More stories

  • in

    Caveats on COVID-19 herd immunity threshold: the Spain case

    Generation timeDuring the infectious period, an infected individual may produce a secondary infection. However, the individual’s infectiousness is not constant during the infectious period, but it can be approximated by the probability distribution of the generation time (GT), which accounts for the time between the infection of a primary case and the infection of a secondary case. Unfortunately, such distribution is not as easy to estimate as that of the serial interval, which accounts for the time between the onset of symptoms in a primary case to the onset of symptoms of a secondary case. This is because the time of infection is more difficult to detect than the time of symptoms onset. Ganyani et al.27 developed a methodology to estimate the distribution of the GT from the distributions of the incubation period and the serial interval. Assuming an incubation period following a gamma distribution with a mean of 5.2 days and a standard deviation (SD) of 2.8 days, they estimated the serial interval from 91 and 135 pairs of documented infector-infectee in Singapore and Tianjin (China). Then, they found that the GT followed a gamma distribution with mean = 5.20 (95% CI = [3.78, 6.78]) days and SD = 1.72 (95% CI = [0.91, 3.93]) for Singapore (hereafter GT1), and with mean = 3.95 (95% CI = [3.01, 4.91]) days and SD = 1.51 (95% CI = [0.74, 2.97]) for Tianjin (hereafter GT2). Ng et al.28 applied the same methodology to 209 pairs of infector-infectee in Singapore and determined a gamma distribution with mean = 3.44 (95% CI = [2.79, 4.11]) days and SD 2.39 (95% CI = [1.27, 3.45]; hereafter GT3). Figure 3 shows the probability density functions (PDF) of such distributions, fGT. The differences between them are remarkable. For example, the 54.5%, 81.0%, and 80.7% of the contagions are produced in a pre-symptomatic stage (in the first 5.2 days after primary infection) assuming GT1, GT2, and GT3, respectively.Figure 3Probability density function of the generation time distribution, fGT(t), of GT1 (blue line; Singapore27), GT2 (yellow line; Tianjin27), GT3 (red line; Singapore28), and GTth (black line; theoretical distribution). Bars are the discretized version, (widetilde{{f_{GT} }}left( n right)), of the PDF of GTth.Full size imageTheoretically, assuming that the incubation periods of two individuals are independent and identically distributed, which is quite plausible, the expected/mean values of the GT and the serial interval should be equal29,30. The mean of the serial interval is easier to estimate than that of the GT. For that reason, we assume a mean serial interval as estimated from a meta-analysis of 13 studies involving a total of 964 pairs of infector-infectee, which is 4.99 days (95% CI = [4.17, 5.82])31, is more reliable than the aforementioned means of the GT. This value is within the error estimates of the means of GT1 and GT2, but not for GT3. Then, we construct a theoretical distribution for the GT that follows a gamma distribution (hereafter GTth) with mean = 4.99 days and SD = 1.88 days. This theoretical distribution can be seen in Fig. 3 and approximates the average PDF of three gamma distributions with mean = 4.99 and the SD of GT1, GT2, and GT3. We assume a conservative CI = [1.51, 2.39] for the theoretical SD, defined with the minimum and maximum SD values of GT1, GT2, and GT3. GTth shows 63.1% of pre-symptomatic contagions.
    R

    0

    from r
    In theory, the basic reproduction number R0 can be estimated as far as the intrinsic growth rate r, and the distributions of both the latent and infectious periods are known26,32,33,34. The latent period accounts for the period during which an infected individual cannot infect other individuals. It is observed in diseases for which the infectious period starts around the end of the incubation period, as happened with influenza35 and SARS36. However, from Fig. 3 it is inferred that COVID-19 is transmissible from the moment of infection, and we will assume a null latent period. Then, if the GT follows a gamma distribution, R0 can be estimated from the formulation of Anderson and Watson32, which was adapted to null latent periods by Yan26 as$$ R_{0} = frac{{mean_{GT} }}{{1 – left( {1 + mean_{GT} cdot r cdot frac{1}{{shape_{GT} }}} right)^{{ – shape_{GT} }} }} cdot r, $$
    (4)
    where meanGT is the mean GT and shapeGT is one of the two parameters defining the gamma distribution, which can be estimated as$$ shape_{GT} = frac{{left( {mean_{GT} } right)^{2} }}{{left( {SD_{GT} } right)^{2} }}. $$
    (5)
    For GTth, we get R0 = 1.50 (CI = [1.41, 1.61]) for REMEDID I(n) and R0 = 1.76 (CI = [1.60, 1.94]) for official I(n). For the other three GT distributions, R0 ranges from 1.39 (CI = [1.27, 1.58]) to 1.51 (CI = [1.34, 1.80]) for REMEDID I(n) and from 1.59 (CI = [1.40, 1.88]) to 1.78 (CI = [1.51, 2.23]) for official I(n) (Table 1). In all cases, R0 from GTth are within those from the three known GT distributions and indistinguishable from them within the error estimates. The lower (upper) bound of the CI is estimated as the minimum (maximum) R0 obtained from all the possible combinations of 100 evenly spaced values covering the CI of r, meanGT and SDGT. Then, following the Bonferroni correction, the reported CI present at least a 85% of confidence level for GT1, GT2, and GT3, but it cannot be assured for GTth since the CI of its SD is unknown. In general, all these R0 estimates are lower than those summarised by Park et al.20.Table 1 R0 and HIT values of the ancestral SARS-CoV-2 variant estimated from GT1, GT2, GT3, and GTth, and REMEDID and official infections. For date0, “Dec.” means December 2019, and “Jan.” means January 2020.Full size tableAlternatively, R0 can be estimated by applying the Euler–Lotka equation29,33,$$ R_{0} = frac{1}{{mathop smallint nolimits_{0}^{ + infty } e^{ – rt} cdot f_{GT} left( t right)dt}}. $$
    (6)
    In this case, we get values closer to previous estimates20. In particular, for GTth, we get R0 = 2.12 (CI = [1.81, 2.48]) for REMEDID I(n) and R0 = 2.92 (CI = [2.28, 3.75]) for official I(n). For the other three GT distributions, R0 ranges from 1.63 (CI = [1.43, 1.90]) to 2.21 (CI = [1.59, 2.95]) for REMEDID I(n) and from 1.97 (CI = [1.59, 2.54]) to 3.11 (CI = [1.84, 4.90]) for official I(n) (Table 1). The CI are estimated as in Eq. (4).R0 from a dynamical modelWe designed a dynamic model with Susceptible-Infected-Recovered (SIR) as stocks that accounts for the infectiousness of the infectors. Such a model is a generalisation of the Susceptible-Exposed-Infected-Recovered (SEIR) model37. Births, deaths, immigration and emigration are ignored, which seems reasonable since the timescale of the outbreak is too short to produce significant demographic changes. For the sake of simplicity, the recovered stock includes recoveries and fatalities, and it is denoted as R(t). A random mixing population is assumed, that is a population where contacts between any two people are equally probable. Time is discretized in days, so the real time variable t is replaced by the integer variable n. As a consequence, the derivatives in the differential equations defining the dynamic model explained below are discrete derivatives.The size of the population is fixed at N = 100,000, and then, for any day n we get$$ tilde{S}left( n right) + left( {mathop sum limits_{k = 0}^{20} tilde{I}left( n-k right)} right) + tilde{R}left( n right) = N, $$
    (7)
    where (tilde{S}left( n right)), (tilde{I}left( n right)), and (tilde{R}left( n right)) are the discretized versions of S(t), I(t), and R(t) and (tilde{I}) is assumed to be null for negative integers. The summation is a consequence of the infectiousness, which is approximated according to the GT, whose PDF is discretized as$$ widetilde{{f_{GT} }}left( n right) = mathop smallint limits_{n – 1}^{n} f_{GT} left( t right) dt, $$
    (8)
    from n = 1 to 20. Figure 3 shows (widetilde{{f_{GT} }}left( n right)) for GTth. Truncating at n = 20 accounts for 99.99% of the area below the PDF of all the GT. Then, an infected individual at day n0 is expected to produce on average$$ widetilde{{R_{e} }}left( {n_{0} + n} right) cdot widetilde{{f_{GT} }}left( n right) $$
    (9)
    infections n days later, where (widetilde{{R_{e} }}left( n right)) is the discretized version of Re(t). From this expression, it is obvious that values of (widetilde{{R_{e} }}left( n right) < 1) will produce a decline of infections. Conversely, infections at day n0 are produced by all individuals infected during the previous 20 days as$$ tilde{I}(n_{0} ) = tilde{R}_{e} left( {n_{0} } right) cdot left( {mathop sum limits_{n = 1}^{20} tilde{I}left( {n_{0} - n} right) cdot widetilde{{f_{GT} }}left( n right)} right), $$ (10) whose continuous version has been reported in previous studies29,38. The expression in brackets is called total infectiousness of infected individuals at day n039. According to Eq. (1), Eq. (10) can be expressed in terms of R0 as$$ tilde{I}(n_{0} ) = R_{0} cdot frac{{tilde{S}left( {n_{0} } right)}}{N} cdot left( {mathop sum limits_{n = 1}^{20} tilde{I}left( {n_{0} - n} right) cdot widetilde{{f_{GT} }}left( n right)} right). $$ (11) As we want a dynamic model capable of providing (tilde{I}left( {n_{0} } right)) from the stocks at time step n0 − 1, we replaced (tilde{S}left( {n_{0} } right)) by (tilde{S}left( {n_{0} - 1} right)) in Eq. (11). This assumption makes sense in a discrete domain since the infections at time n0 take place in the susceptible population at time n0 − 1. Then, assuming that all stocks are set to zero for negative integers, our dynamic model can be expressed in terms of Eq. (7) and the following differential equations:$$ delta tilde{I}(n_{0} ) = R_{0} cdot frac{{tilde{S}left( {n_{0} - 1} right)}}{N} cdot left( {mathop sum limits_{n = 1}^{20} tilde{I}left( {n_{0} - n} right) cdot widetilde{{{text{f}}_{GT} }}left( n right)} right) - tilde{I}(n_{0} - 1), $$ (12) $$ delta tilde{S}left( {n_{0} } right) = {-}tilde{I}left( {n_{0} } right), $$ (13) $$ delta tilde{R}left( {n_{0} } right) = tilde{I}left( {n_{0} - 21} right), $$ (14) where (delta tilde{I}), (delta tilde{S}), and (delta tilde{R}) are the (discrete) derivatives of (tilde{I}), (tilde{S}), and (tilde{R}), respectively. Applying the initial conditions (tilde{S}left( 0 right) = N - 1), (tilde{I}left( 0 right) = 1), and (tilde{R}left( 0 right) = 0), it is assumed that the outbreak was produced by only one infector. The latter is not true in Spain, since several independent introductions of SARS-CoV-2 were detected40. However, for modelling purposes it is equivalent to introducing a single infection at day 0 or M infections produced by the single infection n days later. Then, the date of the initial time n = 0 is accounted as a parameter date0, which is optimised, as well as R0, to minimise the root-mean square of the residual between the model simulated (tilde{I}left( n right)) and the REMEDID and official I(n) for the period from date0 to n0.The model was implemented in Stella Architect software v2.1.1 (www.iseesystems.com) and exported to R software v4.1.1 with the help of deSolve (v1.28) and stats (v4.1.1) packages, and the Brent optimisation algorithm was implemented. For REMEDID I(n) and GTth, we obtained date0 = 13 December 2019 and R0 = 2.71 (CI = [2.33, 3.15]). Optimal solutions combine lower/higher R0 and earlier/later date0 (Fig. 4), which highlights the importance of providing an accurate first infection date to estimate R0. When the other three GT distributions were considered, we obtained similar date0, ranging from 12 to 17 December 2019, and R0 values ranging from 2.08 (CI = [1.86, 2.42]) to 2.85 (CI = [2.05, 3.25]; see Table 1). For official infections, date0 was set to 1 January 2020 for all cases, and R0 ranged from 1.81 (CI = [1.64, 2.07]) to 2.41 (CI = [1.80, 2.91]). The CI are estimated as in Eq. (4).Figure 4Root-mean square (RMS) of the residuals between infections from the model, which depends on date0 (x-axis) and R0 (y-axis), and REMEDID (from MoMo ED) and official infections. Parameters optimizing the model are highlighted in purple. RMS larger than 1275 (left panel) and 103 (right panel) are saturated in white.Full size image More

  • in

    Challenging the sustainability of urban beekeeping using evidence from Swiss cities

    1.Federal Office for the Environment (FOEN). Action Plan for the Swiss Biodiversity Strategy (FOEN, Bern, 2017).2.Geldmann, J. & González-Varo, J. P. Conserving honey bees does not help wildlife. Science 359, 392–393 (2018).Article 

    Google Scholar 
    3.Egerer, M. & Kowarik, I. Confronting the modern gordian knot of urban beekeeping. Trends Ecol. Evol. 35, 956–959 (2020).Article 

    Google Scholar 
    4.Herrera, C. M. Gradual replacement of wild bees by honeybees in flowers of the Mediterranean Basin over the last 50 years. Proc. R. Soc. B Biol. Sci. 287, 16–20 (2020).
    Google Scholar 
    5.Torné-Noguera, A., Rodrigo, A., Osorio, S. & Bosch, J. Collateral effects of beekeeping: impacts on pollen-nectar resources and wild bee communities. Basic Appl. Ecol. 17, 199–209 (2016).Article 

    Google Scholar 
    6.Magrach, A., González-Varo, J. P., Boiffier, M., Vilà, M. & Bartomeus, I. Honeybee spillover reshuffles pollinator diets and affects plant reproductive success. Nat. Ecol. Evol. 1, 1299–1307 (2017).Article 

    Google Scholar 
    7.Prendergast, K. S., DIxon, K. W. & Bateman, P. W. Interactions between the introduced European honey bee and native bees in urban areas varies by year, habitat type and native bee guild. Biol. J. Linn. Soc. 133, 725–743 (2021).Article 

    Google Scholar 
    8.Herbertsson, L., Lindström, S. A. M., Rundlöf, M., Bommarco, R. & Smith, H. G. Competition between managed honeybees and wild bumblebees depends on landscape context. Basic Appl. Ecol. 17, 609–616 (2016).Article 

    Google Scholar 
    9.Stevenson, P. C. et al. The state of the world’s urban ecosystems: what can we learn from trees, fungi, and bees? Plants People Planet 2, 482–498 (2020).Article 

    Google Scholar 
    10.Ropars, L., Dajoz, I., Fontaine, C., Muratet, A. & Geslin, B. Wild pollinator activity negatively related to honey bee colony densities in urban context. PLoS ONE 14, e0222316 (2019).CAS 
    Article 

    Google Scholar 
    11.Venter, Z. S. & Sydenham, M. A. K. Continental-scale land cover mapping at 10 m resolution over Europe (ELC10). Remote Sens. 13, 2301 (2021).Article 

    Google Scholar 
    12.Hardin, G. The tragedy of the commons. Science 162, 1243–1248 (1968).CAS 
    Article 

    Google Scholar 
    13.Tew, N. E. et al. Quantifying nectar production by flowering plants in urban and rural landscapes. J. Ecol. 109, 1747–1757 (2021).Article 

    Google Scholar 
    14.Baldock, K. C. R. et al. A systems approach reveals urban pollinator hotspots and conservation opportunities. Nat. Ecol. Evol. 3, 363–373 (2019).Article 

    Google Scholar 
    15.Casanelles-Abella, J. et al. Applying predictive models to study the ecological properties of urban ecosystems: A case study in Zürich, Switzerland. Landsc. Urban Plan. 214, 104137 (2021).Article 

    Google Scholar 
    16.IPBES. Summary for policymakers of the global assessment report on biodiversity and ecosystem services of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services (IPBES, 2019).17.IUCN. IUCN’s Key Messages. First Draft of the Post-2020 Global Biodiversity Framework. In Convention on Biological Diversity Third meeting of the Open-Ended Working Group on the Post-2020 Global Biodiversity Framework (OEWG3) 15 (IUCN, 2021).18.Baldock, K. C. R. Opportunities and threats for pollinator conservation in global towns and cities. Curr. Opin. Insect Sci. 38, 63–71 (2020).Article 

    Google Scholar 
    19.Henry, M. & Rodet, G. The apiary influence range: a new paradigm for managing the cohabitation of honey bees and wild bee communities. Acta Oecologica 105, 103555 (2020).Article 

    Google Scholar 
    20.Ignatieva, M. & Hedblom, M. An alternative urban green carpet. Science 362, 148–149 (2018).CAS 
    Article 

    Google Scholar 
    21.Vega, K. A. & Küffer, C. Promoting wildflower biodiversity in dense and green cities: The important role of small vegetation patches. Urban For. Urban Green. 62, 127165 (2021).Article 

    Google Scholar 
    22.Fabián, D., González, E., Sánchez Domínguez, M. V., Salvo, A. & Fenoglio, M. S. Towards the design of biodiverse green roofs in Argentina: assessing key elements for different functional groups of arthropods. Urban For. Urban Green. 61 (2021).23.Vega, K. A., Schläpfer-Miller, J. & Kueffer, C. Discovering the wild side of urban plants through public engagement. Plants People Planet 3, 389–401 (2021).Article 

    Google Scholar 
    24.Von Büren, R. S., Oehen, B., Kuhn, N. J. & Erler, S. High-resolution maps of Swiss apiaries and their applicability to study spatial distribution of bacterial honey bee brood diseases. PeerJ 2019, 1–21 (2019).
    Google Scholar 
    25.QGIS Development Team. QGIS Geographic Information System (Open Source Geospatial Foundation Project, 2020).26.R Core Team. R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, Vienna, 2019).27.R Studio Team. R studio: Integrated Development for R (RStudio, Boston, 2020).28.Casanelles-Abella, J. & Moretti M. Challenging the sustainability of urban beekeeping: evidence from Swiss cities. Envidat. https://doi.org/10.16904/envidat.239 (2021).29.Casanelles-Abella, J. Code for the paper: Challenging the sustainability of urban beekeeping: evidence from Swiss cities (v. 1.0). Zenodo https://doi.org/10.5281/zenodo.5618254 (2021).30.Swiss Fedearl Office of Topography Swisstopo. SWISSIMAGE 25. https://www.swisstopo.admin.ch/en/geodata/images/ortho/swissimage25.html#links (Swisstopo, 2021). More

  • in

    Plant neighborhood shapes diversity and reduces interspecific variation of the phyllosphere microbiome

    1.Lindow SE, Brandl MT. Microbiology of the phyllosphere. Appl Environ Microbiol. 2003;69:1875LP–1883.
    Google Scholar 
    2.Morella NM, Zhang X, Koskella B. Tomato seed-associated bacteria confer protection of seedlings against foliar disease caused by Pseudomonas syringae. Phytobiomes J. 2019;3:177–90.
    Google Scholar 
    3.Innerebner G, Knief C, Vorholt JA. Protection of Arabidopsis thaliana against leaf-pathogenic Pseudomonas syringae by Sphingomonas strains in a controlled model system. Appl Environ Microbiol. 2011;77:3202–10.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    4.Fu S-F, Sun P-F, Lu H-Y, Wei J-Y, Xiao H-S, Fang W-T, et al. Plant growth-promoting traits of yeasts isolated from the phyllosphere and rhizosphere of Drosera spatulata Lab. Fungal Biol. 2016;120:433–48.CAS 
    PubMed 

    Google Scholar 
    5.Laforest-Lapointe I, Paquette A, Messier C, Kembel SW. Leaf bacterial diversity mediates plant diversity and ecosystem function relationships. Nature. 2017;546:145–7.CAS 
    PubMed 

    Google Scholar 
    6.Lindow SE, Leveau JHJ. Phyllosphere microbiology. Curr Opin Biotechnol. 2002;13:238–43.CAS 
    PubMed 

    Google Scholar 
    7.Fürnkranz M, Wanek W, Richter A, Abell G, Rasche F, Sessitsch A. Nitrogen fixation by phyllosphere bacteria associated with higher plants and their colonizing epiphytes of a tropical lowland rainforest of Costa Rica. ISME J. 2008;2:561–70.PubMed 

    Google Scholar 
    8.Ottesen AR, Gorham S, Reed E, Newell MJ, Ramachandran P, Canida T, et al. Using a control to better understand phyllosphere microbiota. PLoS ONE. 2016;11:e0163482.PubMed 
    PubMed Central 

    Google Scholar 
    9.Jones JDG, Dangl JL. The plant immune system. Nature. 2006;444:323–9.CAS 
    PubMed 

    Google Scholar 
    10.Bodenhausen N, Bortfeld-miller M, Ackermann M, Vorholt JA. A synthetic community approach reveals plant genotypes affecting the phyllosphere microbiota. PLoS Biol. 2014; 10. https://doi.org/10.1371/journal.pgen.1004283.11.Horton MW, Bodenhausen N, Beilsmith K, Meng D, Muegge BD, Subramanian S, et al. Genome-wide association study of Arabidopsis thaliana leaf microbial community. Nat Commun. 2014;5:5320.PubMed 

    Google Scholar 
    12.Hacquard S, Spaepen S, Garrido-Oter R, Schulze-Lefert P. Interplay between innate immunity and the plant microbiota. Annu Rev Phytopathol. 2017;55:565–89.CAS 
    PubMed 

    Google Scholar 
    13.Zhalnina K, Louie KB, Hao Z, Mansoori N, Nunes U, Shi S et al. Dynamic root exudate chemistry and substrate preferences drive patterns in rhizosphere microbial community assembly. Nat Microbiol. 2018. https://doi.org/10.1038/s41564-018-0129-3.14.Humphrey PT, Whiteman NK. Insect herbivory reshapes a native leaf microbiome. Nat Ecol Evol. 2020;4:221–9.PubMed 
    PubMed Central 

    Google Scholar 
    15.Yadav RKP, Karamanoli K, Vokou D. Bacterial colonization of the phyllosphere of Mediterranean perennial species as influenced by leaf structural and chemical features. Micro Ecol. 2005;50:185–96.CAS 

    Google Scholar 
    16.Morella NM, Weng FCH, Joubert PM, Metcalf CJE, Lindow S, Koskella B. Successive passaging of a plant-associated microbiome reveals robust habitat and host genotype-dependent selection. Proc Natl Acad Sci USA. 2020;117:1148–59.CAS 
    PubMed 

    Google Scholar 
    17.Wagner MR, Busby PE, Balint-Kurti P. Analysis of leaf microbiome composition of near-isogenic maize lines differing in broad-spectrum disease resistance. N Phytol. 2019;225:2152–65.
    Google Scholar 
    18.Wagner MR, Lundberg DS, Del Rio TG, Tringe SG, Dangl JL, Mitchell-Olds T. Host genotype and age shape the leaf and root microbiomes of a wild perennial plant. Nat Commun. 2016;7:1–15.CAS 

    Google Scholar 
    19.Horner-Devine MC, Bohannan BJM. Phylogenetic clustering and overdispersion in bacterial communities. Ecology. 2006;87:S100–8.PubMed 

    Google Scholar 
    20.Kembel SW, O’Connor TK, Arnold HK, Hubbell SP, Wright SJ, Green JL. Relationships between phyllosphere bacterial communities and plant functional traits in a neotropical forest. Proc Natl Acad Sci USA 2014; 1–6.21.Burns AR, Stephens WZ, Stagaman K, Wong S, Rawls JF, Guillemin K, et al. Contribution of neutral processes to the assembly of gut microbial communities in the zebrafish over host development. ISME J. 2016;10:655–64.CAS 
    PubMed 

    Google Scholar 
    22.Sloan WT, Woodcock S, Lunn M, Head IM, Curtis TP. Modeling taxa-abundance distributions in microbial communities using environmental sequence data. Micro Ecol. 2007;53:443–55.
    Google Scholar 
    23.Laforest-Lapointe I, Messier C, Kembel SW. Host species identity, site and time drive temperate tree phyllosphere bacterial community structure. Microbiome 2016; 1–10.24.Schlaeppi K, Dombrowski N, Oter RG, Ver Loren van Themaat E, Schulze-Lefert P. Quantitative divergence of the bacterial root microbiota in Arabidopsis thaliana relatives. Proc Natl Acad Sci USA. 2014;111:585LP–592.
    Google Scholar 
    25.Gallart M, Adair KL, Love J, Meason DF, Clinton PW, Xue J, et al. Host genotype and nitrogen form shape the root microbiome of Pinus radiata. Micro Ecol. 2018;75:419–33.CAS 

    Google Scholar 
    26.Hambäck PA, Inouye BD, Andersson P, Underwood N. Effects of plant neighborhoods on plant–herbivore interactions: resource dilution and associational effects. Ecology. 2014;95:1370–83.PubMed 

    Google Scholar 
    27.Underwood N, Inouye BD, Hambäck PA. A conceptual framework for associational effects: when do neighbors matter and how would we know? Q Rev Biol. 2014;89:1–19.PubMed 

    Google Scholar 
    28.Barbosa P, Hines J, Kaplan I, Martinson H, Szczepaniec A, Szendrei Z. Associational resistance and associational susceptibility: having right or wrong neighbors. Annu Rev Ecol Evol Syst. 2009;40:1–20.
    Google Scholar 
    29.Janzen DH. Herbivores and the number of tree species in tropical forests. Am Nat. 1970;104:501–28.
    Google Scholar 
    30.Connell JH. On the role of natural enemies in preventing competitive exclusion in some marine animals and in rain forest trees. in Den Boer PJ, Gradwell G, editors. Dynamics of populations. PUDOC, 1971, p. 298–312.31.Mangan SA, Schnitzer SA, Herre EA, Mack KML, Valencia MC, Sanchez EI, et al. Negative plant–soil feedback predicts tree-species relative abundance in a tropical forest. Nature. 2010;466:752–5.CAS 
    PubMed 

    Google Scholar 
    32.Miller EC, Perron GG, Collins CD. Plant‐driven changes in soil microbial communities influence seed germination through negative feedbacks. Ecol Evol. 2019;0:1–14.
    Google Scholar 
    33.Antonovics J, Ellstrand NC. Experimental studies of the evolutionary significance of sexual reproduction. I. A test of the frequency-dependent selection hypothesis. Evolution. 1984;38:103–15.PubMed 

    Google Scholar 
    34.Ellstrand NC, Antonovics J. Experimental studies of the evolutionary significance of sexual reproduction II. A test of the density-dependent selection hypothesis. Evolution. 1985;39:657–66.PubMed 

    Google Scholar 
    35.Naeem S, Tjossem SF, Byers D, Bristow C, Li S. Plant neighborhood diversity and production. Ecoscience. 1999;6:355–65.
    Google Scholar 
    36.Worrich A, Musat N. Associational effects in the microbial neighborhood. ISME J 2019; 2143–9.37.Copeland JK, Yuan L, Layeghifard M, Wang PW, Guttman DS. Seasonal community succession of the phyllosphere microbiome. Mol Plant Microbe Interact. 2015;28:274–85.CAS 
    PubMed 

    Google Scholar 
    38.Lajoie G, Kembel SW. Host neighborhood shapes bacterial community assembly and specialization on tree species across a latitudinal gradient. Ecol Monogr. 2021;0:1–18.
    Google Scholar 
    39.Lymperopoulou D, Adams R, Lindow SE, Löffler F. Contribution of vegetation to the microbial composition of nearby outdoor air. Appl Environ Microbiol. 2016;82:3822–33.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    40.Lindow SE, Andersen G. Influence of immigration on epiphytic bacterial populations on navel orange leaves. Appl Environ Microbiol. 1996;62:2978–87.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    41.Massoni J, Bortfeld-miller M, Widmer A, Vorholt JA. Capacity of soil bacteria to reach the phyllosphere and convergence of floral communities despite soil microbiota variation. Proc Natl Acad Sci USA 2021; 118. https://doi.org/10.1073/pnas.2100150118.42.Leibold MA, Holyoak M, Mouquet N, Amarasekare P, Chase JM, Hoopes MF, et al. The metacommunity concept: a framework for multi-scale community ecology. Ecol Lett. 2004;7:601–13.
    Google Scholar 
    43.Fodelianakis S, Lorz A, Valenzuela-cuevas A, Barozzi A, Booth JM, Daffonchio D. Dispersal homogenizes communities via immigration even at low rates in a simplified synthetic bacterial metacommunity. Nat Commun. 2019;10:1–12.
    Google Scholar 
    44.Burns AR, Miller E, Agarwal M, Rolig AS, Milligan-Myhre K, Seredick S et al. Interhost dispersal alters microbiome assembly and can overwhelm host innate immunity in an experimental zebrafish model. Proc Natl Acad Sci USA 2017;114. https://doi.org/10.1073/pnas.1702511114.45.Chelius MK, Triplett EW. The diversity of archaea and bacteria in association with the roots of Zea mays L. Micro Ecol. 2001;41:252–63.CAS 

    Google Scholar 
    46.Bodenhausen N, Horton MW, Bergelson J. Bacterial communities associated with the leaves and the roots of Arabidopsis thaliana. PLoS ONE. 2013;8:e56329.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    47.Lundberg DS, Yourstone S, Mieczkowski P, Jones CD, Dangl JL. Practical innovations for high-throughput amplicon sequencing. Nat Methods. 2013;10:999–1002.CAS 
    PubMed 

    Google Scholar 
    48.Callahan BJ, McMurdie PJ, Rosen MJ, Han AW, Johnson AJA, Holmes SP. DADA2: high-resolution sample inference from Illumina amplicon data. Nat Methods. 2016;13:581.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    49.R Core Team. R: A language and environment for statistical computing. 2020. http://cran.r-project.org.50.Morgan M, Anders S, Lawrence M, Aboyoun P, Pagès H, Gentleman R. ShortRead: a bioconductor package for input, quality assessment and exploration of high-throughput sequence data. Bioinformatics. 2009;25:2607–8.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    51.Pagès H, Aboyoun P, Gentleman R, DebRoy S. Biostrings: efficient manipulation of biological strings. 2020.52.McMurdie P, Holmes S. phyloseq: An R package for reproducible interactive analysis and graphics of microbiome census data. PLoS ONE 2013; 8. https://doi.org/10.1371/journal.pone.0061217tle.53.Wang Q, Garrity GM, Tiedje JM, Cole JR. Naive bayesian classifier for rapid assignment of rRNA sequences into the new bacterial taxonomy. Appl Environ Microbiol. 2007;73:5261–7.CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    54.Quast C, Pruesse E, Yilmaz P, Gerken J, Schweer T, Yarza P, et al. The SILVA ribosomal RNA gene database project: improved data processing and web-based tools. Nucleic Acids Res. 2013;41:D590–D596.CAS 
    PubMed 

    Google Scholar 
    55.Davis NM, Proctor DM, Holmes SP, Relman DA, Callahan BJ. Simple statistical identification and removal of contaminant sequences in marker-gene and metagenomics data. Microbiome. 2018;6:226.PubMed 
    PubMed Central 

    Google Scholar 
    56.Morella NM, Gomez AL, Wang G, Leung MS, Koskella B. The impact of bacteriophages on phyllosphere bacterial abundance and composition. Mol Ecol. 2018;27:2025–38.PubMed 

    Google Scholar 
    57.Oksanen J, Blanchet FG, Roeland K, Legendre P, Minchin P, O’Hara RB et al. vegan: Community ecology package. 2015. http://cran.r-project.org.58.Anderson MJ. A new method for non-parametric multivariate analysis of variance. Austral Ecol. 2001;26:32–46.
    Google Scholar 
    59.De Cáceres M, Legendre P. Associations between species and groups of sites: indices and statistical inference. Ecology. 2009;90:3566–74.PubMed 

    Google Scholar 
    60.Ersts PJ. Geographic Distance Matrix Generator. http://biodiversityinformatics.amnh.org/open_source/gdmg.61.Sprockett D. reltools: Microbiome Amplicon Analysis and Visualization. 2021.62.Sloan WT, Lunn M, Woodcock S, Head IM, Nee S, Curtis TP. Quantifying the roles of immigration and chance in shaping prokaryote community structure. Environ Microbiol. 2006;8:732–40.PubMed 

    Google Scholar 
    63.Wright ES. Using DECIPHER v2.0 to analyze big biological sequence data in R. R J. 2016;8:352–9.
    Google Scholar 
    64.Schliep KP. phangorn: phylogenetic analysis in R. Bioinformatics. 2011;27:592–3.CAS 
    PubMed 

    Google Scholar 
    65.Kembel SW, Cowan PD, Helmus MR, Cornwell WK, Morlon H, Ackerly DD, et al. Picante: R tools for integrating phylogenies and ecology. Bioinformatics. 2010;26:1463–4.CAS 
    PubMed 

    Google Scholar 
    66.Koskella B. The phyllosphere. Curr Biol. 2020;30:R1143–R1146.CAS 
    PubMed 

    Google Scholar 
    67.Chaparro JM, Badri DV, Vivanco JM. Rhizosphere microbiome assemblage is affected by plant development. ISME J. 2014;8:790–803.CAS 
    PubMed 

    Google Scholar 
    68.İnceoğlu Ö, Al-Soud WA, Salles JF, Semenov AV, van Elsas JD. Comparative analysis of bacterial communities in a potato field as determined by pyrosequencing. PLoS ONE. 2011;6:e23321.PubMed 
    PubMed Central 

    Google Scholar 
    69.Christian N, Herre EA, Mejia LC, Clay K. Exposure to the leaf litter microbiome of healthy adults protects seedlings from pathogen damage. Proc R Soc B Biol Sci. 2017;284:20170641.
    Google Scholar 
    70.Leigh EG, Davidar P, Dick CW, Terborgh J, Puyravaud J-P, ter Steege H, et al. Why do some tropical forests have so many species of trees? Biotropica. 2004;36:447–73.
    Google Scholar 
    71.Hyatt LA, Rosenberg MS, Howard TG, Bole G, Fang W, Anastasia J, et al. The distance dependence prediction of the Janzen-Connell hypothesis: a meta-analysis. Oikos. 2003;103:590–602.
    Google Scholar 
    72.Carson W, Anderson J, Leigh E, Schnitzer S. Challenges associated with testing and falsifying the Janzen_Connell hypothesis: a review and critique. In: Carson W, Schnitzer SA, editors. Tropical forest community ecology. Wiley Blackwell; 2008. p. 210–41. More

  • in

    The formation of avian montane diversity across barriers and along elevational gradients

    Genome sequencing and assemblyGenome assemblies ranged in size from 799.9 Mbp in Melanocharis versteri to 1053.5 Mbp in Sericornis nouhuysi. The number of scaffolds ranged from 14,086 scaffolds in Melipotes ater to 87,957 scaffolds in Ficedula hyperythra and N50 ranged between ca. 40 Kbp to and 25 Mbp. Benchmarking Universal Single-Copy Orthologs (BUSCO) analyses of genome completeness ranged from a high proportion of complete BUSCOs in Melipotes ater, 86.8% to only 66.7% complete BUSCOs in Rhipidura albolimbata. For most species, the proportions of complete BUSCOs were 75–80%. Overall, the proportion of missing BUSCOs was low, ranging from 6.6% in Melipotes ater to 15.2% in Rhipidura albolimbata (see Supplementary Table 1 for all genome assembly statistics and Supplementary Fig. 1 for the number of SNP variants per species).Kinship analyses of individuals within populationsSampling of closely related individuals can dramatically bias estimates of population structure and demographics. Two Pachycephala schlegelii individuals (A117 and A118) showed a pairwise kinship coefficient of 0.144, indicative of being half-siblings. The two individuals were collected at the same locality on the same date. Similarly, two Ifrita kowaldi individuals (D116 and D117) showed a pairwise kinship coefficient of 0.135, also suggestive of being half-siblings. In this case, the individuals were collected on the same sampling locality on two consecutive days. To not bias downstream demographic analyses, one of the P. schlegelii (A118) and one of the I. kowaldi (D117) individuals were excluded from all subsequent analyses. For all other species, no closely related individuals were identified.Genetic differentiationEstimated levels of differentiation between populations were initially based on three approaches; (i) calculation of FST (the fixation index), which quantifies the degree of genetic differentiation between populations based on the variation in allele frequencies, ranging between 0 (complete mixing of individuals) and 1 (complete separation of populations) (Fig. 1), (ii) Standardized pairwise FST used to conduct a Principal Component Analysis (PCA) in order to visualize population structure (Supplementary Fig. 1) and (iii) Admixture analysis as implemented in STRUCTURE (a clustering algorithm that infers the most likely number of groups [K]), in which individuals are grouped into clusters according to the proportion of their ancestry components (Supplementary Fig. 1). As a preliminary analysis, we calculated FST and constructed PCA plots for the four congeneric (incl. Sericornis/Aethomyias [until recently placed in the genus Sericornis]) species pairs in our study (Supplementary Fig. 2), which were aligned using the same reference genome. This was done to ascertain that no samples had been misidentified and to gauge levels of differentiation between distinct species. All species were genetically well separated and FST values ranged from 0.08 for the two Ptiloprora species to 0.20 for the two Ficedula species.For five out of six species from Buru/Seram, genetic differentiation (FST) was high between islands (Fig. 1), and comparable to differentiation between named congeneric species in this study (e.g. Ptiloprora and Melipotes); Ceyx lepidus (FST = 0.16), Thapsinillas affinis (FST = 0.15), Ficedula buruensis (FST = 0.13) and Pachycephala macrorhyncha (FST = 0.09). In contrast, differentiation in Ficedula hyperythra was consistent with population-level differentiation (FST = 0.04). In all cases, individuals from Buru and Seram were clearly differentiated in the PCA and STRUCTURE plots (Supplementary Fig. 1A). For Ceyx lepidus, Ficedula buruensis and Pachycephala macrorhyncha, samples were collected at multiple elevations and we therefore calculated genetic differentiation between elevations (Buru: 1097 m versus 1435 m and Seram: 1000 m versus 1300 m) to determine any potential parapatric differentiation along the gradients. In all possible comparisons, FST values did not differ significantly from 0. Moreover, PCA plots showed that samples did not cluster according to elevation (Supplementary Fig. 3A).Three of the thirteen New Guinean population pairs occurring in Mount Wilhelm and Huon showed relatively high genetic divergences: Melipotes fumigatus/ater (FST = 0.08), Paramythia montium (FST = 0.09) and Ifrita kowaldi (FST = 0.07) (Fig. 1) with populations clearly separated (Supplementary Fig. 1). By contrast, the two lowland species Toxorhamphus novaeguineae and Melilestes megarhynchus showed little genetic differentiation, FST = 0.00. For the remaining species, genetic differentiation between Mount Wilhelm and Huon ranged between FST = 0.01–0.05. Despite this moderate level of genetic differentiation, the populations of Mount Wilhelm and Huon could be clearly distinguished in the PCA plots. In all cases STRUCTURE suggested a scenario with K = 2 with some mixing of individuals, except for Rhipidura albolimbata, in which K = 1 was suggested.For five bird species we included an additional population from Mount Scratchley, which is also situated in the Central Range but ~400 km to the southeast of Mount Wilhelm. Genetic differentiation of this population from the other two populations was comparable with that between Mount Wilhelm and Huon. The highest genetic differentiation was found in Paramythia montium (FST = 0.10 both between Mount Wilhelm and Mount Scratchley and between Huon and Mount Scratchley). In the case of Peneothello sigillata, the Mount Scratchley population appeared genetically well-differentiated from both the populations of Mount Wilhelm (FST = 0.06) and Huon (FST = 0.07). In both cases, STRUCTURE suggested a scenario of K = 3, with individual assignments matching the three geographically circumscribed populations. For Pachycephala schlegelii, genetic differentiation was relatively high between Huon and Mount Scratchley (FST = 0.05), but low between Mount Wilhelm and Mount Scratchley (FST = 0.01). Accordingly, STRUCTURE suggested a scenario with K = 2 groups. For the remaining two species Sericornis nouhuysi showed some differentiation (FST = 0.03) between Mount Wilhelm and Huon and Aethomyias papuensis showed minor differentiation (FST = 0.02 between Mount Scratchley and Huon (Supplementary Table 2), but for both species, STRUCTURE suggested a scenario of K = 2 with considerable mixing of individuals between populations.Samples from Mount Wilhelm were collected at elevations ranging from 1700 to 3700 m, again allowing us to test for differences within populations on a single slope, a finding that would be consistent with incipient parapatric speciation. No species showed significant differences in FST when comparing individuals from different elevations, and concordantly there was little clustering of individuals by elevation in the PCA plots. Even when individuals were collected as far as 2000 elevational meters apart (as in the case of Origma robusta), genetic differentiation was low (FST = 0.01). In Huon, all samples were collected at the same elevation, except for Ifrita kowaldi, for which genetic differentiation of FST = 0.03 was found between individuals collected at 2300 m and 2950 m (Supplementary Fig. 3B, Supplementary Table 2). These analyses however, suffer from very small sample sizes that hinder a thorough analysis of parapatric speciation events. Furthermore, we note that divergence with gene flow may not manifest as a genome-wide phenomenon (at least, not until the taxa are so differentiated that gene flow has ceased). Instead, it may proceed via selection acting to create small ‘islands of differentiation’ within the genome against a background of negligible differentiation22,23. Such analyses require large sample sizes and are therefore not possible herein.Correlations between genetic divergence and elevationIf lineages colonize mountains from the lowlands, followed by range contraction and differentiation in the highlands, we would expect a signature of larger genetic differentiation (FST) between populations inhabiting higher elevations. We found no relationship between genetic differentiation (FST) and the altitudinal floor (the lowest elevation at which a species/population occurs) for the five Moluccan species, but for all New Guinean taxa with the exception of Melipotes fumigatus/ater we found a significant positive correlation (r = 0.83, p  More

  • in

    Local adaptation and colonization are potential factors affecting sexual competitiveness and mating choice in Anopheles coluzzii populations

    1.Kawecki, T. J. & Ebert, D. Conceptual issues in local adaptation. Ecol. Lett. 7, 1225–1241 (2004).
    Google Scholar 
    2.Fisher, T. W. et al. Handbook of Biological Control: Principles and Applications of Biological Control (Academic Press, London, 1999).
    Google Scholar 
    3.Dyck, V. A., Hendrichs, J. & Robinson, A. S. Sterile insect technique: Principles and practice in area-wide integrated pest management. In Sterile Insect Technique: Principles and Practice in Area-Wide Integrated Pest Management. https://doi.org/10.1007/1-4020-4051-2. (2005)4.Etges, W. J. & Noor, M. A. F. Genetics of Mate Choice: From Sexual Selection to Sexual Isolation. (Kluwer Academic Publishers, 2002).5.Harbach, R. E. Review of the internal classification of the genus Anopheles (Diptera: Culicidae): The foundation for comparative systematics and phylogenetic research. Bull. Entomol. Res. 84, 331–342 (1994).
    Google Scholar 
    6.Rogers, D. J., Randolph, S. E., Snow, R. W. & Hay, S. I. Satellite imagery in the study and forecast of malaria. Nature 415, 710–715 (2002).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    7.Bayoh, M. N. N., Thomas, C. J. J. & Lindsay, S. W. W. Mapping distributions of chromosomal forms of Anopheles gambiae in West Africa using climate data. Med. Vet. Entomol. 15, 267–274 (2001).CAS 
    PubMed 

    Google Scholar 
    8.Namountougou, M. et al. Multiple insecticide resistance in Anopheles gambiae s. l. Populations from Burkina Faso. West Africa. PLoS One 7, e48412 (2012).CAS 
    PubMed 
    ADS 

    Google Scholar 
    9.Benedict, M. Q. & Robinson, A. S. The first releases of transgenic mosquitoes: An argument for the sterile insect technique. Trends Parasitol. 19, 349–355 (2003).PubMed 

    Google Scholar 
    10.Maïga, H. et al. Mating competitiveness of sterile male Anopheles coluzzii in large cages. Malar. J. 13, 460 (2014).PubMed 
    PubMed Central 

    Google Scholar 
    11.Clements, A. N. The Biology of Mosquitoes. Sensory Reception and Behaviour Behaviour, Vol. 2. (Wallingford, 1999).12.Doug, P. et al. Genetic and environmental factors associated with laboratory rearing affect survival and assortative mating but not overall mating success in Anopheles gambiae Sensu Stricto. PLoS One 8, e82631 (2013).
    Google Scholar 
    13.Baeshen, R. et al. Differential effects of inbreeding and selection on male reproductive phenotype associated with the colonization and laboratory maintenance of Anopheles gambiae. Malar. J. 13, 19 (2014).PubMed 
    PubMed Central 

    Google Scholar 
    14.Bargielowski, I., Kaufmann, C., Alphey, L., Reiter, P. & Koella, J. Flight performance and teneral energy reserves of two genetically-modified and one wild-type strain of the yellow fever mosquito Aedes aegypti. Vector-Borne Zoonotic Dis. 12, 1053–1058 (2012).PubMed 
    PubMed Central 

    Google Scholar 
    15.Harris, A. F. et al. Field performance of engineered male mosquitoes. Nat. Biotechnol. 29, 1034–1037 (2011).CAS 
    PubMed 

    Google Scholar 
    16.Alphey, L. et al. Genetic control of Aedes mosquitoes. Pathogens and Global Health 107, 170–179 (2013).PubMed 
    PubMed Central 

    Google Scholar 
    17.Lehmann, T. et al. Tracing the origin of the early wet-season Anopheles coluzzii in the Sahel. Evol. Appl. 10, 704–717 (2017).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    18.Huestis, D. L. et al. Windborne long-distance migration of malaria mosquitoes in the Sahel. Nature 574, 404–408 (2019).CAS 
    PubMed 

    Google Scholar 
    19.Wondji, C., Simard, F. & Fontenille, D. Evidence for genetic differentiation between the molecular forms M and S within the Forest chromosomal form of Anopheles gambiae in an area of sympatry. Insect Mol. Biol. 11, 11–19 (2002).
    CAS 
    PubMed 

    Google Scholar 
    20.Simard, F., Nchoutpouen, E., Toto, J. C. & Fontenille, D. Geographic distribution and breeding site preference of Aedes albopictus and Aedes aegypti (Diptera: Culicidae) in Cameroon, Central Africa. J. Med. Entomol. 42, 726–731 (2005).PubMed 

    Google Scholar 
    21.Roux, O., Diabaté, A. & Simard, F. Divergence in threat sensitivity among aquatic larvae of cryptic mosquito species. J. Anim. Ecol. 83, 702–711 (2014).PubMed 

    Google Scholar 
    22.Costantini, C. et al. Living at the edge: Biogeographic patterns of habitat segregation conform to speciation by niche expansion in Anopheles gambiae. BMC Ecol. 9 (2009).23.The Anopheles gambiae 1000 Genomes Consortium. Genetic diversity of the African malaria vector Anopheles gambiae. Nature 552, 96–100 (2017).PubMed Central 

    Google Scholar 
    24.Oliva, C. F., Benedict, M. Q., Lempérière, G. & Gilles, J. Laboratory selection for an accelerated mosquito sexual development rate. Malar. J. 10, 135 (2011).PubMed 
    PubMed Central 

    Google Scholar 
    25.Munhenga, G. et al. Evaluating the potential of the sterile insect technique for malaria control: Relative fitness and mating compatibility between laboratory colonized and a wild population of Anopheles arabiensis from the Kruger National Park, South Africa. Parasit. Vectors 4, 208 (2011).PubMed 
    PubMed Central 

    Google Scholar 
    26.Lee, H. L. et al. Mating compatibility and competitiveness of transgenic and wild type Aedes aegypti (L.) under contained semi-field conditions. Transgenic Res. 22, 47–57 (2013).CAS 
    PubMed 

    Google Scholar 
    27.Damiens, D. et al. Cross-Mating compatibility and competitiveness among Aedes albopictus strains from distinct geographic origins-implications for future application of sit programs in the south west Indian ocean islands. PLoS One 11, e0163788 (2016).PubMed 
    PubMed Central 

    Google Scholar 
    28.Zheng, X. et al. Incompatible and sterile insect techniques combined eliminate mosquitoes. Nature 572, 56–61 (2019).CAS 
    ADS 

    Google Scholar 
    29.Aguilar, R. et al. Genome-wide analysis of transcriptomic divergence between laboratory colony and field Anopheles gambiae mosquitoes of the M and S molecular forms. Insect Mol. Biol. 19, 695–705 (2010).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    30.Sawadogo, P. S. et al. Swarming behaviour in natural populations of Anopheles gambiae and An. coluzzii: Review of 4 years survey in rural areas of sympatry, Burkina Faso (West Africa). Acta Trop. 130, 24–34 (2014).
    Google Scholar 
    31.Poda, S. B. et al. Sex aggregation and species segregation cues in swarming mosquitoes: Role of ground visual markers. Parasit. Vectors 12, 589 (2019).PubMed 
    PubMed Central 

    Google Scholar 
    32.Ekechukwu, N. E. et al. Heterosis increases fertility, fecundity, and survival of laboratory-produced F1 hybrid males of the malaria mosquito Anopheles coluzzii. G3 Genes Genomes Genet. 5, 2693–2709 (2015).CAS 

    Google Scholar 
    33.Ng’habi, K. R. et al. Colonization of malaria vectors under semi-field conditions as a strategy for maintaining genetic and phenotypic similarity with wild populations. Malar. J. 14, 10 (2015).PubMed 
    PubMed Central 

    Google Scholar 
    34.Huho, B. J. et al. Nature beats nurture: A case study of the physiological fitness of free-living and laboratory-reared male Anopheles gambiae s.l. J. Exp. Biol. 210, 2939–2947 (2007).CAS 
    PubMed 

    Google Scholar 
    35.Ferguson, H. M., John, B., Ng, K. & Knols, B. G. J. Redressing the sex imbalance in knowledge of vector biology. Trends Ecol. Evol. 20, 202–209 (2005).PubMed 

    Google Scholar 
    36.Hassan, M., El-Motasim, W. M., Ahmed, R. T. & El-Sayed, B. B. Prolonged colonisation, irradiation, and transportation do not impede mating vigour and competitiveness of male Anopheles arabiensis mosquitoes under semi-field conditions in Northern Sudan. Malar. World J. 1 (2010).37.Yamada, H., Vreysen, M. J. B., Gilles, J. R. L., Munhenga, G. & Damiens, D. D. The effects of genetic manipulation, dieldrin treatment and irradiation on the mating competitiveness of male Anopheles arabiensis in field cages. Malar. J. 13, 1–10 (2014).
    Google Scholar 
    38.Munhenga, G. et al. Mating competitiveness of sterile genetic sexing strain males (GAMA) under laboratory and semi-field conditions : Steps towards the use of the Sterile Insect Technique to control the major malaria vector Anopheles arabiensis in South Africa. Parasit. Vectors 9, 1–12 (2016).
    Google Scholar 
    39.Assogba, B. S. et al. Characterization of swarming and mating behaviour between Anopheles coluzzii and Anopheles melas in a sympatry area of Benin. Acta Trop. 132S, 1–11 (2013).
    Google Scholar 
    40.Charlwood, J. D. et al. The swarming and mating behaviour of Anopheles gambiae s.s. (Diptera: Culicidae) from São Tomé Island. J. Vector Ecol. 27, 178–183 (2002).CAS 
    PubMed 

    Google Scholar 
    41.Diabate, A. et al. Natural swarming behaviour of the molecular M form of Anopheles gambiae. Trans. R. Soc. Trop. Med. Hyg. 97, 713–716 (2003).CAS 
    PubMed 

    Google Scholar 
    42.Manoukis, N. C. et al. Structure and dynamics of male swarms of Anopheles gambiae. J. Med. Entomol. 46, 227–235 (2009).PubMed 

    Google Scholar 
    43.Aldersley, A. et al. Too ‘sexy’ for the field? Paired measures of laboratory and semi-field performance highlight variability in the apparent mating fitness of Aedes aegypti transgenic strains. Parasit. Vectors 12, 357 (2019).PubMed 
    PubMed Central 

    Google Scholar 
    44.Pantoja-Sánchez, H., Gomez, S., Velez, V., Avila, F. W. & Alfonso-Parra, C. Precopulatory acoustic interactions of the New World malaria vector Anopheles albimanus (Diptera: Culicidae). Parasit. Vectors 12, 386 (2019).PubMed 
    PubMed Central 

    Google Scholar 
    45.Gibson, G., Warren, B. & Russell, I. J. Humming in tune: Sex and species recognition by mosquitoes on the wing. JARO 540, 527–540 (2010).
    Google Scholar 
    46.Pennetier, C., Warren, B., Dabiré, K. R., Russell, I. J. & Gibson, G. ‘Singing on the wing’ as a mechanism for species recognition in the malarial mosquito Anopheles gambiae. Curr. Biol. 20, 131–136 (2010).CAS 
    PubMed 

    Google Scholar 
    47.Caputo, B. et al. Comparative analysis of epicuticular lipid profiles of sympatric and allopatric field populations of Anopheles gambiae s.s. molecular forms and An. arabiensis from Burkina Faso (West Africa). Insect Biochem. Mol. Biol. 37, 389–398 (2007).CAS 
    PubMed 

    Google Scholar 
    48.Ferguson, H. M. & Read, A. F. Genetic and environmental determinants of malaria parasite virulence in mosquitoes. Proc. R. Soc. B Biol. Sci. 269, 1217–1224 (2002).CAS 

    Google Scholar 
    49.Niang, A. et al. Semi-field and indoor setups to study malaria mosquito swarming behavior. Parasit. Vectors 12, 446 (2019).PubMed 
    PubMed Central 

    Google Scholar 
    50.Santolamazza, F. et al. Insertion polymorphisms of SINE200 retrotransposons within speciation islands of Anopheles gambiae molecular forms. Malar. J. 7, 163 (2008).
    PubMed 
    PubMed Central 

    Google Scholar 
    51.Vantaux, A. et al. Larval nutritional stress affects vector life history traits and human malaria transmission. Sci. Rep. 6, 36778 (2016).CAS 
    PubMed 
    PubMed Central 
    ADS 

    Google Scholar 
    52.Crawley, M. J. The R Book. (Ltd, Sons, 2012). https://doi.org/10.1002/9780470515075.53.Hothorn, T., Bretz, F., Westfall, P. & Heiberger, R. M. Package ‘multcomp’ title simultaneous inference in general parametric models. Biom. J. 50, 346–363 (2016).
    Google Scholar  More

  • in

    Newly identified HMO-2011-type phages reveal genomic diversity and biogeographic distributions of this marine viral group

    General characterization of seven newly isolated HMO-2011-type phagesIn this study, we used four Roseobacter strains (FZCC0040, FZCC0042, FZCC0012, and FZCC0089) and one SAR11 strain (HTCC1062) to isolate phages. FZCC0040 and FZCC0042 belong to the Roseobacter RCA lineage [22], FZCC0012 shares 99.8% 16S rRNA gene identity with Roseobacter strain HIMB11 [57], and FZCC0089 belongs to a newly identified Roseobacter lineage located close to HIMB11 and SAG-019 lineages (Supplementary Fig. 1).A total of seven phages were newly isolated and analyzed in this study (Table 1). The complete phage genomes range in size from 52.7 to 54.9 kb, harbor 62 to 84 open reading frames (ORFs), and feature a G + C content ranging from 33.8 to 48.6%. Compared to other HMO-2011-type phages, pelagiphage HTVC033P has a relatively lower G + C content of 33.8%, similar to the G + C content of its host HTCC1062 (29.0%) and of other described pelagiphages [21, 26,27,28]. The G + C content of other six roseophages ranges from 42.2 to 48.6%, which is also similar to the G + C content of the hosts they infect (44.8 to 54.1%).Despite their distinct host origins, these phage genomes show considerable similarity in terms of gene content and genome architecture (Fig. 1). They all display clear similarity with the previously reported SAR116 phage HMO-2011 [20] and HMO-2011-type RCA phages [22]. Overall, these phages share 19.2 to 79.1% of their genes with previously reported HMO-2011-type phages and all contain homologues of HMO-2011-type DNA replication and metabolism genes, structural genes, and DNA packaging genes. Moreover, their overall genome structure is conserved with that of HMO-2011-type phages. Considering these observations, we tentatively classified these seven phages into the HMO-2011-type group. Of the 11 currently known HMO-2011-type isolates, one infects the SAR116 strain IMCC1322, one infects the SAR11 strain HTCC1062, and the remaining nine all infect Roseobacter strains; this suggest that HMO-2011-type phages infect diverse bacterial hosts. HTVC033P is the first pelagiphage identified to belong to the HMO-2011-type viral group. Our study has also increased the number of known types of pelagiphages. To date, pelagiphages belonging to a total of nine distinct viral groups have been isolated and analyzed [21, 26,27,28].Fig. 1: Alignment and comparison of genomes of HMO-2011-type isolates and representative HMO-2011-type MVGs from major subgroups.HMO-2011-type phage isolates are shown in red. Phages isolated in this study are indicated with red asterisks. Predicted open reading frames (ORFs) are represented by arrows, with the left or right arrow points indicating the direction of their transcription. The numbers inside the arrows indicate ORF numbers. ORFs annotated with known functions are marked using distinct colors according to their functions. HMO-2011-type core genes are indicated with blue asterisks. The color of the shading connecting homologous genes indicates the level of amino acid identity between the genes. To clearly present the genomic comparison, several MVGs were rearranged to start from the same gene as in the HMO-2011-type phages. DNAP DNA polymerase, Endo endonuclease, RNR ribonucleoside-triphosphate reductase, PhoH phosphate starvation-inducible protein, MazG MazG nucleotide pyrophosphohydrolase domain protein, ThyX thymidylate synthase, GRX glutaredoxin, TerS terminase small subunit, TerL terminase large subunit.Full size imageIdentification and sequence analyses of HMO-2011-type MVGsTo identify HMO-2011-type MVGs, we performed a metagenomic mining and retrieved a total of 207 HMO-2011-type MVGs (≥50% genome completeness) from viromes in the worldwide ocean, from tropical to polar oceans (Supplementary Table 1). These MVGs range in size from 29.2 to 67.9 kb and their G + C content range from 31.3 to 52.4%. In addition, 45 HMO-2011-type MVGs were also identified from some non-marine habitats, suggesting that HMO-2011-type phages are widely distributed worldwide (Supplementary Table 1).Genomic analysis confirmed that all HMO-2011-type MVGs exhibit genomic synteny with HMO-2011-type phages (Fig. 1). Although some of these HMO-2011-type MVGs are highly similar to their cultivated relatives, most MVGs appear to have more genomic variations. To resolve the evolutionary relationship among the HMO-2011-type phages, a phylogenetic tree was constructed based on the concatenated sequences of five core genes. We found that HMO-2011-type phages are evolutionarily diverse and can be separated into at least 10 well-supported subgroups ( >2 members), with 140 MVGs clustering into previously identified HMO-2011-type groups (subgroups I and III in Fig. 2A) [22], and the remaining 67 MVGs forming new subgroups (Fig. 2A). Among these HMO-2011-type subgroups, three contain cultivated representatives (subgroups I, III, and IX). Subgroup I contains the greatest number of phages, including six cultivated representatives and 123 MVGs (Fig. 2A). The cultivated representatives in subgroup I include a phage that infect SAR116 strain and five phages that infect Roseobacter strains. Subgroup III contains four cultivated representatives that infect two Roseobacter strains, and 17 MVGs. Pelagiphage HTVC033P and nine MVGs form subgroup IX. Other subgroups have no cultivated representatives yet. The results of phylogenomic analysis showed that subgroups I to VI are closely related, whereas subgroups VII to X are located on a separate branch and are more distinct from the subgroups I to VI, which suggests that these subgroups are more evolutionarily distant. A phylogenomic-based approach with GL-UVAB workflow [53] was also performed to cluster these HMO-2011-type genomes, which showed similar grouping results (Supplementary Fig. 2).Fig. 2: Phylogenomic and shared-gene analyses of HMO-2011-type phages.A A maximum-likelihood tree was constructed using concatenated sequences of five hallmark genes. HMO-2011-type phages were grouped into 10 subgroups based on the phylogeny. Shading is used to indicate the subgroups. HMO-2011-type phage isolates are shown in red. Genomes containing an integrase gene are indicated by red triangles. The G + C content and completeness of the genomes are indicated. Scale bar indicates the number of amino acid substitutions per site. B Heatmap showing the percentage of shared genes between HMO-2011-type genomes. Phages in the same subgroup are boxed.Full size imageA previous study suggested the use of the percentage of shared proteins as a means of defining phage taxonomic ranks and proposed that phages with ≥20 and ≥40% orthologous proteins in common can be grouped at the taxonomic ranks of subfamily and genus, respectively [58]. Overall, most of the calculated percentages between HMO-2011-type genomes fall within the 20 to 100% range and most of the percentages between genomes within the same subgroup fall within the 40 to 100% range (Fig. 2B). Therefore, our results suggest that the HMO-2011-type is roughly a subfamily-level phage taxonomic group containing at least ten genus-level subgroups in the Podoviridae family.Conserved genomic structure and variation in HMO-2011-type phagesOf the 1235 orthologous protein groups (≥2 members) identified in HMO-2011-type genomes, only 254 proteins groups could be assigned putative biological functions (Supplementary Table 2). Comparative genomic analysis clearly revealed the conserved functional module structure of all HMO-2011-type genomes. All HMO-2011-type phage genomes can be roughly divided into the DNA metabolism and replication module, structural module and DNA packaging module (Fig. 1). Most of the homologous genes are scattered in similar loci of the HMO-2011-type genomes. Core genome analysis based on complete HMO-2011-type genomes revealed that HMO-2011-type genomes share a common set of ten core genes (Fig. 1). These core genes are mostly genes related to essential function in phage replication and development, including genes encoding DNA helicase, DNA primase, DNA polymerase (DNAP), portal protein, capsid protein, and terminase small and large subunits (TerL and TerS) as well as several genes with no known function, suggesting that phages in this group employ similar overall infection and propagation processes (Fig. 1).Most members in subgroups I and III and one member in subgroup II possess a tyrosine integrase gene (int) located upstream of the DNA replication and metabolism module, whereas all subgroup IV to X genomes contain no identifiable lysogeny-related genes. This result suggests that members of subgroups IV to X might be obligate lytic phages. Integrase genes typically occur in the genomes of temperate phages and are responsible for site-specific recombination between phage and host bacterial genomes [59, 60]. In subgroup III, RCA phage CRP-3 has been experimentally demonstrated to be capable of integrating into the host genome [22]. Thus, certain int-containing HMO-2011-type phages are also likely to be temperate phages.In the DNA metabolism and replication modules, genes encoding DNA primase, DNA helicase, DNAP, ribonucleotide reductase (RNR), and endonuclease can be identified; and DNA helicase, DNA primase, and DNAP are core to all HMO-2011-type phages. All reported HMO-2011-type phages contain an atypical DNAP, in which a partial DnaJ central domain is located between the exonuclease domain and the DNA polymerase domain [20, 22]. The Escherichia coli DnaJ protein, a co-chaperone [61], has been shown to be involved in diverse functions [62] and to be critical for the replication of phage Lambda [63,64,65]. The sequence analysis revealed that DNAP sequences of these seven new HMO-2011-type phages and 207 MVGs also present this unusual domain structure and contain two repeats of the CXXCXGXG motifs involved in zinc binding [66] in the partial DnaJ domain (Supplementary Fig. 3). RNR gene is frequently detected in subgroups I, II, III, IV, V, and X genomes but not in the other subgroup genomes. RNRs, which are widely distributed in diverse phage genomes, are involved in catalyzing the reduction of ribonucleotides to deoxyribonucleotides, and thus play a crucial role in providing deoxyribonucleoside triphosphates for phage DNA biosynthesis and repair [67,68,69]. RNR genes clustered with the RNR gene in phage HMO-2011 were previously reported to dominate the class II viral RNRs in examined marine viromes [69]. In the remaining two modules, genes involved in phage structure (e.g., genes encoding capsid and portal proteins), packaging of DNA (TerL and TerS genes), and cell lysis were detected. The proteins encoded by these genes play key roles in phage morphogenesis and virion release.Examination of the distribution of the orthologous groups among the subgroups revealed clear pan-genome differences in various subgroups (Fig. 3). Most subgroups harbor subgroup-specific genes not identified in other subgroups, although  no function has yet been assigned to most of these genes. Notably, the phages in subgroups VII, VIII, and IX possess genomic features that differentiate them from phages in other subgroups, specifically with regard to the G + C content and gene content. The members of these three subgroups are closely related to each other in the phylogenetic tree and harbor several subgroup-specific genes. The G + C content of the phage genomes in these subgroups ranges from 31.9 to 35.4%, significantly smaller than other subgroups but similar to the G + C content of SAR11 bacteria and other known pelagiphages. HTVC033P is the only cultivated representative of subgroup IX. The aforementioned results suggest that the phages in subgroup VII, VIII, and IX might have related bacterial hosts and are highly likely to be pelagiphages. The host prediction using RaFAH tool also assigned Pelagibacter as their potential hosts (Supplementary Table 1). Subgroup X is located near these three subgroups in the phylogenetic tree, and the G + C content of the phages in this subgroup ranges from 34.4 to 39.0%. The host prediction assigned Roseobacter as their potential hosts. The hosts of this subgroup still remain to be experimentally investigated.Fig. 3: Distribution and functional classification of orthologous protein groups across HMO-2011-type genomes.Only orthogroups containing >10 members or showing subgroup-specific features are shown. Subgroup-specific genes are boxed in red. Genes that are absent in a specific subgroup are boxed in orange.Full size imageMetabolic capabilities of HMO-2011-type phagesAll HMO-2011-type phage genomes harbor several host-derived auxiliary metabolic genes (AMGs) potentially involved in diverse metabolic processes. Some AMGs in HMO-2011-type phages have been discussed previously [20, 22].Subgroups VII, VIII, IX, and X possess distinct AMGs as compared with the other subgroups. For example, the genes encoding FAD-dependent thymidylate synthase (ThyX, PF02511) and MazG pyrophosphohydrolase domains are absent in all subgroups VII, VIII, IX, and X genomes but frequently detected in other subgroup genomes. ThyX protein is essential for the conversion of dUMP to dTMP mediated by an FAD coenzyme and is therefore a key enzyme involved in DNA synthesis [70, 71]. The thyX gene is commonly found in microbial genomes and phage genomes. Phage-encoded ThyX has been suggested to compensate for the loss of host-encoded ThyA and thus play crucial roles in phage nucleic acid synthesis and metabolism during infection [72]. Except in the case of subgroups VII, VIII, IX, and X genomes, the mazG gene, which encodes a nucleoside triphosphate pyrophosphohydrolase is sporadically distributed in HMO-2011-type genomes. MazG protein is predicted to be a regulator of nutrient stress and programmed cell death [73] and has been hypothesized to promote phage survival by keeping the host alive during phage propagation [74]. The Escherichia coli MazG can interfere with the function of the MazEF toxin–antitoxin system by decreasing the cellular level of (p)ppGpp [73]. However, a recent study showed that a cyanophage MazG has no binding or hydrolysis activity against alarmone (p)ppGpp but has high hydrolytic activity toward dGTP and dCTP, and it was speculated to play a role in hydrolyzing high G + C host genome for phage replication [75]. Whether the MazG proteins encoded by HMO-2011-type phages play a similar role in phage propagation remained to be investigated.Five MVGs in subgroup I contain a gene encoding a DraG-like family ADP-ribosyl hydrolase (ARH). In cellular ADP-ribosylation systems, ARH catalyzes the cleavage of the ADP-ribose moiety, and thereby counteract the effects of ADP-ribosyl transferases [76]. It has been reported that ARH in Rhodospirillum rubrum regulates the nitrogen fixation [77]. However, the function of this phage-encoded ARH in the phage propagation process remains unclear.We also observed that several MVGs possess genes involved in iron–sulfur (Fe–S) cluster biosynthesis, including an Fe–S cluster assembly scaffold gene (iscU) that involved in Fe–S cluster assembly and transfer [78] and an Fe–S cluster insertion protein gene (erpA). Fe–S cluster participates in a wide variety of cellular biological processes [79]. The discovery of these genes suggests that these phages may play important roles in Fe–S cluster biogenesis and function.The gene encoding sodium-dependent phosphate transport protein (PF02690) has been identified in eight subgroup I genomes. The Na/Pi cotransporter family protein is responsible for high-affinity, sodium-dependent Pi uptake, and thus the protein plays a critical role in maintaining phosphate homeostasis [80]. This gene might function in the transport of phosphate into cells during phage infection. The presence of Na/Pi cotransporter genes suggests that some HMO-2011-type phages may have the potential to regulate host phosphate uptake in phosphate-limited ocean environments in order to benefit phage replication and propagation.Identification and phylogenetic analysis of HMO-2011-type DNAPsThe genetic diversity and geographically distribution of HMO-2011-type phages in marine environments was further inferred from DNAP gene analyses. A total of 2433 HMO-2011-type DNAP sequences with sequence sizes ranging from 540 to 779 amino acids were identified and subjected to phylogenetic analysis (Supplementary Table 3).Among the identified HMO-2011-type DNAPs, 2030 sequences were retrieved from the GOV 2.0 Tara expedition upper-ocean viral populations (0–1000 m), from tropical to polar regions. HMO-2011-type DNAP genes were identified from all analyzed upper-ocean viromes, suggesting the global prevalence of HMO-2011-type phages in upper oceans.A previous study revealed that marine viromes contain various types of tailed phage genomes that encode a family A DNAP gene [81]. To estimate the importance of HMO-2011-type phages, we calculated the proportion of HMO-2011-type DNAPs based on the number of HMO-2011-type DNAP sequences and the total number of family A DNAP sequences ( >470 aa) in each GOV 2.0 viral population dataset. This analysis revealed that HMO-2011-type DNAPs accounted for up to 19.7% of all family A DNAPs in each GOV 2.0 dataset (Supplementary Table 4). We found that the HMO-2011-type DNAP sequences appear to be more dominant in epipelagic viromes than in mesopelagic viromes (p  More

  • in

    Potential distribution of fall armyworm in Africa and beyond, considering climate change and irrigation patterns

    Research model and softwareCLIMEX modelFAW growth and development are primarily related to climate conditions, especially temperature patterns17. The current study used CLIMEX (version 4)42, a semi-mechanistic niche modeling platform, to project FAW distribution in relation to climate. The model parameters that describe the species’ response to climate were overlaid onto FAW occurrence data and climate data to project the species’ potential global distribution. Briefly, the annual growth index (GI) was used to describe the potential for FAW population growth during favorable climatic conditions, while stress indices (SI: cold, wet, hot, and dry) and interaction stresses (SX: hot-dry, hot-wet, cold-dry, and cold-wet) (Table 1) were applied to describe the probability that FAW populations could survive unfavorable conditions. The Ecoclimatic index (EI) was derived from a combination of GI, SI, and SX indices to provide an overall annual index of climatic suitability on a scale of 0–10042. An EI value of 0 indicates that the location is not suitable for the long-term survival of the species, whereas an EI value of 100 indicates maximum climatic suitability comparable to conditions in incubators. EI values of more than 30 indicate the optimal climate for a species. In this study, the climatic suitability was classified into four arbitrary categories; unsuitable for EI = 0, marginal for 0  More

  • in

    Physical geography, isolation by distance and environmental variables shape genomic variation of wild barley (Hordeum vulgare L. ssp. spontaneum) in the Southern Levant

    Abdel-Ghani AH, Parzies HK, Omary A, Geiger HH (2004) Estimating the outcrossing rate of barley landraces and wild barley populations collected from ecologically different regions of Jordan Theor Appl Genet 109(3):588–595PubMed 

    Google Scholar 
    Akerman A, Bürger R (2014) The consequences of gene flow for local adaptation and differentiation: a two-locus two-deme model J Math Biol 68(5):1135–1198PubMed 

    Google Scholar 
    Al-Asadi H, Petkova D, Stephens M, Novembre J (2019) Estimating recent migration and population-size surfaces PLoS Genet 15(1):e1007908PubMed 
    PubMed Central 

    Google Scholar 
    Baker HG (1967) Support for Baker’s law-as a rule Evolution 21(4):853–856PubMed 

    Google Scholar 
    Baker K, Baker K, Bayer M, Cook N, Dreißig S, Dhillon T, Russell J, Hedley PE, Morris J, Ramsay L, Colas I et al. (2014) The low-recombining pericentromeric region of barley restricts gene diversity and evolution but not gene expression Plant J 79(6):981–992CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Battey C, Ralph PL, Kern AD (2020) Space is the place: effects of continuous spatial structure on analysis of population genetic data Genetics 215(1):193–214CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Baum M, Grando S, Backes G, Jahoor A, Sabbagh A, Ceccarelli S (2003) QTLs for agronomic traits in the mediterranean environment identified in recombinant inbred lines of the cross’ Arta’ × H. spontaneum 41-1 Theor Appl Genet 107(7):1215–1225CAS 
    PubMed 

    Google Scholar 
    Bedada G, Westerbergh A, Nevo E, Korol A, Schmid KJ (2014) DNA sequence variation of wild barley Hordeum spontaneum (L.) across environmental gradients in Israel Heredity 112(6):646–655CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Berner D, Roesti M (2017) Genomics of adaptive divergence with chromosome-scale heterogeneity in crossover rate Mol Ecol 26(22):6351–6369CAS 
    PubMed 

    Google Scholar 
    Bhatia G, Patterson N, Sankararaman S, Price AL (2013) Estimating and interpreting FST: the impact of rare variants Genome Res 23(9):1514–1521CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bohra A, Kilian B, Kilian B, Sivasankar S, Caccamo M, Mba, C, McCouch SR, Varshney RK (2021) Reap the crop wild relatives for breeding future crops. Trends Biotechnol. https://doi.org/10.1016/j.tibtech.2021.08.009Bradburd GS, Coop GM, Ralph PL (2018) Inferring continuous and discrete population genetic structure across space. Genetics 210(1):33–52PubMed 
    PubMed Central 

    Google Scholar 
    Bürger R, Akerman A (2011) The effects of linkage and gene flow on local adaptation: a two-locus continent–island model. Theor Popul Biol 80(4):272–288PubMed 
    PubMed Central 

    Google Scholar 
    Cabreros I, Storey JD (2019) A likelihood-free estimator of population structure bridging admixture models and principal components analysis. Genetics 212(4):1009–1029PubMed 
    PubMed Central 

    Google Scholar 
    Caldwell KS, Russell J, Langridge P, Powell W (2006) Extreme population-dependent linkage disequilibrium detected in an inbreeding plant species, Hordeum vulgare. Genetics 172(1):557–567CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Capblancq T, Luu K, Blum MG, Bazin E (2018) Evaluation of redundancy analysis to identify signatures of local adaptation. Mol Ecol Resour 18(6):1223–1233CAS 
    PubMed 

    Google Scholar 
    Caye K, Jumentier B, Lepeule J, François O (2019) LFMM 2: fast and accurate inference of gene-environment associations in genome-wide studies. Mol Biol Evol 36(4):852–860CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Contreras-Moreira B, Serrano-Notivoli R, Mohammed NE, Cantalapiedra CP, Beguería S, Casas AM, Igartua E (2019) Genetic association with high-resolution climate data reveals selection footprints in the genomes of barley landraces across the Iberian Peninsula. Mol Ecol 28(8):1994–2012PubMed 
    PubMed Central 

    Google Scholar 
    Dawson IK, Russell J, Powell W, Steffenson B, Thomas WTB, Waugh R (2015) Barley: a translational model for adaptation to climate change. New Phytol 206(3):913–931PubMed 

    Google Scholar 
    Dray S, Legendre P, Peres-Neto PR (2006) Spatial modelling: a comprehensive framework for principal coordinate analysis of neighbour matrices (pcnm). Ecol Model 196(3-4):483–493
    Google Scholar 
    Dray S, Bauman D, Blanchet G, Borcard D, Clappe S, Guenard G, Jombart T, Larocque G, Legendre P, Madi N, Wagner HH (2019) adespatial: multivariate multiscale spatial analysis. R package version 0.3-7. https://CRAN.R-project.org/package=adespatialElshire RJ, Glaubitz JC, Sun Q, Poland JA, Kawamoto K, Buckler ES, Mitchell SE (2011) A robust, simple genotyping-by-sequencing (GBS) approach for high diversity species. PLoS ONE 6(5):e19379CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Excoffier L, Hofer T, Foll M (2009) Detecting loci under selection in a hierarchically structured population. Heredity 103(4):285–298CAS 
    PubMed 

    Google Scholar 
    Fang Z, Gonzales AM, Clegg MT, Smith KP, Muehlbauer GJ, Steffenson BJ, Morrell PL (2014) Two genomic regions contribute disproportionately to geographic differentiation in wild barley. G34(7):1193–1203PubMed 
    PubMed Central 

    Google Scholar 
    Fick SE, Hijmans RJ (2017) Worldclim 2: new 1-km spatial resolution climate surfaces for global land areas. Int J Climatol 37(12):4302–4315
    Google Scholar 
    Forester BR, Lasky JR, Wagner HH, Urban DL (2018) Comparing methods for detecting multilocus adaptation with multivariate genotype–environment associations. Mol Ecol 27(9):2215–2233CAS 
    PubMed 

    Google Scholar 
    Forester BR, Jones MR, Joost S, Landguth EL, Lasky JR (2016) Detecting spatial genetic signatures of local adaptation in heterogeneous landscapes. Mol Ecol 25(1):104–120CAS 
    PubMed 

    Google Scholar 
    Galkin E, Dalal A, Evenko A, Fridman E, Kan I, Wallach R, Moshelion M (2018) Risk-management strategies and transpiration rates of wild barley in uncertain environments. Physiol Plant 164(4):412–428CAS 
    PubMed 

    Google Scholar 
    Gautier M (2015) Genome-wide scan for adaptive divergence and association with population-specific covariates. Genetics 201(4):1555–1579CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Gibson MJ, Moyle LC (2020) Regional differences in the abiotic environment contribute to genomic divergence within a wild tomato species. Mol Ecol 29(12):2204–2217CAS 
    PubMed 

    Google Scholar 
    Günther T, Coop G (2013) Robust identification of local adaptation from allele frequencies. Genetics 195(1):205–220PubMed 
    PubMed Central 

    Google Scholar 
    Gutaker RM, Groen SC, Bellis ES, Choi JY, Pires IS, Bocinsky RK, Slayton ER, Wilkins O, Castillo CC, Negrão S et al. (2020) Genomic history and ecology of the geographic spread of rice. Nat Plants 6(5):492–502PubMed 

    Google Scholar 
    Hämälä T, Savolainen O (2019) Genomic patterns of local adaptation under gene flow in arabidopsis lyrata. Mol Biol Evol 36(11):2557–2571
    Google Scholar 
    Harlan JR, Zohary D (1966) Distribution of wild wheats and barley. Science 153(3740):1074–1080CAS 
    PubMed 

    Google Scholar 
    Hartfield M, Bataillon T, Glémin S (2017) The evolutionary interplay between adaptation and self-fertilization. Trends Genet 33(6):420–431CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hendrick MF, Finseth FR, Mathiasson ME, Palmer KA, Broder EM, Breigenzer P, Fishman L (2016) The genetics of extreme microgeographic adaptation: an integrated approach identifies a major gene underlying leaf trichome divergence in yellowstone mimulus guttatus. Mol Ecol 25(22):5647–5662CAS 
    PubMed 

    Google Scholar 
    Hengl T, Mendes de Jesus J, Heuvelink GBM, Ruiperez Gonzalez M, Kilibarda M, Blagotić A, Shangguan W, Wright MN, Geng X, Bauer-Marschallinger B et al. (2017) Soilgrids250m: Global gridded soil information based on machine learning. PLoS ONE 12(2):e0169748PubMed 
    PubMed Central 

    Google Scholar 
    Herzig P, Herzig P, Maurer A, Draba V, Sharma R, Draicchio F, Bull H, Milne L, Thomas WTB, Flavell AJ, Pillen K (2018) Contrasting genetic regulation of plant development in wild barley grown in two European environments revealed by nested association mapping. J Exp Bot 69(7):1517–1531CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hill W, Weir B (1988) Variances and covariances of squared linkage disequilibria in finite populations. Theor Popul Biol 33(1):54–78CAS 
    PubMed 

    Google Scholar 
    Hodgins KA, Yeaman S (2019) Mating system impacts the genetic architecture of adaptation to heterogeneous environments. New Phytol 224(3):1201–1214PubMed 

    Google Scholar 
    House GL, Hahn MW (2018) Evaluating methods to visualize patterns of genetic differentiation on a landscape. Mol Ecol Resour 18(3):448–460PubMed 

    Google Scholar 
    Hübner S, Korol AB, Schmid KJ (2015) Rna-seq analysis identifies genes associated with differential reproductive success under drought-stress in accessions of wild barley hordeum spontaneum. BMC Plant Biol15(1):1–14
    Google Scholar 
    Hübner S, Bdolach E, Ein-Gedy S, Schmid KJ, Korol A, Fridman E (2013) Phenotypic landscapes: phenological patterns in wild and cultivated barley. J Evol Biol 26(1):163–174PubMed 

    Google Scholar 
    Hübner S, Günther T, Flavell A, Fridman E, Graner A, Korol A, Schmid KJ (2012) Islands and streams: clusters and gene flow in wild barley populations from the Levant. Mol Ecol 21(5):1115–1129PubMed 

    Google Scholar 
    Hübner S, Höffken M, Oren E, Haseneyer G, Stein N, Graner A, Schmid K, Fridman E (2009) Strong correlation of wild barley (Hordeum spontaneum) population structure with temperature and precipitation variation. Mol Ecol 18(7):1523–1536PubMed 

    Google Scholar 
    Jakob SS, Rödder D, Engler JO, Shaaf S, Özkan H, Blattner FR, Kilian B (2014) Evolutionary history of wild barley (Hordeum vulgare subsp. spontaneum) analyzed using multilocus sequence data and paleodistribution modeling. Genome Biol Evol 6(3):685–702CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Jayakodi M, Padmarasu S, Haberer G, Bonthala VS, Gundlach H, Monat C, Lux T, Kamal N, Lang D, Himmelbach A, Ens J, Zhang X-Q, Angessa TT, Zhou G, Tan C, Hill C, Wang P, Schreiber M, Boston LB, Plott C, Jenkins J, Guo Y, Fiebig A, Budak H, Xu D, Zhang J, Wang C, Grimwood J, Schmutz J, Guo G, Zhang G, Mochida K, Hirayama T, Sato K, Chalmers KJ, Langridge P, Waugh R, Pozniak CJ, Scholz U, Mayer KFX, Spannagl M, Li C, Mascher M, Stein N (2020) The barley pan-genome reveals the hidden legacy of mutation breeding Nature 588:284–289. https://doi.org/10.1038/s41586-020-2947-8CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kawecki TJ, Ebert D (2004) Conceptual issues in local adaptation. Ecol Lett 7(12):1225–1241
    Google Scholar 
    Kilian B, Özkan H, Kohl J, von Haeseler A, Barale F, Deusch O, Brandolini A, Yucel C, Martin W, Salamini F (2006) Haplotype structure at seven barley genes: relevance to gene pool bottlenecks, phylogeny of ear type and site of barley domestication. Mol Genet Genom 276(3):230–241CAS 

    Google Scholar 
    Lasky JR, Des Marais DL, McKAY JK, Richards JH, Juenger TE, Keitt TH (2012) Characterizing genomic variation of arabidopsis thaliana: the roles of geography and climate. Mol Ecol 21(22):5512–5529PubMed 

    Google Scholar 
    Lasky JR, Upadhyaya HD, Ramu P, Deshpande S, Hash CT, Bonnette J, Juenger TE, Hyma K, Acharya C, Mitchell SE et al. (2015) Genome-environment associations in sorghum landraces predict adaptive traits. Sci Adv 1(6):e1400218PubMed 
    PubMed Central 

    Google Scholar 
    Lawson DJ, Van Dorp L, Falush D (2018) A tutorial on how not to over-interpret STRUCTURE and ADMIXTURE bar plots. Nat Commun 9(1):1–11CAS 

    Google Scholar 
    Lee C-R, Mitchell-Olds T (2011) Quantifying effects of environmental and geographical factors on patterns of genetic differentiation. Mol Ecol 20(22):4631–4642PubMed 
    PubMed Central 

    Google Scholar 
    Leek JT (2011) Asymptotic conditional singular value decomposition for high-dimensional genomic data. Biometrics 67(2):344–352PubMed 

    Google Scholar 
    Legendre P, Legendre L (2012) Canonical analysis. In: Numerical ecology, 3rd English edn, chap. 11. Elsevier Science BV, The Netherlands, pp 625–710López-Goldar X, Agrawal AA (2021) Ecological interactions, environmental gradients, and gene flow in local adaptation Trends Plant Sci 26(8):796–809PubMed 

    Google Scholar 
    Lotterhos KE, Whitlock MC (2015) The relative power of genome scans to detect local adaptation depends on sampling design and statistical method. Mol Ecol 24(5):1031–1046PubMed 

    Google Scholar 
    Lundgren E, Ralph PL (2019) Are populations like a circuit? Comparing isolation by resistance to a new coalescent-based method. Mol Ecol Resour 19(6):1388–1406PubMed 

    Google Scholar 
    Makowski D, Ben-Shachar M, Lüdecke D (2019) bayestestR: describing effects and their uncertainty, existence and significance within the Bayesian framework. J Open Source Softw 4(40):1541
    Google Scholar 
    Mascher M, Gundlach H, Himmelbach A, Beier S, Twardziok SO, Wicker T, Radchuk V, Dockter C, Hedley PE, Russell J et al. (2017) A chromosome conformation capture ordered sequence of the barley genome. Nature 544(7651):427–433CAS 
    PubMed 

    Google Scholar 
    Mascher M (2019) Pseudomolecules and annotation of the second version of the reference genome sequence assembly of barley cv. morex [morex v2]. https://doi.ipk-gatersleben.de:443/DOI/83e8e186-dc4b-47f7-a820-28ad37cb176b/d1067eba-1d08-42e2-85ec-66bfd5112cd8/2McVean G (2009) A genealogical interpretation of principal components analysis. PLoS Genet 5(10):e1000686Mee JA, Yeaman S (2019) Unpacking conditional neutrality: genomic signatures of selection on conditionally beneficial and conditionally deleterious mutations. Am Nat 194(4):529–540PubMed 

    Google Scholar 
    Milner SG, Jost M, Taketa S, Mazón ER, Himmelbach A, Oppermann M, Weise S, Knüpffer H, Basterrechea M, König P et al. (2019) Genebank genomics highlights the diversity of a global barley collection. Nat Genet 51(2):319–326CAS 
    PubMed 

    Google Scholar 
    Morrell PL, Toleno DM, Lundy KE, Clegg MT (2005) Low levels of linkage disequilibrium in wild barley (Hordeum vulgare ssp. spontaneum) despite high rates of self-fertilization. Proc Natl Acad Sci USA 102(7):2442–2447CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Navarro JAR, Willcox M, Burgueño J, Romay C, Swarts K, Trachsel S, Preciado E, Terron A, Delgado HV, Vidal V et al. (2017) A study of allelic diversity underlying flowering-time adaptation in maize landraces. Nat Genet 49(3):476
    Google Scholar 
    Nevo E, Zohary D, Brown A, Haber M (1979) Genetic diversity and environmental associations of wild barley, Hordeum spontaneum, in Israel. Evolution 33(3):815–833CAS 
    PubMed 

    Google Scholar 
    Nevo E, Beharav A, Meyer RC, Hackett CA, Forster BP, Russell JR, Powell W (2005) Genomic microsatellite adaptive divergence of wild barley by microclimatic stress in ‘Evolution Canyon’, Israel. Biol J Linn Soc84(2):205–224
    Google Scholar 
    Oksanen J, Blanchet FG, Friendly M, Kindt R, Legendre P, McGlinn D, Minchin PR, O’Hara RB, Simpson GL, Solymos P, Stevens MHH, Szoecs E, Wagner H (2019) vegan: community ecology package. R package version 2.5-6. https://CRAN.R-project.org/package=veganPankin A, Altmüller J, Becker C, von Korff M (2018) Targeted resequencing reveals genomic signatures of barley domestication. New Phytol 218(3):1247–1259CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Pekel J-F, Cottam A, Gorelick N, Belward AS (2016) High-resolution mapping of global surface water and its long-term changes. Nature 540(7633):418–422CAS 

    Google Scholar 
    Pembleton L, Cogan N, Forster J (2013) StAMPP: an R package for calculation of genetic differentiation and structure of mixed-ploidy level populations Mol Ecol Res 13:946–952. https://doi.org/10.1111/1755-0998.12129CAS 
    Article 

    Google Scholar 
    Peterman WE (2018) Resistancega: an r package for the optimization of resistance surfaces using genetic algorithms. Methods Ecol Evol 9(6):1638–1647
    Google Scholar 
    Petkova D, Novembre J, Stephens M (2016) Visualizing spatial population structure with estimated effective migration surfaces. Nat Genet 48(1):94CAS 
    PubMed 

    Google Scholar 
    Pham A-T, Maurer A, Pillen K, Brien C, Dowling K, Berger B, Eglinton JK, March TJ (2019) Genome-wide association of barley plant growth under drought stress using a nested association mapping population. BMC Plant Biol 19(1):134PubMed 
    PubMed Central 

    Google Scholar 
    Poland JA, Brown PJ, Sorrells ME, Jannink J-L (2012) Development of high-density genetic maps for barley and wheat using a novel two-enzyme genotyping-by-sequencing approach. PLoS ONE 7(2):e32253CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Pyhäjärvi T, Hufford MB, Mezmouk S, Ross-Ibarra J (2013) Complex patterns of local adaptation in teosinte. Genome Biol Evol 5(9):1594–1609PubMed 
    PubMed Central 

    Google Scholar 
    Rellstab C, Gugerli F, Eckert AJ, Hancock AM, Holderegger R (2015) A practical guide to environmental association analysis in landscape genomics. Mol Ecol 24(17):4348–4370PubMed 

    Google Scholar 
    Renaut S, Grassa CJ, Yeaman S, Moyers BT, Lai Z, Kane NC, Bowers JE, Burke JM, Rieseberg LH (2013) Genomic islands of divergence are not affected by geography of speciation in sunflowers. Nat Commun 4(1):1–8
    Google Scholar 
    Russell J, Mascher M, Dawson IK, Kyriakidis S, Calixto C, Freund F, Bayer M, Milne I, Marshall-Griffiths T, Heinen S et al. (2016) Exome sequencing of geographically diverse barley landraces and wild relatives gives insights into environmental adaptation. Nat Genet 48(9):1024CAS 
    PubMed 

    Google Scholar 
    Samuk K, Samuk K, Owens GL, Delmore KE, Miller SE, Rennison DJ, Schluter D (2017) Gene flow and selection interact to promote adaptive divergence in regions of low recombination. Mol Ecol 26(17):4378–4390PubMed 

    Google Scholar 
    Sato K, Mascher M, Himmelbach A, Haberer G, Spannagl M, Stein N (2021) Chromosome-scale assembly of wild barley accession ‘OUH602’. G3 11(10):jkab244PubMed 
    PubMed Central 

    Google Scholar 
    Schmid K, Kilian B. Russell J (2018) Barley domestication, adaptation and population genomics. In: The Barley Genome, Springer International Publishing: Cham, pp 317–336Szkiba D, Kapun M, von Haeseler A, Gallach M (2014) SNP2GO: functional analysis of genome-wide association studies. Genetics 197(1):285–289PubMed 
    PubMed Central 

    Google Scholar 
    Terrazas RA, Balbirnie-Cumming K, Morris J, Hedley PE, Russell J, Paterson E, Baggs EM, Fridman E, Bulgarelli D (2020) A footprint of plant eco-geographic adaptation on the composition of the barley rhizosphere bacterial microbiota. Sci Rep 10(1):1–13
    Google Scholar 
    Tiffin P, Ross-Ibarra J (2014) Advances and limits of using population genetics to understand local adaptation. Trends Ecol Evol 29(12):673–680PubMed 

    Google Scholar 
    Tsuda Y, Chen J, Stocks M, Källman T, Sønstebø JH, Parducci L, Semerikov V, Sperisen C, Politov D, Ronkainen T et al. (2016) The extent and meaning of hybridization and introgression between siberian spruce (picea obovata) and norway spruce (picea abies): cryptic refugia as stepping stones to the west? Mol Ecol 25(12):2773–2789CAS 
    PubMed 

    Google Scholar 
    Turner-Hissong SD, Mabry ME, Beissinger TM, Ross-Ibarra J, Pires JC (2020) Evolutionary insights into plant breeding Curr Opin Plant Biol 54:93–100. https://doi.org/10.1016/j.pbi.2020.03.003CAS 
    Article 
    PubMed 

    Google Scholar 
    de Villemereuil P, Frichot É, Bazin É, François O, Gaggiotti OE (2014) Genome scan methods against more complex models: when and how much should we trust them? Mol Ecol 23(8):2006–2019PubMed 

    Google Scholar 
    Volis S (2011) Adaptive genetic differentiation in a predominantly self-pollinating species analyzed by transplanting into natural environment, crossbreeding and QST-FST test. New Phytol 192(1):237–248CAS 
    PubMed 

    Google Scholar 
    Volis S, Mendlinger S, Ward D (2002a) Differentiation in populations of Hordeum spontaneum along a gradient of environmental productivity and predictability: life history and local adaptation. Biol J Linn Soc 77(4):479–490
    Google Scholar 
    Volis S, Mendlinger S, Ward D (2002b) Adaptive traits of wild barley plants of Mediterranean and desert origin. Oecologia 133(2):131–138PubMed 

    Google Scholar 
    Volis S, Zaretsky M, Shulgina I (2010) Fine-scale spatial genetic structure in a predominantly selfing plant: role of seed and pollen dispersal. Heredity 105(4):384–393CAS 
    PubMed 

    Google Scholar 
    Volis S, Shulgina I, Ward D, Mendlinger S (2003) Regional subdivision in wild barley allozyme variation: adaptive or neutral? J Hered 94(4):341–351CAS 
    PubMed 

    Google Scholar 
    Volis S, Verhoeven K, Mendlinger S, Ward D (2004) Phenotypic selection and regulation of reproduction in different environments in wild barley. J Evol Biol 17(5):1121–1131CAS 
    PubMed 

    Google Scholar 
    Volis S, Yakubov B, Shulgina I, Ward D, Mendlinger S (2005) Distinguishing adaptive from nonadaptive genetic differentiation: comparison of q st and f st at two spatial scales. Heredity 95(6):466–475CAS 
    PubMed 

    Google Scholar 
    Volis S, Yakubov B, Shulgina I, Ward D, Zur V, Mendlinger S (2001) Tests for adaptive RAPD variation in population genetic structure of wild barley, Hordeum spontaneum Koch. Biol J Linn Soc 74(3):289–303
    Google Scholar 
    Wang X, Chen Z-H, Yang C, Zhang X, Jin G, Chen G, Wang Y, Holford P, Nevo E, Zhang G et al. (2018) Genomic adaptation to drought in wild barley is driven by edaphic natural selection at the Tabigha Evolution Slope. Proc Natl Acad Sci USA 115(20):5223–5228CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wiegmann M, Wiegmann M, Maurer A, Pham A, March TJ, Al-Abdallat A, Thomas WTB, Bull HJ, Shahid M, Eglinton J, Baum M, Flavell AJ, Tester M, Pillen K (2019) Barley yield formation under abiotic stress depends on the interplay between flowering time genes and environmental cues Sci Rep 9(1):6397. https://doi.org/10.1038/s41598-019-42673-1CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Yeaman S, Whitlock MC (2011) The genetic architecture of adaptation under migration–selection balance. Evolution 65(7):1897–1911PubMed 

    Google Scholar 
    Zheng X, Levine D, Shen J, Gogarten S, Laurie C, Weir B (2012) A high-performance computing toolset for relatedness and principal component analysis of snp data Bioinformatics 28(24):3326–3328. https://doi.org/10.1093/bioinformatics/bts606CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar  More