}} right|dt $$ (4) Thresholds of temperature, dissolved oxygen and pH (16 °C57, 4.8525 mg L−1, and pH of 7.523 respectively) representing non-interactive negative impacts on juvenile red abalone growth were based on literature values5,23,57. Integrated exposure quantified the time and degree of stress an organism experiences, similar to the degree heating week with units of oC w, or day (oC d) measure used to estimate thermal stress on coral reefs58, and has been previously used to understand the exposure of juvenile abalone populations to similar stressors in an empirical field study57. Overall, red abalone threshold values chosen had a strong negative effect on the species5,23,57, therefore, we used Eint as a proxy for estimating the potential impact of future conditions on abalone growth and survival. Fertilization response We estimated fertilization success using results of Boch et al.25 where the fertilization response of red abalone (Haliotis rufescens) was quantified in response to multiple stressor climate conditions (high temperature, low DO, and low pH). Fertilization in abalone occurs over relatively short periods, therefore Eint would not provide an appropriate estimate in such cases. While the process of fertilization occurs over short periods, adult red abalone exhibit an extended spawning season, over which environmental conditions may vary greatly based on our modeling results. Thus, we used the equations from Boch et al.25 to examine how fertilization success over a one-month period might be affected by environmental variability, specifically the interactive effects of ph and temperature. Changes in DO did not show a strong effect on fertilization in their experiments (Fig. 4; see Supplementary Table 3 for parameter values): Figure 4 Proportional Fertilization (Prop. Fert.) as a function of pH for red abalone Eq. (5)—blue line; Eq. (6)—black line based on Boch et al.25 (Fig. 4a,c). For our study we used 15.5 °C as a transition between the two curves shown. Full size image For temperatures = 13 °C: $$ {text{Logit}};left( {% Fert.} right) = left{ {begin{array}{*{20}l} {{upbeta}_{0} + {upbeta}_{{{text{pH}}}} {text{pH}}} hfill & {{text{for}};{text{pH}} le {text{BP}}} hfill \ {left( {{upbeta}_{{{text{pH}}}} + left( {{upbeta}_{2} - {upbeta} } right)} right){text{pH } - text{ offset}}} hfill & {{text{for}};{text{pH}} > {text{BP}}} hfill \ end{array} } right. $$
    (5)

    For Temperatures = 18 °C:

    $$mathrm{Logit};(mathrm{%}Fert.)= {upbeta }_{0} + ({upbeta}_{mathrm{p}H} + {upbeta }_{mathrm{A}})mathrm{p}H + {upbeta }_{mathrm{B}}$$
    (6)

    where ({upbeta }_{0}) and ({upbeta }_{mathrm{pH}}) are intercepts, (upbeta ) and ({upbeta }_{2}) are slope segments, ({upbeta }_{mathrm{A}}) is slope of the pH-temperature interaction (pH × Temperature Group), ({upbeta }_{mathrm{B}}) is accounts for high temperature effects, and BP is the curve breaking point. Since only curves for 13 °C and 18 °C were available, we used 15.5 °C as a transition where Eq. (5) was applied for temperature less than 15.5 °C and Eq. (6) was applied for temperature greater than 15.5 °C. Since more complex interpolation schemes yielded similar results, we used this straightforward method for clarity.
    Model evaluation
    We first assessed whether the oceanographic model was able to reproduce current (observed) oceanographic conditions in Monterey Bay. The model was expected to reproduce the main dominant semi-diurnal and diurnal periods of oscillations observed in the region as well as primary production regulation of DO and pH (excluding kelp) in the biogeochemical model. We had two requisites to consider the model performance satisfactory. First, we required that the model was capable of reproducing local minima for DO. Second, the model needed to be capable of reproducing the mean and extremes for temperature and pH. In addition, we anticipated observing lower values of DO in surface waters since we did not account for the higher primary productivity observed in kelp forests20, and therefore, DO oversaturation. To evaluate the biogeochemical model, we compared chlorophyll concentration in the model to satellite-derived chlorophyll estimates for the region. Monthly mean depth averaged chlorophyll by area (mg Chl m−1) was calculated for the model and for the months of May, June, and July from Sea-Viewing Wide Field-of-View Sensor (SeaWiFS)59 for the period of 2010–2017 to the closest region with data available next to our model simulation (cross section region in Fig. 1). Chlorophyll concentrations in the model were 2.12 mg Chl m−3 compared to 4.02 mg Chl m−3 estimated from SeaWifs. Thus, modeled values were within the range observed in the satellite data over this period.
    In order to validate temporal variability in the model results, we estimated power spectra using the Thomson Multi-taper method (MTM)60. This method was chosen due to its robustness for stationary data with low variance. Power spectra allowed us to quantify the variability by frequency and confirm that the model was reproducing variability at dominant periods (M2, K1) observed in the region, as initial and boundary conditions in the model were based on regional observations. We applied the analysis over a 3-week window of upwelling for temperature, DO, and pH and compared with Booth et al.7 data (Fig. 5). Spectral analysis was used to ascertain the dominate temporal constituents of temperature, DO, and pH for all scenarios (present and RCP) between 12- and 24-h periods. The spectra were then integrated numerically to calculate the variability of temperature, DO, and pH, and the 95% confidence interval at each frequency band. Bootstrap analysis was applied to the integrated exposure calculation for each variable. In order to assure convergence in our estimates, 1,000 iterations with replacement were done using a sample size of 50% of the data. In the end, the confidence interval (CI) was calculated based on 2.5% percentile and 97.5% percentile of the distribution of the estimated means.
    Figure 5

    Time series for in situ5 and model data and power spectra of temperature (a,b), DO (c,d), and pH (e,f). Time series for the model was shifted to match tidal phase of the in situ data. cpd cycles per day.

    Full size image

    Our model results reflect current variability in temperature, DO, and pH in southern Monterey Bay (Fig. 5). In addition, mean temperature and pH values were not significantly different from in situ data. However, mean DO in our model was lower than present day averages, likely due to the lack of oxygen-producing kelp in our model61. Thus, the model accurately simulates diurnal, semi-diurnal, and higher order tidal components62. For Monterey Bay, diurnal (cycles per day (cpd) = 1) and semi-diurnal (cpd = 1.93) are the main temporal components of variability in temperature, oxygen, and pH and are well represented by our model with overlapping 95% confidence intervals (Fig. 5b,d,f). Higher frequency variability (cpd = 3) is also within the 95% confidence interval when comparing model and observations. However, frequencies occurring between peaks are not well resolved (Fig. 5b,d,f). We expect this is because we used a 2D model, and therefore, the model was unable to resolve all the physical processes occurring in the Monterey Bay. The oscillations between peaks in the observed data were likely due other coastal ocean processes such 3D circulation and ocean surface waves36 (as well as noise in the instruments). However, the variability at these periods did not have an appreciable effect on exposure calculations. Importantly, the model preserved the observed diurnal and semi-diurnal variability that is not present in global and regional scale climate models for future RCP scenarios (Fig. 5a,c,e).
    Variability at 15 m
    The model was able to simulate the main drivers of variability in temperature, DO, and pH (predominantly internal waves) for the region in study. Internal waves in the domain were seen as vertical changes in the u-component of the velocity (Fig. 2). Cross-shelf velocities ranged from − 0.05 to 0.1 m s−1, and were within the range found in other studies8,18,27. Before the arrival of the internal wave crest, isotherms were tilted downwards indicating previous downwelling. Flow in opposite directions between crests was observed during retreating of internal waves in the domain, as it has been observed in studies on internal waves with in situ data8,9.
    High variability in temperature, DO, and pH was observed in all model runs (Supplementary Figs. 7–9). Mean temperatures were 10.63 °C (SD = 0.39), 13.81 °C (SD = 0.46), 15.46 °C (SD = 0.50), 15.45 °C (SD = 0.49), 16.96 °C (SD = 0.95) for present, RCP 2.6, RCP 4.5, RCP 6.0, and RCP 8.5, respectively (Supplementary Fig. 7). Overall, mean temperature increased as expected and exhibited similar variability (standard deviation [SD]) across all RCPs except for RCP 8.5 scenario. This was likely due to increased temperature stratification where surface waters warm faster than deeper waters resulting in higher temperature variability.
    Mean dissolved oxygen values were 5.15 mg L−1 (SD = 0.74) for present, 4.80 mg L−1 (SD = 0.88) for RCP 2.6, 5.32 mg L−1 (SD = 0.77) for RCP 4.5, 5.02 mg L−1 (SD = 0.81) for RCP 6.0, and 4.93 mg L−1 (SD = 1.21) for RCP 8.5 (Supplementary Fig. 8). The SD for RCP 8.5 was almost twice that of other scenarios. This was likely due to a stronger gradient in oxygen (surface remains saturated while the values at depth are lower). RCP 2.6 had the lowest mean DO but not different than the rest and RCP 4.5 was 0.52 mg L−1 higher, though not significantly different than the present scenario. The consequence for the lowest mean DO in RCP 2.6 was related to the strength of density stratification61, and therefore, low oxygen at 15 m. Another scenario (not shown) was used where RCP 4.5 oxygen profile was applied using the density stratification from RCP 2.6 and the same low oxygen values found previously were also observed in the alternate run, supporting this mechanism.
    pH variability was also high across all model runs (Supplementary Fig. 9). The mean value for pH decreased from 7.73 (SD = 0.07) for present to 7.44 (SD = 0.12) for the RCP 8.5 scenario. RCP 8.5 again had the highest variability among all scenarios. Otherwise, the pH range was ~ 0.075 for all other scenarios. Lower pH for the most extreme scenarios (RCP 6.0 and 8.5) has also been observed in large scale models12. More

  • in

    Why locusts congregate in billion-strong swarms — and how to stop them

    EDITORIAL
    26 August 2020

    Researchers are starting to understand the behaviour of insects ravaging parts of Asia, Africa and the Middle East. This work must be furthered, funded and field-tested.

    Desert locusts have been swarming in some areas since late 2019.Credit: Fredrik Lerneryd/Getty

    On top of coronavirus, many countries are dealing with a second dangerous plague. Since the end of 2019, gigantic swarms of the desert locust Schistocerca gregaria have been devouring crops across East Africa, the Middle East and southwest Asia. It is the worst locust crisis some regions have seen for 70 years.
    The upsurge — which has been linked to unusually heavy rains and a tropical cyclone on the Arabian Peninsula — has produced devastating swarms in Kenya, Ethiopia, Somalia, Yemen and India, with many more countries under threat. At least 20 million people are at risk of losing their food supplies and livelihoods, according to the Food and Agriculture Organization (FAO) of the United Nations. Swarms typically contain between 4 billion and 8 billion locusts, and can eat in a day the equivalent of what at least 3.5 million people would consume.
    Governments and research organizations in the affected countries are fighting to control the insects, largely by spraying pesticides from planes. But it can seem like a losing battle. The swarms are being dealt with at the 11th hour: only after the juvenile insects, which are known as hoppers, gather to take flight.

    But researchers are making progress. They are starting to understand how the insects communicate; some have used data from other outbreaks to design tools to predict when and where the next ones will happen. They are calling for more real-time data to inform agricultural policies.
    All this is crucial work, but just first steps. Equally important is the need to test, improve and eventually act on these findings. The results must be turned into something practical that can be used in the fight against the desert locust.
    Chemical attraction
    One long-standing mystery is what causes the locusts to come together periodically in sky-blackening swarms. In this issue of Nature, Xiaojiao Guo and her colleagues report one answer: they identify a sweet-smelling pheromone produced by the migratory locust Locusta migratoria, a different species that also forms swarms. The researchers, at the Chinese Academy of Sciences and Hebei University, isolated 35 compounds emitted by this insect (X. Guo et al. Nature 584, 584–588; 2020). They tested a handful for their ability to attract other locusts, and found that the pheromone 4-vinylanisole (4VA) had the strongest results. The researchers also discovered that when just four or five locusts congregate, they start to produce 4VA, which then attracts others to create a swarm (see New & Views).
    The researchers identified a gene, Or35, which produces a receptor that detects the pheromone. Using CRISPR–Cas9 gene editing, they showed that locusts with a mutated Or35 were unable to detect or respond to 4VA.
    Locust forecast
    In a different study, published last month, Emily Kimathi and her colleagues created the first draft of a machine-learning algorithm designed to predict desert-locust breeding sites (E. Kimathi et al. Sci. Rep. 10, 11937; 2020). The team at three institutions in Kenya, working with the FAO, combined more than 9,000 locust records from Mauritania, Morocco and Saudi Arabia with information on rainfall, temperature and soil and sand moisture. The algorithm performed well at predicting breeding sites in all three locations.

    All these promising findings could, at least in theory, be used in complementary ways. The model could point to potential breeding sites, where an artificial pheromone might be released to attract locusts so that they can be trapped and destroyed before they breed in large numbers. But first, the findings must clearly be validated, extended and tested in the field. The machine-learning model needs to be refined. Researchers must establish whether 4VA has the same effect on the destructive desert locust as on the migratory locust and whether other signals are involved; much more work would be needed before an artificial pheromone could be created; and researchers must investigate practical issues such as how, where and when to distribute traps.
    Major locust upsurges happen infrequently — the last event was 15 years ago — and so national and international funders have not prioritized such research. That is one reason countries have not been prepared for attacks: locust surveillance, including in-country research, has been weakened by years of under-funding. This cannot be allowed to continue. It isn’t known how quickly swarms will return after the present outbreak. But countries must be prepared when they do.

    Nature 584, 497 (2020)
    doi: 10.1038/d41586-020-02453-8

    Latest on:

    Agriculture

    Developing world

    An essential round-up of science news, opinion and analysis, delivered to your inbox every weekday.

    Related Articles More

  • in

    Bimodal diel pattern in peatland ecosystem respiration rebuts uniform temperature response

    1.
    Turunen, J., Tomppo, E., Tolonen, K. & Reinikainen, A. Estimating carbon accumulation rates of undrained mires in Finland–application to boreal and subarctic regions. Holocene 12, 69–80 (2002).
    ADS  Google Scholar 
    2.
    Gorham, E. Northern Peatlands: role in the carbon cycle and probable responses to climatic warming. Ecol. Appl. 1, 182–195 (1991).
    PubMed  Google Scholar 

    3.
    Loisel, J. et al. A database and synthesis of northern peatland soil properties and Holocene carbon and nitrogen accumulation. Holocene 24, 1028–1042 (2014).
    ADS  Google Scholar 

    4.
    Yu, Z., Loisel, J., Brosseau, D. P., Beilman, D. W. & Hunt, S. J. Global peatland dynamics since the Last Glacial Maximum. Geophys. Res. Lett. 37, L13402 (2010).
    ADS  Google Scholar 

    5.
    Batjes, N. H. Total carbon and nitrogen in the soils of the world. Eur. J. Soil Sci. 65, 10–21 (2014).
    CAS  Google Scholar 

    6.
    Dorrepaal, E. et al. Carbon respiration from subsurface peat accelerated by climate warming in the subarctic. Nature 460, 616–619 (2009).
    ADS  CAS  Google Scholar 

    7.
    Nijp, J. J. et al. Rain events decrease boreal peatland net CO2 uptake through reduced light availability. Glob. Change Biol. 21, 2309–2320 (2015).
    ADS  Google Scholar 

    8.
    Petrescu, A. M. R. et al. The uncertain climate footprint of wetlands under human pressure. Proc. Natl Acad. Sci. USA. 112, 4594–4599 (2015).
    ADS  PubMed  Google Scholar 

    9.
    Wu, J. & Roulet, N. T. Climate change reduces the capacity of northern peatlands to absorb the atmospheric carbon dioxide: the different responses of bogs and fens. Glob. Biogeochem. Cycles 28, 1005–1024 (2014).
    ADS  CAS  Google Scholar 

    10.
    Wang, X. et al. Soil respiration under climate warming: differential response of heterotrophic and autotrophic respiration. Glob. Change Biol. 20, 3229–3237 (2014).
    ADS  Google Scholar 

    11.
    Mäkiranta, P. et al. Indirect regulation of heterotrophic peat soil respiration by water level via microbial community structure and temperature sensitivity. Soil Biol. Biochem. 41, 695–703 (2009).
    Google Scholar 

    12.
    Bond-Lamberty, B., Wang, C. & Gower, S. T. A global relationship between the heterotrophic and autotrophic components of soil respiration? Glob. Change Biol. 10, 1756–1766 (2004).
    ADS  Google Scholar 

    13.
    Järveoja, J., Nilsson, M. B., Gažovič, M., Crill, P. M. & Peichl, M. Partitioning of the net CO2 exchange using an automated chamber system reveals plant phenology as key control of production and respiration fluxes in a boreal peatland. Glob. Change Biol. 24, 3436–3451 (2018).
    ADS  Google Scholar 

    14.
    Lloyd, J. & Taylor, J. A. On the temperature dependence of soil respiration. Funct. Ecol. 8, 315 (1994).
    Google Scholar 

    15.
    van’t Hoff, J. H. Lectures on theoretical and physical chemistry: chemical dynamics. Part I (Edward Arnold, London, 1898).

    16.
    Arrhenius, S. Uber die Reaktionsgeschwindigkeit bei der Inversion von Rohrzucker durch Sauren. Z. f.ür. Phys. Chem. 4, 226–248 (1889).
    Google Scholar 

    17.
    Reichstein, M. et al. On the separation of net ecosystem exchange into assimilation and ecosystem respiration: review and improved algorithm. Glob. Change Biol. 11, 1424–1439 (2005).
    ADS  Google Scholar 

    18.
    Lasslop, G. et al. Separation of net ecosystem exchange into assimilation and respiration using a light response curve approach: critical issues and global evaluation. Glob. Change Biol. 16, 187–208 (2010).
    ADS  Google Scholar 

    19.
    Lasslop, G. et al. On the choice of the driving temperature for eddy-covariance carbon dioxide flux partitioning. Biogeosciences 9, 5243–5259 (2012).
    ADS  CAS  Google Scholar 

    20.
    Wohlfahrt, G. & Galvagno, M. Revisiting the choice of the driving temperature for eddy covariance CO2 flux partitioning. Agric. Meteorol. 237–238, 135–142 (2017).
    Google Scholar 

    21.
    Phillips, S. C. et al. Interannual, seasonal, and diel variation in soil respiration relative to ecosystem respiration at a wetland to upland slope at Harvard Forest. J. Geophys. Res. Biogeosciences 115, G02019 (2010).

    22.
    Savage, K., Davidson, E. A. & Tang, J. Diel patterns of autotrophic and heterotrophic respiration among phenological stages. Glob. Change Biol. 19, 1151–1159 (2013).
    ADS  CAS  Google Scholar 

    23.
    Thorne, R., Khomik, M., Hayman, E. & Arain, A. Response of soil CO2 efflux to shelterwood harvesting in a mature temperate pine forest. Forests 11, 304 (2020).
    Google Scholar 

    24.
    Carbone, M. S., Winston, G. C. & Trumbore, S. E. Soil respiration in perennial grass and shrub ecosystems: linking environmental controls with plant and microbial sources on seasonal and diel timescales. J. Geophys. Res. Biogeosciences 113, G02022 (2008).

    25.
    Keane, J. Ben & Ineson, P. Technical note: differences in the diurnal pattern of soil respiration under adjacent Miscanthus × giganteus and barley crops reveal potential flaws in accepted sampling strategies. Biogeosciences 14, 1181–1187 (2017).
    ADS  CAS  Google Scholar 

    26.
    Wutzler, T. et al. Basic and extensible post-processing of eddy covariance flux data with REddyProc. Biogeosciences 15, 5015–5030 (2018).
    ADS  CAS  Google Scholar 

    27.
    Luo, Y. & Zhou, X. Soil respiration and the environment. (Academic Press, An Imprint of Elsevier Science, London, 2006).

    28.
    Hoffmann, M. et al. Automated modeling of ecosystem CO2 fluxes based on periodic closed chamber measurements: a standardized conceptual and practical approach. Agric. Meteorol. 200, 30–45 (2015).
    Google Scholar 

    29.
    Rochette, P. & Hutchinson, G. Measurement of soil respiration in situ: chamber techniques. in Micrometeorology in agricultural systems, Agron. Monogr. 47 (ASA, CSSA and SSSA, Madison, WI 2005).

    30.
    Peichl, M. et al. A 12-year record reveals pre-growing season temperature and water table level threshold effects on the net carbon dioxide exchange in a boreal fen. Environ. Res. Lett. 9, 055006 (2014).
    ADS  Google Scholar 

    31.
    Nilsson, M. et al. Contemporary carbon accumulation in a boreal oligotrophic minerogenic mire – a significant sink after accounting for all C-fluxes. Glob. Change Biol. 14, 2317–2332 (2008).
    ADS  Google Scholar 

    32.
    Qiu, C. et al. ORCHIDEE-PEAT (revision 4596), a model for northern peatland CO2, water, and energy fluxes on daily to annual scales. Geosci. Model Dev. 11, 497–519 (2018).
    ADS  CAS  Google Scholar 

    33.
    Abdalla, M. et al. Simulation of CO2 and attribution analysis at Six European Peatland sites using the ECOSSE model. Water, Air, Soil Pollut. 225, 2182 (2014).
    ADS  Google Scholar 

    34.
    Metzger, C., Nilsson, M. B., Peichl, M. & Jansson, P.-E. Parameter interactions and sensitivity analysis for modelling carbon heat and water fluxes in a natural peatland, using CoupModel v5. Geosci. Model Dev. 9, 4313–4338 (2016).
    ADS  CAS  Google Scholar 

    35.
    Jung, M. et al. Global patterns of land-atmosphere fluxes of carbon dioxide, latent heat, and sensible heat derived from eddy covariance, satellite, and meteorological observations. J. Geophys. Res. 116, G00J07 (2011).
    Google Scholar 

    36.
    Ai, J. et al. MODIS-based estimates of global terrestrial ecosystem respiration. J. Geophys. Res. Biogeosciences 123, 326–352 (2018).
    ADS  Google Scholar 

    37.
    Xiao, J. et al. Remote sensing of the terrestrial carbon cycle: a review of advances over 50 years. Remote Sens. Environ. 233, 111383 (2019).

    38.
    Badawy, B., Arora, V. K., Melton, J. R. & Nassar, R. Modeling the diurnal variability of respiratory fluxes in the Canadian Terrestrial Ecosystem Model (CTEM). J. Adv. Model. Earth Syst. 8, 614–633 (2016).
    ADS  Google Scholar 

    39.
    Wu, Y., Verseghy, D. L. & Melton, J. R. Integrating peatlands into the coupled Canadian Land Surface Scheme (CLASS) v3.6 and the Canadian Terrestrial Ecosystem Model (CTEM) v2.0. Geosci. Model Dev. 9, 2639–2663 (2016).
    ADS  CAS  Google Scholar 

    40.
    Waddington, J. M., Rotenberg, P. A. & Warren, F. J. Peat CO2 production in a natural and cutover peatland: implications for restoration. Biogeochemistry 54, 115–130 (2001).
    CAS  Google Scholar 

    41.
    Glatzel, S., Basiliko, N. & Moore, T. Carbon dioxide and methane production potentials of peats from natural, harvested and restored sites, eastern Québec, Canada. Wetlands 24, 261–267 (2004).
    Google Scholar 

    42.
    Hoyos-Santillan, J. et al. Quality not quantity: organic matter composition controls of CO2 and CH4 fluxes in neotropical peat profiles. Soil Biol. Biochem. 103, 86–96 (2016).
    CAS  Google Scholar 

    43.
    Moore, T. R. & Dalva, M. Methane and carbon dioxide exchange potentials of peat soils in aerobic and anaerobic laboratory incubations. Soil Biol. Biochem. 29, 1157–1164 (1997).
    CAS  Google Scholar 

    44.
    Nilsson, M. & Öquist, M. Partitioning litter mass loss into carbon dioxide and methane in peatland ecosystems. in Carbon Cycling in Northern Peatlands (American Geophysical Union, Washington, DC, 2009).

    45.
    Blodau, C., Basiliko, N. & Moore, T. R. Carbon turnover in peatland mesocosms exposed to different water table levels. Biogeochemistry 67, 331–351 (2004).
    CAS  Google Scholar 

    46.
    D’Angelo, B. et al. Soil temperature synchronisation improves representation of diel variability of ecosystem respiration in Sphagnum peatlands. Agric. Meteorol. 223, 95–102 (2016).
    Google Scholar 

    47.
    Phillips, C. L., Nickerson, N., Risk, D. & Bond, B. J. Interpreting diel hysteresis between soil respiration and temperature. Glob. Change Biol. 17, 515–527 (2011).
    ADS  Google Scholar 

    48.
    Chapin, F. S. III, Matson, P. A. & Mooney, H. A. Carbon input to terrestrial ecosystems. in principles of terrestrial ecosystem. Ecology. (Springer, New York, NY, 2002).
    Google Scholar 

    49.
    Vargas, R. & Allen, M. F. Environmental controls and the influence of vegetation type, fine roots and rhizomorphs on diel and seasonal variation in soil respiration. N. Phytol. 179, 460–471 (2008).
    CAS  Google Scholar 

    50.
    Bahn, M., Schmitt, M., Siegwolf, R., Richter, A. & Brüggemann, N. Does photosynthesis affect grassland soil-respired CO2 and its carbon isotope composition on a diurnal timescale? N. Phytol. 182, 451–460 (2009).
    CAS  Google Scholar 

    51.
    Laine, A. M. et al. Abundance and composition of plant biomass as potential controls for mire net ecosytem CO2 exchange. Botany 90, 63–74 (2012).
    CAS  Google Scholar 

    52.
    Goulden, M. L. & Crill, P. M. Automated measurements of CO2 exchange at the moss surface of a black spruce forest. Tree Physiol. 17, 537–542 (1997).
    CAS  PubMed  Google Scholar 

    53.
    Bubier, J., Crill, P., Mosedale, A., Frolking, S. & Linder, E. Peatland responses to varying interannual moisture conditions as measured by automatic CO2 chambers. Glob. Biogeochem. Cycles 17, 1066 (2003).
    ADS  Google Scholar 

    54.
    Bond-Lamberty, B., Bronson, D., Bladyka, E. & Gower, S. T. A comparison of trenched plot techniques for partitioning soil respiration. Soil Biol. Biochem. 43, 2108–2114 (2011).
    CAS  Google Scholar 

    55.
    Lai, D. Y. F., Roulet, N. T., Humphreys, E. R., Moore, T. R. & Dalva, M. The effect of atmospheric turbulence and chamber deployment period on autochamber CO2 and CH4 flux measurements in an ombrotrophic peatland. Biogeosciences 9, 3305–3322 (2012).
    ADS  CAS  Google Scholar 

    56.
    Brændholt, A., Larsen, K. S., Ibrom, A. & Pilegaard, K. Overestimation of closed-chamber soil CO2 effluxes at low atmospheric turbulence. Biogeosciences 14, 1603–1616 (2017).
    ADS  Google Scholar 

    57.
    Peichl, M., Sonnentag, O. & Nilsson, M. B. Bringing color into the picture: using digital repeat photography to investigate phenology controls of the carbon dioxide exchange in a boreal mire. Ecosystems 18, 115–131 (2015).
    CAS  Google Scholar 

    58.
    Grinsted, A., Moore, J. C. & Jevrejeva, S. Application of the cross wavelet transform and wavelet coherence to geophysical time series. Nonlinear Process. Geophys. 11, 561–566 (2004).
    ADS  Google Scholar 

    59.
    Welch, P. The use of fast Fourier transform for the estimation of power spectra: a method based on time averaging over short, modified periodograms. IEEE Trans. Audio Electroacoust. 15, 70–73 (1967).
    ADS  Google Scholar  More