More stories

  • in

    A unified framework for herbivore-to-producer biomass ratio reveals the relative influence of four ecological factors

    A unified framework model of the H/P ratio
    Based on the Lotka–Volterra equations24,25, the biomass dynamics of primary producers (P) and herbivores (H) are described as follows:

    $$begin{array}{l}dP/dt = gleft( P right)P – xP – fleft( P right)PH,\ dH/dt = kfleft( P right)PH – mHend{array}$$
    (1)

    where g(P) is biomass-specific primary production (gC gC−1 d−1) and may be a function of P (gC m−2) owing to density-dependent growth, f(P) is per capita grazing rate of herbivores (m2 gC−1 d−1) and also may be function of P depending on the functional response, x is biomass-specific loss rate of primary producers other than due to grazing loss (gC gC−1 d−1), k is the conversion efficiency of herbivores as a fraction of ingested food converted into herbivore biomass (dimensionless: 0~1), and m is per capita mortality rate of herbivores owing to predation and other factors (gC gC−1 d−1). A list of model variables is listed in Supplementary Table 1. If we assume both g(P) and f(P) are constants, Eq. (1) is basically the Lotka–Volterra model, whereas it is an expansion of the Rosenzweig–MacArthur model if we assume logistic growth for g(P) and Michaelis–Menten (Holling type II) functional responses for f(P)24. At the equilibrium state, i.e., dP/dt = 0 and dH/dt = 0, the abundance of producers (P*) and consumers (H*) can be represented as:

    $$H ast /P ast = left{ {left[ {gleft( {P ast } right) – x} right]} right./left. {left. {fleft( {P ast } right)} right]} right}/left{ {{mathrm{m/}}left[ {kfleft( {P ast } right)} right]} right}$$
    (2)

    Thus, the relationship between H and P is not affected by the types of the functional response in herbivores (f(P)). If we set g(P) as the biomass-specific primary production rate at equilibrium, as in simple Lotka–Volterra equations (i.e., g(P*) = g), then the H/P ratio can be expressed with log transformation as:

    $${mathrm{log}}left( {H ast /P ast } right) = {mathrm{log}}left( k right) + {mathrm{log}}(g – x) – {mathrm{log}}left( m right)$$
    (3)

    At equilibrium, (g − x)P* is the amount of primary production that herbivores consume per unit of time (f(P*)P*H*). Thus, if this amount is divided by primary production per unit of time (P*g), it corresponds to the fraction of primary production that herbivores consume (0~1). We define it as β (= 1 − x/g). Large β values imply that producers are efficiently grazed at the equilibrium state. Thus, β is a gauge of inefficiency in the producers’ defensive traits. Using these parameters, the H*/P* ratio can be expressed as:

    $${mathrm{log}}left( {H ast /P ast } right) = {mathrm{log}}left( k right) + {mathrm{log}}left( beta right) + {mathrm{log}}left( g right) – {mathrm{log}}left( m right)$$
    (4)

    This equation implies that the H*/P* biomass ratio on a log scale is affected additively by the specific primary production rate (log(g)), the grazeable fraction of primary production (log(β)), the conversion efficiency (log(k)), and the mortality rate of herbivores (log(m)). According to this equation, communities with relatively low carnivore abundance would have a correspondingly low value of m and will exhibit high herbivore biomass relative to producer biomass (H*/P*), whereas those with low primary production (with low value of g) owing to, for example, low light supply will have a low H*/P* ratio. An increase in defended producers such as armored plants or a decrease in edible producers will decrease β by increasing the loss rate x owing to the cost of defense, and will result in a decreased H*/P* ratio. Finally, when the nutritional value of producers decreases, the conversion efficiency of herbivores (k) should be low, which in turn decreases the H*/P*biomass ratio.
    The model for a test with plankton communities
    To apply Eq. (4) to a natural community, some modifications are necessary. Here, we consider a plankton community composed of algae and zooplankton. A theory of ecological stoichiometry suggests that the carbon content of primary producers relative to their nutrient content such as nitrogen or phosphorus is an important property affecting growth efficiency in herbivores4. Supporting the theory, a number of studies have shown that growth rate in terms of carbon accumulation relative to ingestion rate strongly depends on the carbon contents of the food relative to nutrients26,27,28,29. Thus, k can be expressed as:

    $$k = q_1 times a_{{mathrm{nut}}}^{varepsilon 1}$$
    (5)

    where αnut is carbon content relative to nutrient content of primary producers and q1 is the conversion factor adjusting to biomass units. In this study, we applied a power function with coefficient of ε1 as a first order approximation because effects of this factor on the H*/P* biomass ratio may not be proportionally related to plant nutrient content. For example, if ε1 is much smaller than zero, it means that negative effects of the carbon to phosphorus ratio of algal food on an herbivore’s k are more substantial when the carbon to phosphorus ratio is high compared with the case when the carbon to phosphorus ratio is low. However, if this factor does not affect the H*/P* biomass ratio, ε1 = 0 and k is constant.
    As herbivorous zooplankton cannot efficiently graze on larger phytoplankton due to gape limitation30, the feeding efficiency of herbivores or the defense efficiency of the producers’ resistance traits, β, would be related to the fraction of edible algae in terms of size as follows:

    $$beta = q_2 times a_{{mathrm{edi}}}^{varepsilon 2}$$
    (6)

    where αedi is a trait determining producer edibility, q2 is a factor for converting the traits to edible efficiency, and ε2 is how effective the trait is in defending against grazing. We expect ε2 = 0 if this factor does not matter in regulating the H*/P* biomass ratio but ε2  > 0 if it has a role. Similarly, g can be described as

    $$g = q_3 times mu ^{varepsilon 3}$$
    (7)

    where μ is the specific growth rate of producers, q3 is a conversion factor, and ε3 is the effect of µ on growth rate. Again, we expect that ε3 ≠ 0 if g has a role in determining the H*/P* ratio. Finally, assuming a Holling type I functional response of carnivores, the mortality rate of herbivores, m, is expressed as:

    $$m = q_4 times theta ^{varepsilon 4}$$
    (8)

    where θ is abundance of carnivores, q4 is specific predation rate, and ε4 is the effect size of carnivore abundance on m.
    By inserting Eqs. (5–8) to Eq. (4), effects of factors on the H*/P* biomass ratio is formulated as:

    $${!}{mathrm{log}}left( {H ast /P ast } right) {!}= varepsilon _1{mathrm{log}}(a_{{mathrm{nut}}}) + varepsilon _2{mathrm{log}}(a_{{mathrm{edi}}}) + varepsilon _3{mathrm{log}}(mu ) – varepsilon _4{mathrm{log}}(theta ) + gamma$$
    (9)

    where γ is log(q1) + log(q2) + log(q3) − log(q4). If differences in the H*/P* ratio among communities are regulated by growth rate (μ), edibility (αedi), and nutrient contents (αnut) of producers as well as by predation by carnivores (θ), we expect non-zero values for ε1–ε4. Thus, Eq. (9) can be used to evaluate the relative importance of the four hypothesized agents if all of them simultaneously affect the H*/P* ratio. Here we undertake this analysis using data for natural plankton communities in experimental ponds where primary production rate was manipulated with different abundance of carnivore fish.
    Experimental test by plankton communities
    The experiment was carried out at two ponds (pond ID 217 and 218) located at the Cornell University Experimental Ponds Facility in Ithaca, NY, USA during 4 June to 28 August 2016 (Fig. 1). Each pond has a 0.09 ha surface area (30 × 30 m) and is 1.5 m deep. To initiate the experiment, we equally divided each of the two ponds into four sections using vinyl-coated canvas curtains, and randomly assigned the four sections to either high-shade (64% shading), mid-shade (47% shading), low-shade (33% shading), or no-shade treatments (no shading). Shading in each treatment was made using opaque floating mats (6 m diameter; Solar-cell SunBlanket, Century Products, Inc., Georgia, USA)31. The floating mats were deployed silvered side up to reflect sunlight and blue side down to avoid pond heating. Sampling was performed biweekly for water chemistry and abundance of phytoplankton and zooplankton with measurements of vertical profiles of water temperature, dissolved oxygen (DO) concentration and photosynthetic active radiation (PAR).
    Fig. 1: Ponds used in the experimental test.

    Pond 217 (a) and Pond 218 (b) in the Cornell University Experimental Ponds Facility divided into four sections by vinyl-canvas curtains and partially shaded by floating mats to regulate primary production rate. Floating docks were placed at the center of the ponds for sampling.

    Full size image

    PAR in the water column was lower in the sections with larger shaded areas throughout the experiment (Fig. 2b). Water temperature varied from 18 to 25 °C during the experiment but showed no notable differences in mean values of the water columns or the vertical profiles among the four treatments of the two ponds regardless of the shading treatment (Supplementary Fig. 1a, Supplementary Fig. 2). In all treatments, pH values gradually decreased towards the end of experiment and were higher in pond 218 (Supplementary Fig. 1b). DO concentration varied among the treatments and between the ponds, but were within the range of 5–12 mg L−1 (Supplementary Fig. 1c).
    Fig. 2: Relationships between zooplankton biomass and phytoplankton biomass, and between H/P mass ratio and photosynthetic active radiation.

    Biplots of zooplankton biomass (μg C L−1) and phytoplankton biomass (mg C L−1) (a) and H/P mass ratio and photosynthetic active radiation (PAR, mol photon m−2 d−1) (b) in the water column during the experiment in no-shade (blue), low-shade (orange), mid-shade (red), and high-shade (gray) treatments in pond 217 (circles) and 218 (squares). In each panel, small symbols denote values at each sampling date, and large symbols denote the mean values among the sampling dates. Bars denote standard errors on the means (n = 7 sampling date in each section). Correlation coefficients (r) with p values between the mean values are inserted in each panel.

    Full size image

    Phytoplankton biomass (mg C L−1) correlated significantly with chlorophyll a (µg L−1) (r = 0.702, p  More

  • in

    Peptide signaling without feedback in signal production operates as a true quorum sensing communication system in Bacillus subtilis

    The concentration of signal molecule ComX increases by the square of bacterial density
    In order to determine the dynamics of signal molecule (ComX) production, we have used experimental and mathematical modeling approaches. We quantified the ComX concentration over time in the spent medium of PS-216 (ΔcomP) with the biosensor strain BD2876 (for strain-description see Supplementary Table 1), which produces β-galactosidase in response to the exogenous addition of ComX34. The assay included proper controls and calibrations to assure the biosensor-derived ComX concentrations are accurate (for details see Materials and methods). We found that the ComX concentration correlated positively with population density of PS-216 (ΔcomP) and remained constant at 10 nM after entering the stationary phase (Fig. 2a). Importantly, the representation of ComX concentration versus cell density (OD650) (Fig. 2b) showed a non-linear trend between the two parameters. The experimental data were fitted by an allometric function:

    $$SMleft( t right) = aNleft( t right)^b$$
    (1)

    Where SM(t) is a signal molecule (ComX) concentration in time, N(t) is bacterial cell density in time, expressed as optical density of the bacterial suspension (OD650). The fitting results for parameters a and b were (9.6 ± 0.6) nM a.u.−2.09 and 2.09 ± 0.10, respectively. The value of parameter a means that at OD650 = 1.0 a.u., which corresponds to the stationary growth phase in our experimental conditions and the bacterial density of 4 × 108 cells mL‒1, the ComX concentration is about 10 nM. In the early exponential growth phase the concentration was about 0.1 nM. Considering parameter b, the value obtained (2.09 ± 0.10) indicates that the ComX concentration increases by the square of bacterial density. This means that with increasing population density, the SM concentration (ComX) increases by the second power, while the amount of ComX per cell increases linearly. This relationship suggests that ComQXPA has an ultra-sensitive encoder module9, where signal molecule production is very sensitive to cell density. The same mathematical relationship can be obtained by assuming that SM production rate per cell corresponds to the product of a specific cell growth rate and a cell density (i.e., population growth rate, for details, see Supplementary Methods, Derivation of ComQXPA communication system model). The dependence of the SM production rate per cell on (a) cell density and (b) the specific cell growth rate can be seen as an alternative way to obtain the ultra-sensitivity of encoders, which is usually achieved by SM dependent positive feedback in many QS systems9. This makes ComX a true indicator of population density, which also encodes information about the cell growth rate.
    Fig. 2: The accumulation of signal molecule (SM) during the growth of B. subtilis and fitness cost of SM production.

    a The growth curve (OD650) of B. subtilis PS-216 ΔcomP (no signal receptor) producing SM (ComX) that is accumulating in the growth medium of fermenter working in the batch mode of n = 3 biologically independent replicates is presented; b The experimental data n = 3 biologically independent replicates where the data ≥ limit of detection of SM was fitted by Eq. (1); the error bars for SM concertation are standard errors calculated from 8 technical replicates for each biological replicate; c The comparison of growth curves of B. subtilis PS-216 with no signal molecule receptor (ΔcomP) and no signal molecule production and receptor (ΔcomQXP) of n = 3 biologically independent replicates; the OD650 at t = 0 h was corrected with respect to the measured CFU of the inoculum. The slopes of the fitted lines in c correspond to the growth rate divided by log 2; the exponential phase points in the most reliable OD650 region ( >0.1 a.u. and  1.75 h, P  1 values obtained for all fits indicate positive cooperativity (i.e., ultrasensitivity14) in the binding of the transcriptional factor ComA to the srfA promoter43,44. This agrees with the research showing that two molecules of the ComA homodimer cooperatively bind to the two promoter regions located upstream of the RNAP binding sites of srfA13,20,45. The inactivation of the second promoter region decreases the promoter activity of srfA by 100-fold (ref. 13), which underscores the importance of the second binding region, explains n ≥ 2 and the sharp increase in srfA promoter activity with ComX concentration. In addition, we show here that the critical concentration of ComX required to induce quantifiable response (designated here as lower limit response (LLR)) is 0.2–0.5 nM. These results, therefore, suggest that the response per cell depends cooperatively on the ComX concentration and that the cells respond to very low concentrations of ComX.
    Fig. 4: Influence of oxygen on the presence of SM (ComX).

    a Signal molecule deficient B. subtilis PS-216 (∆comQ, PsrfAA-yfp) was incubated in the presence of SM for 4 h and the maximum normalized response was determined from the activity of the srfA promoter. n = 4 biologically independent replicates were performed. Best, concatenated fit to the model in Eq. (2) is presented together with 95% confidence level. b the logistic fit to one of the three growth curves (n = 3, biologically independent replicates) measured as OD650 of the culture B. subtilis PS 216 (srfA-lacZ) producing signal molecule, SM that accumulated in the growth medium of batch system is shown. The response per cell data, obtained as the β-galactosidase activity of srfA promoter of B. subtilis PS-216 (srfA-lacZ) was fitted by Eq. (3), (R2  > 0.99). The time interval at which SM concentration is high enough to cause the measurable response, i.e., lower limit of response (LLR) as predicted from data in experiments in (a) is given as dashed window in (b). One of the five qualitatively and quantitatively similar experimental results is presented. Error bars represent SD of 8 technical replicates. For fits of additional replicates and data variability refer to Supplementary Tables 3 and 4.

    Full size image

    Fully functional ComQXPA communication system does not require a positive feedback loop–the validation of the ComQXPA communication system
    The response curve in Fig. 4a is a function of the SM concentration only. In the more natural setting (i.e., during growth) the cells encounter growth-dependent changes in SM concentrations as well as changes in bacterial density and growth rate over time. We have therefore asked whether the response curve based on the modeling and results presented in Figs. 2b and  4a could fit the response data in the SM producing and responding strain exposed to changes in these three parameters.
    We cultivated the SM producing and responding PS-216 strain carrying the response reporter (PsrfA-lacZ) in a large volume bioreactor system (Fig. 4b). This allowed sterile sampling of spent medium and cells (for response quantification) at several time points, without affecting growth conditions. Immediately after the inoculation of the fresh medium by overnight culture the β-galactosidase activity of PS-216 (PsrfA-lacZ) was high. We assumed that this was a consequence of the accumulation of the expressed PsrfA reporter (β-galactosidase, RM) during the overnight growth. As a consequence of the dilution of the intracellular β-galactosidase (RM) due to cell division, the activity of the β-galactosidase decreased sharply after 2 h incubation (Fig. 4b). Simultaneously, as predicted by (Eq. 1), the concentration of SM (ComX) in the medium was increased exponentially during growth, and soon reached a critical concentration to activate the srfA promoter. In particular, as elucidated by the fits of (Eq. 2) to the data in Fig. 4a, the lower limit of the response (LLR) is reached shortly before upturn of the cell response curve in Fig. 4b. At this point the culture is in exponential growth phase at the cell density of 3 to 8 × 107 cells mL−1. The steep slope of the response curve indicates that the rate by which the response molecule (RM) is synthesized now exceeds the dilution due to the cell division rate. From now on, the response per cell correlates approximately linearly with OD650, suggesting a strong coupling to cell growth. Taking these facts into account and considering that the response molecule (RM) concentration is sensitive to the concentration of the signal, SM, (Eq. 2) and that SM can be expressed in terms of cell density (Eq. 1), the concentration of a response molecule per cell, RM(t)/N(t), can be analytically described (see also Supplementary Methods, Derivation of ComQXPA communication system model) as:

    $$frac{{RM(t)}}{{N(t)}} = frac{{RM0}}{{N(t)}} + frac{{RM1(t)}}{{N(t)}}$$
    (3)

    where RM0/N(t) is the response per cell of overnight culture, i.e., the overnight accumulated β-galactosidase. The second term, RM1(t)/N(t) accounts for the synthesis of the β-galactosidase after inoculation of a fresh medium and comprises the parameters describing the sensitivity of the response to a signal molecule, Wmax, Km, n (Eq. 2), the signal production, a (Eq. 1), cell density, N(t) and proportionality constant, k that gives the magnitude of the response per cell when the potential to respond to the signal is maximally fulfilled (i.e., at Wmax) and the specific growth rate is 1 h−1. The definition of RM1(t) is given in Supplementary Methods (eq S11). Note that for the derivation of Eq. (3) we assumed no degradation of SM occurs, as our experiments suggest SM was stable under the conditions studied (Supplementary Figs. 6a and 7, see also Supplementary Methods, Derivation of ComQXPA communication). All the parameters in (Eq. 3), except k in RM1(t) and RM0, were taken as constants obtained in the independent experiments by fits of (Eq. 1) and (Eq. 2). With k and RM0, as the only fitting parameters, we applied the mathematical model in (Eq. 3) (for details of the model equation see Supplementary Methods, Derivation of ComQXPA communication system model) to fit the experimental cell response data (Fig. 4b). The successful fit (see Supplementary Table 4 for details) indicates that the relationship assumed among cell density, cell growth, signal concentration and response in (Eq. 3) is valid and yields (760 ± 120) M.U. for k and (5.5 ± 1.5) M.U. a.u. for RM0. Again, we did not need to incorporate the SM feedback loop into our model, which is consistent with published results suggesting that this communication system lacks a feedback loop10,11,12,13.
    The ComQXPA dependent signaling and response at the cellular level
    So far, we have focused on the population averages, which is a traditional approach in studies on microbial communication systems17,46. We here report results on communication dynamics of B. subtilis at the single cell level using fluorescence-based molecular tools. This approach provides the means to track temporal changes in expression of genes involved in signal synthesis (signaling) and in response and thus provides the insight into a phenotypic heterogeneity within the population.
    We used the double-labeled fluorescencent strain B. subtilis PS-216 (comQ-yfp, srfA-cfp), in which fluorescent reporters were fused to the comQ and srfA promoters. The two genes code for the ComX signal-processing protein and the communication-activated operon, respectively. Since comQ and comX share the same promoter and their genetic sequences often overlap15 expression level of comQ corresponds to the expression level of comX. The fluorescence of individual cells was observed under the microscope in different growth phases and quantitatively analyzed (Fig. 5).
    Fig. 5: Single cell quantitative fluorescence microscopy of B. subtilis PS-216 (comQ-yfp, srfA-cfp).

    The fluorescence microscopy images were taken periodically during incubation of B. subtilis PS-216 (comQ-yfp, srfA-cfp) in a batch fermenter by the YFP filter (a), CFP filter (b) or DIC (c). The example YFP and CFP images, taken after 3 h of incubation were pseudo-colored. The scale bar represents 10 µm. n = 3 biologically independent experiments were performed. d % of population of cells that are hyper-expressing comQ-yfp or srfA-cfp is depicted. Gene expression level determined by single cell fluorescence microscopy was measured as Na-fluoresceinate standard normalized mean fluorescence intensity per cells expressing comQ-yfp (e) and srfA-cfp (f); ON is the overnight culture. One of the three qualitatively similar cell distributions is shown in (g) and (h) for comQ-yfp and srfA-cfp, respectively; areas under the curves are the same in all time points. For additional replicate see Supplementary Fig. 8.

    Full size image

    The observed expression pattern for the signaling gene (comQ-yfp) follows lognormal distribution (Fig. 5g). A small number of outliers in the comQ expression (on average, 10x brighter than the majority) were easily detected in the qualitative image analysis (Fig. 5a). These hyperproducers were not present in the overnight culture and began to occur during exponential growth, after 1-hour incubation in fresh medium, (Fig. 5d). In general, hyperproducers accounted for about 0.1–1% of the population, and their frequency increased during the first 6 hours. These data suggest that bulk of the ComX is nevertheless produced by the majority of the population as expected for the true QS communication systems. The contribution of the srfA hyperproducers to the total surfactin production is even less pronounced since their occurrence did not exceed 0.1% of the population (Fig. 5b, d).
    The most heterogeneous expression of the communication signaling gene (comQ-yfp) was observed in overnight culture, immediately after its transfer to the fresh medium (Fig. 5g, Supplementary Fig. 8a), but hyperproducers where not detectable at this time (Fig. 5d). Once the cells begun to divide, the distribution shifted to lower fluorescence intensities with a simultaneous decrease in heterogeneity, but from 3 to 4 h onwards single cell fluorescence gradually increased, along with an increase in population heterogeneity (Fig. 5e, g, Supplementary Fig. 8a). This suggests that the expression rate is now higher than the division rate (i.e., the production overpowers the dilution due to cell division). A similar pattern was observed in the communication response (srfA-cfp) (Fig. 5f, h, Supplementary Fig. 8b) with two major differences. The minimum level in comQ-Yfp fluorescence was reached 1 h later than srfA-Cfp fluorescence, which suggests the expression of ComX is in first hours low compared to the srfA expression. Nevertheless, the entire cell population, with the exception of hyperproducers, which represented only a fraction of the comQ/srfA expressing cells, followed unimodal lognormal distribution expression pattern. This suggests that the ComQXPA communication phenomenon in B. subtilis, at least under the conditions in our experiment, is not restricted to individuals and can be studied at the population level, i.e., the averages represent well the population.
    The comQ-yfp and srfA-cfp expression co-localization analysis (Supplementary Table 5) reveals the correlation coefficient of about 0.5, which is significantly (P = 0.01) higher than in the overnight culture. The presence of the correlation suggests that on average, cells that produce the signal more intensively, also respond to signal more intensively, supporting the idea of self-sensing47. However, the correlation coefficient strength was only moderate, suggesting that sensing of the external signal (sensing-of others) still works as expected for a typical QS system.
    The induced response in ComQXPA communication system is graded and almost switch-like
    The perfect QS system does not produce a response until the threshold bacterial density is reached and then immediately switches to a full response. This minimum to maximum transition may be either a perfect switch or a graded induction. By combining the information from Figs. 2b and 4a in the form of eq S9 results in the normalized ComQXPA response curve as a function of bacterial density (Fig. 6a) that resembles a graded switch like induction (compare to Fig. 1a). The perfect switch like communication system is unrealistic, because it requires that all the cells are perfectly synchronized and immediately switch to maximal response, leaving no time for the adaptation to the signal stimuli. It is reasonable to expect that for the true quorum sensing system (QS) most of the response has to occur within the same generation of dividing cells (ngen  More

  • in

    Cross-scale interaction of host tree size and climatic water deficit governs bark beetle-induced tree mortality

    1.
    USDAFS. Press Release: Survey Finds 18 Million Trees Died in California in 2018. https://www.fs.usda.gov/Internet/FSE_DOCUMENTS/FSEPRD609321.pdf (USDAFS, 2019).
    2.
    Griffin, D. & Anchukaitis, K. J. How unusual is the 2012-2014 California drought? Geophys. Res. Lett. 41, 9017–9023 (2014).
    ADS  Article  Google Scholar 

    3.
    Robeson, S. M. Revisiting the recent California drought as an extreme value. Geophys. Res. Lett. 42, 6771–6779 (2015).
    ADS  Article  Google Scholar 

    4.
    Asner, G. P. et al. Progressive forest canopy water loss during the 2012-2015 California drought. Proc. Natl Acad. Sci. USA 113, E249–E255 (2016).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    5.
    Brodrick, P. G. & Asner, G. P. Remotely sensed predictors of conifer tree mortality during severe drought. Environ. Res. Lett. 12, 115013 (2017).
    ADS  Article  Google Scholar 

    6.
    Fettig, C. J. in Managing Sierra Nevada Forests. PSW-GTR-237 Ch. 2 (USDA Forest Service, 2012).

    7.
    Kolb, T. E. et al. Observed and anticipated impacts of drought on forest insects and diseases in the United States. For. Ecol. Manag. 380, 321–334 (2016).
    Article  Google Scholar 

    8.
    Waring, R. H. & Pitman, G. B. Modifying lodgepole pine stands to change susceptibility to mountain pine beetle attack. Ecology 66, 889–897 (1985).
    Article  Google Scholar 

    9.
    Restaino, C. et al. Forest structure and climate mediate drought-induced tree mortality in forests of the Sierra Nevada, USA. Ecol. Appl. 0, e01902 (2019).
    Article  Google Scholar 

    10.
    USDAFS. Press Release: Record 129 million dead trees in California. https://www.fs.usda.gov/Internet/FSE_DOCUMENTS/fseprd566303.pdf (USDAFS, 2017).

    11.
    Young, D. J. N. et al. Long-term climate and competition explain forest mortality patterns under extreme drought. Ecol. Lett. 20, 78–86 (2017).
    PubMed  Article  PubMed Central  Google Scholar 

    12.
    Raffa, K. F. et al. Cross-scale drivers of natural disturbances prone to anthropogenic amplification: The dynamics of bark beetle eruptions. BioScience 58, 501–517 (2008).
    Article  Google Scholar 

    13.
    Boone, C. K., Aukema, B. H., Bohlmann, J., Carroll, A. L. & Raffa, K. F. Efficacy of tree defense physiology varies with bark beetle population density: a basis for positive feedback in eruptive species. Can. J. Res. 41, 1174–1188 (2011).
    Article  Google Scholar 

    14.
    Fettig, C. J., Mortenson, L. A., Bulaon, B. M. & Foulk, P. B. Tree mortality following drought in the central and southern Sierra Nevada, California, U.S. For. Ecol. Manag. 432, 164–178 (2019).
    Article  Google Scholar 

    15.
    Stephenson, N. L., Das, A. J., Ampersee, N. J. & Bulaon, B. M. Which trees die during drought? The key role of insect host-tree selection. J. Ecol. 75, 2383–2401 (2019).
    Article  Google Scholar 

    16.
    Senf, C., Campbell, E. M., Pflugmacher, D., Wulder, M. A. & Hostert, P. A multi-scale analysis of western spruce budworm outbreak dynamics. Landsc. Ecol. 32, 501–514 (2017).
    Article  Google Scholar 

    17.
    Seidl, R. et al. Small beetle, large-scale drivers: how regional and landscape factors affect outbreaks of the European spruce bark beetle. J. Appl Ecol. 53, 530–540 (2016).
    Article  Google Scholar 

    18.
    Fettig, C. J. in Insects and Diseases of Mediterranean Forest Systems (eds Lieutier, F. & Paine, T. D.) 499–528 (Springer International Publishing, 2016).

    19.
    Raffa, K. F. & Berryman, A. A. The role of host plant resistance in the colonization behavior and ecology of bark beetles (Coleoptera: Scolytidae). Ecol. Monogr. 53, 27–49 (1983).
    Article  Google Scholar 

    20.
    Logan, J. A., White, P., Bentz, B. J. & Powell, J. A. Model analysis of spatial patterns in mountain pine beetle outbreaks. Theor. Popul. Biol. 53, 236–255 (1998).
    CAS  PubMed  MATH  Article  PubMed Central  Google Scholar 

    21.
    Wallin, K. F. & Raffa, K. F. Feedback between individual host selection behavior and population dynamics in an eruptive herbivore. Ecol. Monogr. 74, 101–116 (2004).
    Article  Google Scholar 

    22.
    Franceschi, V. R., Krokene, P., Christiansen, E. & Krekling, T. Anatomical and chemical defenses of conifer bark against bark beetles and other pests. N. Phytol. 167, 353–376 (2005).
    CAS  Article  Google Scholar 

    23.
    Raffa, K. F., Grégoire, J.-C. & Staffan Lindgren, B. Natural History and Ecology of Bark Beetles 1–40 (Elsevier, 2015).

    24.
    Bentz, B. J. et al. Climate change and bark beetles of the western United States and Canada: direct and indirect effects. BioScience 60, 602–613 (2010).
    Article  Google Scholar 

    25.
    DeRose, R. J. & Long, J. N. Drought-driven disturbance history characterizes a southern Rocky Mountain subalpine forest. Can. J. Res. 42, 1649–1660 (2012).
    Article  Google Scholar 

    26.
    Hart, S. J., Veblen, T. T., Schneider, D. & Molotch, N. P. Summer and winter drought drive the initiation and spread of spruce beetle outbreak. Ecology 98, 2698–2707 (2017).
    PubMed  Article  PubMed Central  Google Scholar 

    27.
    Netherer, S., Panassiti, B., Pennerstorfer, J. & Matthews, B. Acute drought Is an important driver of bark beetle infestation in Austrian Norway spruce stands. Front. For. Glob. Change 2, 39 (2019).

    28.
    Kaiser, K. E., McGlynn, B. L. & Emanuel, R. E. Ecohydrology of an outbreak: mountain pine beetle impacts trees in drier landscape positions first. Ecohydrology 6, 444–454 (2013).
    Article  Google Scholar 

    29.
    Marini, L. et al. Climate drivers of bark beetle outbreak dynamics in Norway spruce forests. Ecography 40, 1426–1435 (2017).
    Article  Google Scholar 

    30.
    Sambaraju, K. R., Carroll, A. L. & Aukema, B. H. Multiyear weather anomalies associated with range shifts by the mountain pine beetle preceding large epidemics. For. Ecol. Manag. 438, 86–95 (2019).
    Article  Google Scholar 

    31.
    Hayes, C. J., Fettig, C. J. & Merrill, L. D. Evaluation of multiple funnel traps and stand characteristics for estimating western pine beetle-caused tree mortality. J. Econ. Entomol. 102, 2170–2182 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    32.
    Thistle, H. W. et al. Surrogate pheromone plumes in three forest trunk spaces: composite statistics and case studies. For. Sci. 50, 610–625 (2004).

    33.
    Miller, J. M. & Keen, F. P. Biology and Control of the Western Pine Beetle: A Summary of The First Fifty Years of Research (US Department of Agriculture, 1960).

    34.
    Chubaty, A. M., Roitberg, B. D. & Li, C. A dynamic host selection model for mountain pine beetle, Dendroctonus ponderosae Hopkins. Ecol. Model. 220, 1241–1250 (2009).
    Article  Google Scholar 

    35.
    Graf, M., Reid, M. L., Aukema, B. H. & Lindgren, B. S. Association of tree diameter with body size and lipid content of mountain pine beetles. Can. Entomol. 144, 467–477 (2012).
    Article  Google Scholar 

    36.
    Geiszler, D. R. & Gara, R. I. in Theory and Practice of Mountain Pine Beetle Management in Lodgepole Pine Forests: Symposium Proceedings (eds Berryman, A. A., Amman, G. D. & Stark, R. W.) (1978).

    37.
    Klein, W. H., Parker, D. L. & Jensen, C. E. Attack, emergence, and stand depletion trends of the mountain pine beetle in a lodgepole pine stand during an outbreak. Environ. Entomol. 7, 732–737 (1978).
    Article  Google Scholar 

    38.
    Mitchell, R. G. & Preisler, H. K. Analysis of spatial patterns of lodgepole pine attacked by outbreak populations of the mountain pine beetle. For. Sci. 37, 1390–1408 (1991).
    Google Scholar 

    39.
    Preisler, H. K. Modelling spatial patterns of trees attacked by bark-beetles. Appl. Stat. 42, 501 (1993).
    MATH  Article  Google Scholar 

    40.
    Jactel, H. & Brockerhoff, E. G. Tree diversity reduces herbivory by forest insects. Ecol. Lett. 10, 835–848 (2007).
    PubMed  Article  Google Scholar 

    41.
    Faccoli, M. & Bernardinelli, I. Composition and elevation of spruce forests affect susceptibility to bark beetle attacks: implications for forest management. Forests 5, 88–102 (2014).
    Article  Google Scholar 

    42.
    Berryman, A. A. in Bark Beetles in North American Conifers: A System for the Study of Evolutionary Biology 264–314 (University of Texas Press, 1982).

    43.
    Fettig, C. J. et al. The effectiveness of vegetation management practices for prevention and control of bark beetle infestations in coniferous forests of the western and southern United States. For. Ecol. Manag. 238, 24–53 (2007).
    Article  Google Scholar 

    44.
    Moeck, H. A., Wood, D. L. & Lindahl, K. Q. Host selection behavior of bark beetles (Coleoptera: Scolytidae) attacking Pinus ponderosa, with special emphasis on the western pine beetle, Dendroctonus brevicomis. J. Chem. Ecol. 7, 49–83 (1981).
    CAS  PubMed  Article  Google Scholar 

    45.
    Evenden, M. L., Whitehouse, C. M. & Sykes, J. Factors influencing flight capacity of the mountain pine beetle (Coleoptera: Curculionidae: Scolytinae). Environ. Entomol. 43, 187–196 (2014).
    CAS  PubMed  Article  Google Scholar 

    46.
    Raffa, K. F. & Berryman, A. A. Accumulation of monoterpenes and associated volatiles following inoculation of grand fir with a fungus transmitted by the fir engraver, Scolytus ventralis (Coleoptera: Scolytidae). Can. Entomol. 114, 797–810 (1982).
    CAS  Article  Google Scholar 

    47.
    Anderegg, W. R. L. et al. Tree mortality from drought, insects, and their interactions in a changing climate. N. Phytol. 208, 674–683 (2015).
    Article  Google Scholar 

    48.
    Kane, V. R. et al. Assessing fire effects on forest spatial structure using a fusion of Landsat and airborne LiDAR data in Yosemite National Park. Remote Sens. Environ. 151, 89–101 (2014).
    ADS  Article  Google Scholar 

    49.
    Larson, A. J. & Churchill, D. Tree spatial patterns in fire-frequent forests of western North America, including mechanisms of pattern formation and implications for designing fuel reduction and restoration treatments. For. Ecol. Manag. 267, 74–92 (2012).
    Article  Google Scholar 

    50.
    Morris, J. L. et al. Managing bark beetle impacts on ecosystems and society: Priority questions to motivate future research. J. Appl. Ecol. 54, 750–760 (2017).
    Article  Google Scholar 

    51.
    Shiklomanov, A. N. et al. Enhancing global change experiments through integration of remote-sensing techniques. Front. Ecol. Environ. 17, 215–224 (2019).

    52.
    Jeronimo, S. M. A. et al. Forest structure and pattern vary by climate and landform across active-fire landscapes in the montane Sierra Nevada. For. Ecol. Manag. 437, 70–86 (2019).
    Article  Google Scholar 

    53.
    Roussel, J.-R., Auty, D., De Boissieu, F. & Meador, A. S. lidR: Airborne LiDAR Data Manipulation and Visualization for Forestry Applications (2019).

    54.
    McDowell, N. et al. Mechanisms of plant survival and mortality during drought: why do some plants survive while others succumb to drought? N. Phytol. 178, 719–739 (2008).
    Article  Google Scholar 

    55.
    Seybold, S. J. et al. Management of western North American bark beetles with semiochemicals. Annu. Rev. Entomol. 63, 407–432 (2018).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    56.
    Fettig, C. J., McKelvey, S. R. & Huber, D. P. W. Nonhost angiosperm volatiles and verbenone disrupt response of western pine beetle, Dendroctonus brevicomis (Coleoptera: Scolytidae), to attractant-baited traps. J. Econ. Entomol. 98, 2041–2048 (2005).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    57.
    Fettig, C. J., Dabney, C. P., McKelvey, S. R. & Huber, D. P. W. Nonhost angiosperm volatiles and verbenone protect individual ponderosa pines from attack by western pine beetle and red turpentine beetle (Coleoptera: Curculionidae, Scolytinae). West J. Appl. 23, 40–45 (2008).
    Google Scholar 

    58.
    Fettig, C. J. et al. Efficacy of ‘Verbenone Plus’ for protecting ponderosa pine trees and stands from Dendroctonus brevicomis (Coleoptera: Curculionidae) attack in British Columbia and California. J. Econ. Entomol. 105, 1668–1680 (2012).
    PubMed  Article  PubMed Central  Google Scholar 

    59.
    Oliver, W. W. Is self-thinning in ponderosa pine ruled by Dendroctonus bark beetles? In Forest Health Through Silviculture: Proceedings of the 1995 National Silviculture Workshop 6 (1995).

    60.
    Fettig, C. & McKelvey, S. Resiliency of an interior ponderosa pine forest to bark beetle infestations following fuel-reduction and forest-restoration treatments. Forests 5, 153–176 (2014).
    Article  Google Scholar 

    61.
    Fettig, C. J. & Hilszczański, J. Bark Beetles 555–584. https://doi.org/10.1016/B978-0-12-417156-5.00014-9 (Elsevier, 2015).

    62.
    Chesson, P. Mechanisms of maintenance of species diversity. Annu. Rev. Ecol. Syst. 31, 343–366 (2000).
    Article  Google Scholar 

    63.
    Fricker, G. A. et al. More than climate? Predictors of tree canopy height vary with scale in complex terrain, Sierra Nevada, CA (USA). For. Ecol. Manag. 434, 142–153 (2019).
    Article  Google Scholar 

    64.
    Ma, S., Concilio, A., Oakley, B., North, M. & Chen, J. Spatial variability in microclimate in a mixed-conifer forest before and after thinning and burning treatments. For. Ecol. Manag. 259, 904–915 (2010).
    Article  Google Scholar 

    65.
    Stovall, A. E. L., Shugart, H. & Yang, X. Tree height explains mortality risk during an intense drought. Nat. Commun. 10, 1–6 (2019).
    Article  CAS  Google Scholar 

    66.
    Stephenson, N. L. & Das, A. J. Height-related changes in forest composition explain increasing tree mortality with height during an extreme drought. Nat. Commun. 11, 3402 (2020).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    67.
    Stovall, A. E. L., Shugart, H. H. & Yang, X. Reply to ‘Height-related changes in forest composition explain increasing tree mortality with height during an extreme drought’. Nat. Commun. 11, 3401 (2020).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    68.
    Person, H. L. Tree selection by the western pine beetle. J. For. 26, 564–578 (1928).
    Google Scholar 

    69.
    Person, H. L. Theory in explanation of the selection of certain trees by the western pine beetle. J. For. 29, 696–699 (1931).
    CAS  Google Scholar 

    70.
    Pile, L. S., Meyer, M. D., Rojas, R., Roe, O. & Smith, M. T. Drought impacts and compounding mortality on forest trees in the southern Sierra Nevada. Forests 10, 237 (2019).
    Article  Google Scholar 

    71.
    Frey, J., Kovach, K., Stemmler, S. & Koch, B. UAV photogrammetry of forests as a vulnerable process. A sensitivity analysis for a structure from motion RGB-image pipeline. Remote Sens. 10, 912 (2018).
    ADS  Article  Google Scholar 

    72.
    James, M. R. & Robson, S. Mitigating systematic error in topographic models derived from UAV and ground-based image networks. Earth Surf. Process. Landf. 39, 1413–1420 (2014).
    ADS  Article  Google Scholar 

    73.
    Gray, P. C. et al. A convolutional neural network for detecting sea turtles in drone imagery. Methods Ecol. Evol. 10, 345–355 (2019).
    Article  Google Scholar 

    74.
    Millar, C. I., Stephenson, N. L. & Stephens, S. L. Climate change and forests of the future: Managing in the face of uncertainty. Ecol. Appl. 17, 2145–2151 (2007).
    PubMed  Article  PubMed Central  Google Scholar 

    75.
    Vose, J. M. et al. in Impacts, Risks, and Adaptation in the United States: The Fourth National Climate Assessment Vol II (eds Reidmiller, D. R., et al.) 232–267. https://nca2018.globalchange.gov/chapter/6/https://doi.org/10.7930/NCA4.2018.CH6 (2018).

    76.
    Bedard, W. D. et al. Western pine beetle: field response to its sex pheromone and a synergistic host terpene, myrcene. Science 164, 1284–1285 (1969).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    77.
    Byers, J. A. & Wood, D. L. Interspecific inhibition of the response of the bark beetles, Dendroctonus brevicomis and Ips paraconfusus, to their pheromones in the field. J. Chem. Ecol. 6, 149–164 (1980).
    CAS  Article  Google Scholar 

    78.
    Shepherd, W. P., Huber, D. P. W., Seybold, S. J. & Fettig, C. J. Antennal responses of the western pine beetle, Dendroctonus brevicomis (Coleoptera: Curculionidae), to stem volatiles of its primary host, Pinus ponderosa, and nine sympatric nonhost angiosperms and conifers. Chemoecology 17, 209–221 (2007).
    CAS  Article  Google Scholar 

    79.
    DJI. Zenmuse X3 – Creativity Unleashed. DJI Official https://www.dji.com/zenmuse-x3/info (2015).

    80.
    Micasense. MicaSense. https://support.micasense.com/hc/en-us/articles/215261448-RedEdge-User-Manual-PDF-Download- (2015).

    81.
    DJI. DJI – The World Leader in Camera Drones/Quadcopters for Aerial Photography. DJI Official https://www.dji.com/matrice100/info (2015).

    82.
    Wyngaard, J. et al. Emergent challenges for science sUAS data management: fairness through community engagement and best practices development. Remote Sens. 11, 1797 (2019).
    ADS  Article  Google Scholar 

    83.
    Rouse, W., Haas, R. H., Deering, W. & Schell, J. A. Monitoring the Vernal Advancement and Retrogradation (Green Wave Effect) of Natural Vegetation (Remote Sensing Center, Texas A&M Univ., 1973).

    84.
    DronesMadeEasy. Map Pilot for DJI on iOS. App Store https://itunes.apple.com/us/app/map-pilot-for-dji/id1014765000?mt=8 (2018).

    85.
    Farr, T. G. et al. The shuttle radar topography mission. Rev. Geophys. 45, RG2004 (2007).

    86.
    Zhang, W. et al. An easy-to-use airborne LiDAR data filtering method based on cloth simulation. Remote Sens. 8, 501 (2016).
    ADS  Article  Google Scholar 

    87.
    Hijmans, R. J. et al. Raster: Geographic Data Analysis and Modeling (2019).

    88.
    Gitelson, A. & Merzlyak, M. N. Spectral reflectance changes associated with autumn senescence of Aesculus hippocastanum L. and Acer platanoides L. leaves. Spectral features and relation to chlorophyll estimation. J. Plant Physiol. 143, 286–292 (1994).
    CAS  Article  Google Scholar 

    89.
    Coops, N. C., Johnson, M., Wulder, M. A. & White, J. C. Assessment of QuickBird high spatial resolution imagery to detect red attack damage due to mountain pine beetle infestation. Remote Sens. Environ. 103, 67–80 (2006).
    ADS  Article  Google Scholar 

    90.
    Clevers, J. G. P. W. & Gitelson, A. A. Remote estimation of crop and grass chlorophyll and nitrogen content using red-edge bands on Sentinel-2 and -3. Int. J. Appl. Earth Observ. Geoinf. 23, 344–351 (2013).
    ADS  Article  Google Scholar 

    91.
    Li, W., Guo, Q., Jakubowski, M. K. & Kelly, M. A new method for segmenting individual trees from the LiDAR point cloud. Photogramm. Eng. Remote Sens. 78, 75–84 (2012).
    Article  Google Scholar 

    92.
    Jakubowski, M. K., Li, W., Guo, Q. & Kelly, M. Delineating individual trees from LiDAR data: a comparison of vector- and raster-based segmentation approaches. Remote Sens. 5, 4163–4186 (2013).
    ADS  Article  Google Scholar 

    93.
    Shin, P., Sankey, T., Moore, M. & Thode, A. Evaluating unmanned aerial vehicle images for estimating forest canopy fuels in a ponderosa pine stand. Remote Sens. 10, 1266 (2018).
    ADS  Article  Google Scholar 

    94.
    Roussel, J.-R. lidRplugins: Extra Functions and Algorithms for lidR Package (2019).

    95.
    Eysn, L. et al. A benchmark of LiDAR-based single tree detection methods using heterogeneous forest data from the alpine space. Forests 6, 1721–1747 (2015).
    Article  Google Scholar 

    96.
    Vega, C. et al. PTrees: a point-based approach to forest tree extraction from LiDAR data. Int. J. Appl. Earth Observ. Geoinf. 33, 98–108 (2014).
    ADS  Article  Google Scholar 

    97.
    Plowright, A. ForestTools: Analyzing Remotely Sensed Forest Data (2018).

    98.
    Pau, G., Fuchs, F., Sklyar, O., Boutros, M. & Huber, W. EBImage: An R package for image processing with applications to cellular phenotypes. Bioinformatics 26, 979–981 (2010).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    99.
    Meyer, F. & Beucher, S. Morphological segmentation. J. Vis. Commun. Image Represent. 1, 21–46 (1990).
    Article  Google Scholar 

    100.
    Hunziker, P. Velox: Fast Raster Manipulation and Extraction (2017).

    101.
    Kuhn, M. Building predictive models in R using the caret package. J. Stat. Softw. 28, 1–26 (2008).
    Article  Google Scholar 

    102.
    Wang, Y. et al. Is field-measured tree height as reliable as believed – a comparison study of tree height estimates from field measurement, airborne laser scanning and terrestrial laser scanning in a boreal forest. ISPRS J. Photogramm. Remote Sens. 147, 132–145 (2019).
    ADS  Article  Google Scholar 

    103.
    Weinstein, B. G., Marconi, S., Bohlman, S., Zare, A. & White, E. Individual tree-crown detection in RGB imagery using semi-supervised deep learning neural networks. Remote Sens. 11, 1309 (2019).
    ADS  Article  Google Scholar 

    104.
    dos Santos, A. A. et al. Assessment of CNN-based methods for individual tree detection on images captured by RGB cameras attached to UAVs. Sensors (Basel) 19, 3595 (2019).

    105.
    Stephenson, N. Actual evapotranspiration and deficit: biologically meaningful correlates of vegetation distribution across spatial scales. J. Biogeogr. 25, 855–870 (1998).
    Article  Google Scholar 

    106.
    Flint, L. E., Flint, A. L., Thorne, J. H. & Boynton, R. Fine-scale hydrologic modeling for regional landscape applications: The California Basin Characterization Model development and performance. Ecol. Process. 2, 25 (2013).
    Article  Google Scholar 

    107.
    Millar, C. I. et al. Forest mortality in high-elevation whitebark pine (Pinus albicaulis) forests of eastern California, USA: influence of environmental context, bark beetles, climatic water deficit, and warming. Can. J. For. Res. 42, 749–765 (2012).
    Article  Google Scholar 

    108.
    Baldwin, B. G. et al. Species richness and endemism in the native flora of California. Am. J. Bot. 104, 487–501 (2017).
    PubMed  Article  PubMed Central  Google Scholar 

    109.
    Bürkner, P.-C. brms: an R package for bayesian multilevel models using Stan. J. Stat. Softw. 80, 1–28 (2017).
    Article  Google Scholar 

    110.
    Hoffman, M. D. & Gelman, A. The no-U-turn sampler: adaptively setting path lengths in Hamiltonian Monte Carlo. J. Mach. Learn. Res. 15, 31 (2014).
    MathSciNet  MATH  Google Scholar 

    111.
    Carpenter, B. et al. Stan: a probabilistic programming language. J. Stat. Softw. 76, 1–32 (2017).
    Article  Google Scholar 

    112.
    Brooks, S. P. & Gelman, A. General methods for monitoring convergence of iterative simulations. J. Comput. Graph. Stat. 7, 434 (1998).
    MathSciNet  Google Scholar 

    113.
    Gabry, J., Simpson, D., Vehtari, A., Betancourt, M. & Gelman, A. Visualization in Bayesian workflow. J. R. Stat. Soc. Ser. A 182, 389–402 (2019).
    MathSciNet  Article  Google Scholar 

    114.
    Koontz, M. J., Latimer, A. M., Mortenson, L. A., Fettig, C. J. & North, M. P. Drone-derived data supporting “Cross-scale interaction of host tree size and climatic water deficit governs bark beetle-induced tree mortality”. https://doi.org/10.17605/OSF.IO/3CWF9 (2020).

    115.
    Baldwin, B. G. et al. Master spatial file for native California vascular plants used by Baldwin et al. (2017 Amer. J. Bot.), Dryad, Dataset, 2017. https://doi.org/10.6078/D16K5W.

    116.
    R Core Team. R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, 2018).

    117.
    Koontz, M. J., Latimer, A. M., Mortenson, L. A., Fettig, C. J. & North, M. P. Local-structure-wpb-severity. https://doi.org/10.17605/OSF.IO/WPK5Z (2019). More

  • in

    Continued preference for suboptimal habitat reduces bat survival with white-nose syndrome

    Study design
    This study began in 2013, before P. destructans invaded Michigan and Wisconsin, and continued through 2020, after all Michigan and Wisconsin hibernacula were invaded by the pathogen. Throughout this period, we visited 22 hibernacula twice per winter during bat hibernation to quantify bat colony sizes, individual bat roosting temperatures, fungal loads (to which we added a constant 0.0001 before transforming to the log10 scale), and recapture probabilities (Fig. 2). We sampled bats during early hibernation in November, when more than 95% of swarm activity was expected to be finished51, and again during late hibernation in March, when less than 1% of spring emergence activity was expected to have begun51,52. For each sampling event, we counted all bats of all species within the site. We focused our analyses on the little brown bat (Myotis lucifugus). This species suffered severe declines due to WNS24,53, and was the only species abundant enough to provide sufficient sample sizes for our analyses. We divided bat counts by sections within the hibernaculum (i.e., “rooms”) that potentially varied in microclimate, and we used HOBO Pro v2 data loggers to continuously record temperatures every 3 h in these sections (Supplementary Figs. 2 and 7).
    After counting all bats, we haphazardly sampled 20–25 individual little brown bats stratified across sections and roughly in proportion to the number of bats in each section (Fig. 2). For each sampled bat, we used a Fluke 62 MAX IR laser thermometer to quantify the temperature of the substrate directly adjacent to the bat (2 years is defined here as the post-invasion period) and that we had sampled during both the pre-invasion period (up to and including year 0, when WNS was first detected) and invasion period (years 1–2 of invasion). From pre-invasion to post-invasion, bat temperature distributions might have changed across these 12 hibernacula via two mechanisms: (1) bat preferences could have remained the same while populations declined in hibernacula and/or microsites with unfavorable temperatures, causing distribution shifts via selection; and/or (2) bat preferences could have changed towards avoidance of hibernacula and/or microsites with high temperature-mediated disease mortality, causing distribution shifts via selection or learning. In both cases, we would expect the average regional bat roosting temperatures to change over time. In contrast, we would only expect the average bat roosting temperature in any given site (or section) to change over time if the site provided a wide range of temperatures to select from, which was not always the case in the hibernacula we studied. Therefore, we quantified how the regional mean early hibernation roosting temperatures for all counted bats changed across all 12 of these hibernacula from pre-invasion to post-invasion (2013–2020, up to four years after the fungus was first detected) using a Gaussian regression, without including a random site effect. We visualized this change using density plots with 1 °C smoothing bin widths.
    Though we stratified bat roosting temperature sampling by section, roughly in proportion to the total number of bats present in a section, we only sampled up to 25 little brown bats per site, and we could not sample exactly in proportion to the total bats present. Therefore, calculating the mean temperatures based on only sampled bats might be misleading. Instead, we used the observed temperatures from our sampled bats to estimate the roosting temperatures for bats that were counted but not sampled. In particular, we randomly generated an additional set of temperatures for all unmeasured, counted bats based on the mean and standard deviation of the roosting temperatures of sampled bats in the same section during the same survey. We then used both known (sampled bats) and estimated (counted bats) temperatures in our Gaussian regression to determine how average temperatures changed from pre-invasion to post-invasion. We also confirmed that qualitative results were the same if we only used the temperatures from sampled bats, without estimation. By considering how bats’ early hibernation temperature distributions changed throughout invasion, regardless of infection status, we were quantifying bat roosting preferences within and between hibernacula, rather than behavioral responses to infection within a season (i.e., “behavioral fever”).
    Finally, we used the number of bats counted per hibernaculum during early hibernation (November) and late hibernation (March) in all hibernacula that still had bats in post-invasion years to calculate two demographic parameters: over summer population growth rates (reflecting immigration and recruitment to sites between March and November) and over winter mortality rates (population growth rates from November to March). We used a Gaussian linear model to determine whether the log10 population growth rates were correlated with average early hibernation roosting temperatures for sampled bats in each hibernaculum, the demographic season (over summer versus over winter), and the interaction between temperature and demographic season. A significant interaction term would indicate that temperature-correlated bat immigration and recruitment rates were decoupled from temperature-mediated population declines.
    All analyses described above were performed in R version 3.5.161 with packages ‘ggplot2’62, ‘lme4’63, ‘loo’64, ‘pROC’65, and ‘R2jags56. Visual inspection of residuals and predictions plots confirmed acceptable model fits for all regression models.
    Reporting summary
    Further information on research design is available in the Nature Research Reporting Summary linked to this article. More

  • in

    Respiration characteristics and its responses to hydrothermal seasonal changes in reconstructed soils

    1.
    Huang, S. D. et al. Autotrophic and heterotrophic soil respiration responds asymmetrically to drought in a subtropical forest in the Southeast China. Soil Biol. Biochem. 123, 242–249 (2018).
    CAS  Article  Google Scholar 
    2.
    Peichl, M. et al. Management and climate effects on carbon dioxide and energy exchanges in a maritime grassland. Agric. Ecosyst. Environ. 158, 132–146 (2012).
    Article  Google Scholar 

    3.
    Zhang, X. B. et al. How do environmental factors and different fertilizer strategies affect soil CO2 emission and carbon sequestration in the upland soils of southern China?. Appl. Soil Ecol. 72, 109–118 (2013).
    Article  Google Scholar 

    4.
    Curiel, Y. J. et al. Microbial soil respiration and its dependency on carbon inputs, soil temperature and moisture. Glob. Change Biol. 13, 2018–2035 (2010).
    ADS  Article  Google Scholar 

    5.
    Peri, P. L., Bahamonde, H. & Christiansen, R. Soil respiration in Patagonian semiarid grasslands under contrasting environmental and use conditions. J. Arid Environ. 119, 1–8 (2015).
    ADS  Article  Google Scholar 

    6.
    Davidson, E. A. et al. Effects of soil water content on soil respiration in forests and cattle pastures of eastern Amazonia. Biogeochemistry 48, 53–69 (2000).
    CAS  Article  Google Scholar 

    7.
    Buchmann, N. Biotic and abiotic factors controlling soil respiration rates in Piceaavies stands. Soil Biol Biochem. 32, 1625–1635 (2000).
    CAS  Article  Google Scholar 

    8.
    Li, X. et al. Contribution of root respiration to total soil respiration in a semi-arid grassland on the Loess Plateau, China. Sci. Total Environ. 62, 1209–1217 (2018).
    ADS  Google Scholar 

    9.
    Jia, B. & Zhou, G. Integrated diurnal soil respiration model during growing season of a typical temperate steppe: Effects of temperature, soil water content and biomass production. Soil Biol. Biochem. 41, 681–686 (2009).
    CAS  Article  Google Scholar 

    10.
    Jin, Z., Qi, Y. C., Dong, Y. S. & Domroes, M. Seasonal patterns of soil respiration in three types of communities along grass-desert shrub transition in Inner Mongolia, China. Adv. Atmos. Sci. 26, 503–512 (2009).
    CAS  Article  Google Scholar 

    11.
    Li, Y. C., Hou, C. C., Song, C. C. & Guo, Y. D. Seasonal changes in the contribution of root respiration to total soil respiration in a freshwater marsh in Sanjiang Plain, Northeast China. Environ. Earth Sci. 75, 1–10 (2016).
    ADS  Article  Google Scholar 

    12.
    Zhao, C. Y., Zhao, Z. M., Yilihamu, Hong, Z. & Jun, L. Contribution of root and rhizosphere respiration of Haloxylonammodendron to seasonal variation of soil respiration in the central Asian desert. Quatern. Int. 244, 304–309 (2011).
    Article  Google Scholar 

    13.
    Wang, X. et al. Soil respiration under climate warming: Differential response of heterotrophic and autotrophic respiration. Glob. Change Biol. 20, 3229–3237 (2014).
    ADS  Article  Google Scholar 

    14.
    Sun, Z., Han, J. C. & Wang, H. Y. Soft rock for improving crop yield in sandy soil of Mu Us sandy land, China. Arid Land Res. Manag. 33, 136–154 (2019).
    CAS  Article  Google Scholar 

    15.
    Dong, O. G. et al. water infiltration of covering soils with Different textures and bulk densities in gravelmulched areas. Appl. Soil Ecol. 17, 14039–14052 (2019).
    Google Scholar 

    16.
    Lei, N., Han, J. C., Mu, X. M., Sun, Z. H. & Wang, H. Y. Effects of improved materials on reclamation of soil properties and crop yield in hollow villages in China. J. Soil Sediment 19, 2374–2380 (2019).
    CAS  Article  Google Scholar 

    17.
    Fu, G. et al. Response of ecosystem respiration to experimental warming and clipping at daily time scale in an alpine meadow of Tibet. J. Mt. Sci.-Engl. 10, 455–463 (2013).
    Article  Google Scholar 

    18.
    Wu, J. P. et al. Response of soil respiration and ecosystem carbon budget to vegetation removal in Eucalyptus plantations with contrasting ages. Sci. Rep.-UK 4, 6262 (2014).
    Article  Google Scholar 

    19.
    Han, T. F., Huang, W. J., Liu, J. X., Zhou, G. Y. & Xiao, Y. Different soil respiration responses to litter manipulation in three subtropical successional forests. Sci. Rep.-UK 5, 18166 (2015).
    ADS  CAS  Article  Google Scholar 

    20.
    Lei, N. & Han, J. C. Effect of precipitation on respiration of different reconstructed soils. Sci. Rep.-UK 10, 7328 (2020).
    ADS  CAS  Article  Google Scholar 

    21.
    Kosugi, Y. et al. Spatial and temporal variation in soil respiration in a Southeast Asian tropical rainforest. Agric. Forest Meteorol. 147, 35–47 (2007).
    ADS  Article  Google Scholar 

    22.
    Hu, Z. D., Liu, S. R., Shi, Z. M., Liu, X. L. & He, F. Diel variations and seasonal dynamics of soil respirations in subalpine meadow in western Sichuan Province, China. Acta Ecol. Sin. 32, 6376–6386 (2012).
    Article  Google Scholar 

    23.
    Shi, Z. et al. Seasonal variation and temperature sensitivity of soil V respiration under different plant communities along an elevation gradient in Wuyi Mountains of China. J. Appl. Ecol. 19, 2357–2363 (2008).
    Google Scholar 

    24.
    Wu, C. S., Sha, L. Q., Gao, J. M. & Dong, L. Y. Analysis of soil respiration in a montane evergreen broad-leaved forest and an artificial tea garden in Ailao Mountains, Yunnan province. J. Nanjing For. Univ. (Nat. Sci.) 36, 64–68 (2012) (in Chinese).
    Google Scholar 

    25.
    Zhao, J. X. et al. Soil respiration and its affecting factors of Pinusyunnanensis in the Middle Regions of Yunnan. J. Northwest For. Univ. 30, 8–13 (2015) (in Chinese).
    Google Scholar 

    26.
    Rayment, M. B. & Jarvis, P. G. Temporal and spatial variation of soil CO2 efflux in a Canadian boreal forest. Soil Biol. Biochem. 32, 35–45 (2000).
    CAS  Article  Google Scholar 

    27.
    Qi, Y. & Xu, M. Separating the effects of moisture and temperature on soil CO2 efflux in a coniferous forest in the Sierra Nevada Mountains. Plant Soil 237, 15–23 (2001).
    CAS  Article  Google Scholar 

    28.
    Dilustro, J. J., Collins, B., Duncan, L. & Crawford, C. Moisture and soil texture effects on soil CO2 efflux components in southeastern mixed pine forests. Forest Ecol. Manag. 204, 87–97 (2005).
    Article  Google Scholar 

    29.
    Gaumont-guay, D. et al. Influence of temperature and drought on seasonal and interannual variations of soil, bole and ecosystem respiration in a boreal aspen stand. Agric Forest Meteorol. 140, 203–219 (2006).
    ADS  Article  Google Scholar 

    30.
    Yan, J. X., Hao, Z., Jing, X. K. & Li, H. J. Variations of soil respiration and its temperature sensitivity in an orchard in Jinci Region of Taiyuan City. J. Environ. Sci.-China 37, 3625–3633 (2016).
    Google Scholar 

    31.
    Han, G. X., Zhou, G. S. & Xu, Z. Z. Research and prospects for soil respiration of farmland ecosystems in China. J. Plant Ecol.-UK 32, 719–733 (2008).
    CAS  Google Scholar 

    32.
    Yuste, J. C., Janssens, I. A., Carrara, A., Meiresonne, L. & Ceulemans, R. Interactive effects of temperature and precipitation on soil respiration in a temperate maritime pine forest. Tree Physiol. 23, 1263–1270 (2003).
    Article  Google Scholar 

    33.
    Conant, R. T., Dalla-bett, A. P., Klopatek, C. C. & Klopatek, J. Controls on soil respiration in semiarid soils. Soil Biol. Biochem. 36, 945–951 (2004).
    CAS  Article  Google Scholar 

    34.
    Giardina, C. P. & Ryan, M. G. Evidence that decomposition rates of organic carbon in mineral soil do not vary with temperature. Nature 404, 858–861 (2000).
    ADS  CAS  PubMed  Article  Google Scholar 

    35.
    Grace, J. & Rayment, M. Respiration in the balance. Nature 404, 819–820 (2000).
    CAS  PubMed  Article  Google Scholar 

    36.
    Andrews, R. T. et al. Successful embolization of collaterals from the ovarian artery during uterine artery embolization for fibroids: A case report. J. Vasc. Interv. Radiol. 11, 607–610 (2000).
    CAS  PubMed  Article  Google Scholar 

    37.
    Fu, G., Shen, Z. X., Zhang, X. Z. & Zhou, Y. T. Response of soil microbial biomass to short-term experimental warming in alpine meadow on the Tibetan Plateau. Appl. Soil Ecol. 61, 158–160 (2012).
    Article  Google Scholar 

    38.
    Chang, S. X., Shi, Z. & Thomas, B. R. Soil respiration and its temperature sensitivity in agricultural and afforested poplar plantation systems in northern Alberta. Biol. Fertil. Soils 52, 629–641 (2016).
    CAS  Article  Google Scholar 

    39.
    Uribe, C. et al. Effect of wildfires on soil respiration in three typical Mediterranean forest ecosystems in Madrid, Spain. Plant Soil 369, 403–420 (2013).
    CAS  Article  Google Scholar 

    40.
    Bergaust, L., Mao, Y. J., Bakken, L. R. & Frostegard, A. Denitrification response patterns during the transition to anoxic respiration and posttranscriptional effects of suboptimal pH on nitrogen oxide reductase in Paracoccusdenitrificans. Appl. Environ. Microbiol. 76, 6387–6396 (2010).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    41.
    Mu, Z. et al. Linking N2O emission to soil mineral N as estimated by CO2 emission and soil C/N ratio. Soil Biol. Biochem. 41, 2593–2597 (2009).
    CAS  Article  Google Scholar 

    42.
    Fu, G. et al. Response of soil respiration to grazing in an alpine meadow at three elevations in Tibet. Sci. World J. 2014, 1–9 (2014).
    Google Scholar 

    43.
    Yu, C. Q., Han, F. S. & Fu, G. Effects of 7 years experimental warming on soil bacterial and fungal community structure in the Northern Tibet alpine meadow at three elevations. Sci. Total Environ. 655, 814–822 (2019).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar  More

  • in

    Development of hematopoietic syndrome mice model for localized radiation exposure

    Chemicals and reagents
    Sodium chloride (S3014), potassium chloride (P9541), sodium phosphate dibasic (S3264), potassium phosphate monobasic (P9791), potassium bicarbonate (12602), ammonium chloride (A9434), ethylene diamine tetra acetic acid di-sodium salt (E6635), bovine serum albumin (BSA) (5482), absolute ethanol (100983), methanol (34860) were purchased from Sigma–Aldrich, St. Louis, MO, United States. CD 34, Sca 1, and CBA Flex Set which contains IL-3 (558346), IL-6 (558301), TNFα (558299), IFN γ (562233), G-CSF (560152), GM-CSF (558347), IL-1α (560157), IL-1β (560232), Mouse/Rat Soluble Protein Master Buffer Kit (558266) were procured from BD Biosciences, United States. May-Grunwalds Stain (S039) and Giemsa Stain Solution (TCL083) were procured from Hi-Media India.
    1 L 1X PBS (8 g of sodium chloride, 0.2 g of potassium chloride, 1.44 g of sodium phosphate dibasic, 0.25 g of potassium phosphate monobasic to 1 L at pH 7.4), 1 L 1X RBC lysis buffer (1 g of potassium bicarbonate, 8 g of ammonium chloride and 0.03 g of di-sodium EDTA), 1% BSA in PBS, 70% ethanol in PBS, and may-grunwald-giemsa stain mixed in 3:1 ratio were prepared in the laboratory.
    Animals and groupings
    Pathogen free Strain ‘A’ male mice weighing 25.0–30.0 gm were received from Institute of Nuclear Medicine and Allied Science (INMAS) experimental animal facility at the age of 9 weeks, which is equivalent to young adult humans. Six mice per cage were housed under 25 ± 3 °C temperature and relative humidity of 30–70% in 12 h light/dark cycle with standard food and water. The study design strictly adhered to the guidelines approved by Institutional Animal Care and Use Committee (IACUC) of our institute, Institute of Nuclear Medicine and Allied Sciences (INMAS), Defence Research and Development Organization (DRDO), Delhi, India (Institute Animals Ethics Committee number: INM/IAEC/2019/01).
    After 1 week of acclimatization under ambient conditions, mice were grouped for the survival assay according to localized gamma radiation dose i.e. 7, 9, 10, 12, 15, 17 and 19 Gy. Eighteen animals were assigned in each group for this assay.
    For all other parameters mice were divided into two groups with six animals in each. Group 1, consisted of untreated or sham irradiated animals. Group 2 was exposed to 15 Gy localized irradiation. Mice were sacrificed at different time points (1, 2, 4, 7, 10, 15, 20, 25, and 30 days) post treatment. Hematopoietic stem cell marker CD 34 and Sca1 were measured on 1, 4, 7, and 10th day post-irradiation.
    Irradiation
    A Cobalt-60 teletherapy (Model: Bhabhatron-II, Panacea Medical Technologies Pvt Ltd, India) machine was used to irradiate the hinds of the three mice at a time (Fig. 1). The cobalt-60 beam was calibrated following IAEA TRS-398 protocol, with measurement done in water phantom (30 cm × 30 cm × 30 cm), for a field size 10 cm × 10 cm at source to surface distance (SSD) of 80 cm. A Farmer type, 0.6 cc volume, ionization chamber was utilised and the same was recently calibrated for absorbed dose to water (NDW) at Secondary Standard Dosimetry Laboratory (SSDL) of India. The measured absorbed dose rate (water, 10 cm × 10 cm, 80 SSD, Dmax) was 1.39 Gy/min.
    Figure 1

    Representative image of animal exposure with γ-rays in a field size of 2 cm × 35 cm.

    Full size image

    The localized bone marrow irradiation was done using single Posteroanterior (PA) beam with 2 cm × 35 cm field size at 80 cm SSD opened through secondary collimators of the machine. The 5 mm thick acrylic re-strainers were used to immobilize each mouse which also provides the build-up thickness for the irradiation. The dose prescription point was the surface of the hinds i.e. 5 mm below the surface of the re-strainer. The whole irradiation platform was placed on 5 cm thick acrylic slab to provide sufficient back scatter and all the air gaps between the three mice were filled with tissue equivalent gel bolus or wet cotton to compensate for missing side scatter. The prescription doses were absorbed dose to water and not corrected for any tissue heterogeneity.
    The absolute and relative dosimetry of collimated field (2 cm × 35 cm) were performed using pin point (0.04 cc volume) ionization chamber and EBT3 (ISP, Wayne, USA) Gafchromic dosimetry films in solid water phantom. The response of the pin point chamber and EBT3 films was calibrated against the SSDL calibrated ionization chamber in solid water phantom at the depth of 5 cm and field size 10 cm × 10 cm in the same beam quality (i.e. cobalt-60). The film scanning was performed using Epson 10000XL flatbed scanner as per the manufacturers scanning protocol. The beam profiles were obtained from the scanned films and the full width at half maximum (FWHM) in transverse direction was 1.78 cm for the opening of 2.0 cm. The dose uniformity was better than 4% in central 80% of the field. The absolute dose rate (dose to water, dmax, 2 cm × 35 cm, SSD 80 cm) measured at the centre of the field was 1 Gy/min. The overall uncertainty in dose calculations is 6.5% with enhancement or coverage factor (k = 2) at 95% confidence. The main body of the mice was covered with 15 mm lead to prevent any scatter radiation reaching at other body parts of the animal. The surface dose reaching at the shielded body parts of the mice was measured using 0.4 cc ionisation chamber. The measured mean surface dose was 0.024 ± 0.014% of the open field dose. This dose is negligible and considered to be acceptable.
    Survival assay in mice
    For survival assay, 18 mice from each experimental group i.e. 7, 9, 10, 12, 15, 17 and 19 Gy, were irradiated at a fixed dose rate of 1 Gy/min. Groups were inspected daily for their morbidity and mortality status for a total duration of 30 days. All the surviving mice were euthanized humanely using intra-peritoneal injection (i.p.) with dexmedetomidine (0.3 mg/kg) in combination with ketamine (190 mg/kg) at the completion of the observation period.
    Validation of hematopoietic syndrome model
    The model of hematopoietic syndrome was validated with localized radiation of 15 Gy as this was the maximum dose at which all the animals survived as observed in the survival results. Mice were euthanized using intra-peritoneal injection (i.p.) with dexmedetomidine (0.3 mg/kg) in combination with ketamine (190 mg/kg) at 1, 2, 4, 7, 10, 15, 20, 25, and 30th day time points. Further analysis such as organ weight, cell counting, and serum cytokines measurement assays were conducted on various organs such as femur bone, spleen, thymus, lungs, kidneys (L), kidneys (R), liver and heart and blood at above mentioned time points. Hematopoietic stem cell markers (CD 34 and Sca 1) were measured in bone marrow cells of mice at 1, 4, 7, and 10th day time points.
    Body and organ weights
    The control and 15 Gy locally irradiated group animals were sacrificed at 1, 4, 7, 10, 15, 20, 25, and 30th day time points. The organs (spleen, thymus, lungs, kidneys (L), kidneys (R), liver and heart) were dissected out and weighed individually. Relative organ weight was calculated as the ratio between organ weight and body weight.
    Haematology
    The blood samples were collected from untreated and 15 Gy localized radiation group at 1, 4, 7, 10, 15, 20, 25, and 30th day time points in EDTA vial by cardiac puncture. 20 µl of collected blood from each experimental animal was taken for haematology and analysed using Nihon Kohden Celltac α, Tokyo, Japan, a fully automatic 3 part haematology analyzer.
    Bone marrow cell counting
    Total bone marrow cells were flushed out at 1, 4, 7, 10, 15, 20, 25 and 30th day points in 1 ml PBS from both femurs of the untreated mice and 15 Gy localized radiation group. The PBS containing cells was centrifuged at 2000 rpm for 8 min and the supernatant was discarded. The cells were washed with 1 ml PBS and the cell pellet was re-suspended in 1 ml of PBS. The re-suspended sample was diluted as required. The cells were then counted using improved Neubauer chamber under Olympus BX-63 microscope.
    Bone marrow smears study
    Bone marrow cells collected at 1, 4, 7, 10, 15, 20, 25, and 30th day post-irradiation from different treatment groups were centrifuged at 2000 rpm for 8 min and re-suspended in 50 μl of Fetal Bovine Serum (FBS). A small drop of cells suspension was dropped on a clean microscopic slide and then smeared thinly over an area of 2–3 cm by pulling another glass slide held at a 45° angle. Cells were fixed with methanol and stained with may-grunwald-giemsa. Slides were then analysed using Olympus BX-63 microscope.
    Splenocytes and thymocytes counts
    The organs excised at 1, 4, 7, 10, 15, 20, 25, and 30th day time-points were cleaned in chilled PBS. Using sterile-chilled frosted slides both the thymus and spleen were minced into a cell suspension and which was further centrifuged at 2000 rpm for 8 min. After centrifugation the supernatant was discarded. The RBCs were lysed using RBC lysis buffer and the cells were washed with 1 ml PBS. The cells were then counted using improved Neubauer chamber under Olympus BX-63 microscope.
    Measurement of hematopoietic stem cell marker CD 34 and Sca1
    Bone marrow cells were isolated from both the femurs from various treatment groups at different time points. Cell were washed with chilled PBS and fixed with 70% chilled ethanol and stored overnight at – 20 °C. Following day, five million cells were blocked with 1% BSA-PBS and stained with CD 34 and Sca 1 cell-surface markers for 20 min at room temperature. 10,000 cells from each sample were analyzed using FACS LSR II (BD Biosciences, San Jose, CA) and results were visualized by the BD FACSDiva V7 software (BD Biosciences, San Jose, CA).
    Cytokines expression changes
    Mice blood samples were collected in BD serum separation tube at different time-points post treatment (1, 4, 7, 10, 15, 25, and 30 days) via. cardiac puncture from different treatment groups. Samples were incubated at room temperature for 2 h and after incubation centrifuged at 6000g for 15 min. Serum was separated and stored at − 80 °C until analysis. The Levels of IL-3, IL-6, TNFα, IFN γ, G-CSF, GM-CSF, IL-1α,and IL-1β cytokines in mice serum were determined by using respective BD Cytometric Bead Array (CBA) Flex Set (BD Biosciences, USA) on dual laser flow analyzer (LSR-II, Becton Dickinson Biosciences, USA) according to manufacturer’s instructions. Results were analysed using FCAP Array V3 software (BD Biosciences, San Jose, CA).
    Statistical analysis
    All values obtained in our results are represented as mean ± SD of three independent replicates. The 30 day survival data was plotted using Kaplan–Meier analysis. The difference between the experimental groups was evaluated by one-way analysis of variance, with Newman–Keuls multiple comparison test (V, 8.01; GraphPad Prism, San Diego, CA, USA). The comparisons were made among the untreated and 15 Gy locally irradiated groups for all experimental parameters except the survival study. A value of p  More

  • in

    Urban life promotes delayed dispersal and family living in a non-social bird species

    1.
    Lawrence, W. S. Dispersal: an alternative mating tactic conditional on sex ratio and body size. Behav. Ecol. Sociobiol. 21, 367–373 (1987).
    Article  Google Scholar 
    2.
    Russell, E. M. & Rowley, I. Philopatry or dispersal: competition for territory vacancies in the splendid fairy-wren, Malurus splendens. Anim. Behav. 45, 519–539 (1993).
    Article  Google Scholar 

    3.
    Herzig, A. L. Effects of population density on long-distance dispersal in the goldenrod beetle Trirhabda virgata. Ecology 76, 2044–2054 (1995).
    Article  Google Scholar 

    4.
    Verhulst, S., Perrins, C. M. & Riddington, R. Natal dispersal of great tits in a patchy environment. Ecology 78, 864–872 (1997).
    Article  Google Scholar 

    5.
    Johnson, M. L. & Gaines, M. S. Evolution of dispersal: theoretical models and empirical tests using birds and mammals. Annu. Rev. Ecol. Evol. Sys. 21, 449–480 (1990).
    Article  Google Scholar 

    6.
    Clobert, J., Danchin, E., Dhondt, A. A. & Nichols, J. Dispersal (Oxford University Press, Oxford, 2001).
    Google Scholar 

    7.
    Bowler, D. E. & Benton, T. G. Causes and consequences of animal dispersal strategies: relating individual behaviour to spatial dynamics. Biol. Rev. 80, 205–225 (2005).
    PubMed  Article  PubMed Central  Google Scholar 

    8.
    Szulkin, M. & Sheldon, B. C. Dispersal as a means of inbreeding avoidance in a wild bird population. Proc. R. Soc. B 275, 703–711 (2008).
    PubMed  Article  PubMed Central  Google Scholar 

    9.
    Cornwallis, C. K. Cooperative breeding and the evolutionary coexistence of helper and nonhelper strategies. Proc. Nat. Acad. Sc. USA 115, 1684–1686 (2018).
    CAS  Article  Google Scholar 

    10.
    Nelson-Flower, M. J., Ridley, A. R., Wiley, E. M. & Flower, T. P. Individual dispersal delays in a cooperative breeder: ecological constraints, the benefits of philopatry and the social queue for dominance. J. Anim. Ecol. 87, 1227–1238 (2018).
    PubMed  Article  PubMed Central  Google Scholar 

    11.
    Komdeur, J. Importance of habitat saturation and territory quality for evolution of cooperative breeding in the Seychelles warbler. Nature 358, 493 (1992).
    ADS  Article  Google Scholar 

    12.
    Kokko, H. & Lundberg, P. Dispersal, migration, and offspring retention in saturated habitats. Am. Nat. 157, 188–202 (2001).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    13.
    Kokko, H. & Ekman, J. Delayed dispersal as a route to breeding: territorial inheritance, safe havens, and ecological constraints. Am. Nat. 160, 468–484 (2002).
    PubMed  Article  PubMed Central  Google Scholar 

    14.
    Lemel, J. Y., Belichon, S., Clobert, J. & Hochberg, M. E. The evolution of dispersal in a two-patch system: some consequences of differences between migrants and residents. Evol. Ecol. 11, 613–629 (1997).
    Article  Google Scholar 

    15.
    Forero, M. G., Donázar, J. A. & Hiraldo, F. Causes and fitness consequences of natal dispersal in a population of black kites. Ecology 83, 858–872 (2002).
    Article  Google Scholar 

    16.
    Kokko, H. & López-Sepulcre, A. From individual dispersal to species ranges: perspectives for a changing world. Science 313, 789–791 (2006).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    17.
    Bonte, D. et al. Costs of dispersal. Biol. Rev. 87, 290–312 (2012).
    PubMed  Article  PubMed Central  Google Scholar 

    18.
    Serrano, D. & Tella, J. L. Lifetime fitness correlates of natal dispersal distance in a colonial bird. J. Anim. Ecol. 81, 97–107 (2012).
    PubMed  Article  PubMed Central  Google Scholar 

    19.
    Kubisch, A., Holt, R. D., Poethke, H. J. & Fronhofer, E. A. Where am I and why? Synthesizing range biology and the eco-evolutionary dynamics of dispersal. Oikos 123, 5–22 (2014).
    Article  Google Scholar 

    20.
    Drobniak, S. M., Wagner, G., Mourocq, E. & Griesser, M. Family living: an overlooked but pivotal social system to understand the evolution of cooperative breeding. Behav. Ecol. 26, 805–811 (2015).
    Article  Google Scholar 

    21.
    Taborsky, M. Sneakers, satellites, and helpers: parasitic and cooperative behavior in fish reproduction. Adv. Study Behav. 23, e100 (1994).
    Google Scholar 

    22.
    Clutton-Brock, T. H. & Lukas, D. The evolution of social philopatry and dispersal in female mammals. Mol. Ecol. 21, 472–492 (2012).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    23.
    Pruett-Jones, S. G. & Lewis, M. J. Sex ratio and habitat limitation promote delayed dispersal in superb fairy-wrens. Nature 348, 541–542 (1990).
    ADS  Article  Google Scholar 

    24.
    Bergmüller, R., Heg, D. & Taborsky, M. Helpers in a cooperatively breeding cichlid stay and pay or disperse and breed, depending on ecological constraints. Proc. R. Soc. B 272, 325–331 (2005).
    PubMed  Article  PubMed Central  Google Scholar 

    25.
    Carrete, M., Donázar, J. A., Margalida, A. & Bertran, J. Linking ecology, behaviour and conservation: does habitat saturation change the mating system of bearded vultures?. Biol. Lett. 2, 624–627 (2006).
    PubMed  PubMed Central  Article  Google Scholar 

    26.
    Covas, R., Doutrelant, C. & du Plessis, M. A. Experimental evidence of a link between breeding conditions and the decision to breed or to help in a colonial cooperative bird. Proc. R. Soc. B 271, 827–832 (2004).
    PubMed  Article  PubMed Central  Google Scholar 

    27.
    Baglione, V. et al. Does yearround territoriality rather than habitat saturation explain delayed natal dispersal and cooperative breeding in the carrion crow?. J. Anim. Ecol. 74, 842–851 (2005).
    Article  Google Scholar 

    28.
    Hatchwell, B. J. & Komdeur, J. Ecological constraints, life history traits and the evolution of cooperative breeding. Anim. Behav. 59, 1079–1086 (2000).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    29.
    Pen, I. & Weissing, F. J. Towards a unified theory of cooperative breeding: the role of ecology and life history re-examined. Proc. R. Soc. B 267, 2411–2418 (2000).
    Article  Google Scholar 

    30.
    Covas, R. & Griesser, M. Life history and the evolution of family living in birds. Proc. R. Soc. B 274, 1349–1357 (2007).
    PubMed  Article  PubMed Central  Google Scholar 

    31.
    Griesser, M. & Barnaby, J. Role of Nepotism Cooperation and Competition in the Avian Families. Birds-Evolution, Behavior and Ecology Series (Nova Science Pub Inc, New York, 2010).
    Google Scholar 

    32.
    Baglione, V., Canestrari, D., Marcos, J. M., Griesser, M. & Ekman, J. History, environment and social behaviour: experimentally induced cooperative breeding in the carrion crow. Proc. R. Soc B 269, 1247–1251 (2002).
    PubMed  Article  PubMed Central  Google Scholar 

    33.
    Tchabovsky, A. & Bazykin, G. Females delay dispersal and breeding in a solitary gerbil, Meriones tamariscinus. J. Mammal. 85, 105–112 (2004).
    Article  Google Scholar 

    34.
    Ellis, E. C., Fuller, D. Q., Kaplan, J. O. & Lutters, W. G. Dating the Anthropocene: Towards an empirical global history of human transformation of the terrestrial biosphere. Elem. Sci. Anth 1, 18 (2013).
    Article  Google Scholar 

    35.
    Bernardo-Madrid, R. et al. Human activity is altering the world’s zoogeographical regions. Ecol. Lett. 22, 1297–1305 (2019).
    PubMed  PubMed Central  Google Scholar 

    36.
    Rodríguez-Martínez, S., Carrete, M., Roques, S., Rebolo-Ifrán, N. & Tella, J. L. High urban breeding densities do not disrupt genetic monogamy in a bird species. PLoS ONE 9, e91314 (2014).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    37.
    McKinney, M. L. Urbanization as a major cause of biotic homogenization. Biol. Conserv. 127, 247–260 (2006).
    Article  Google Scholar 

    38.
    Grimm, N. B. et al. Global change and the ecology of cities. Science 319, 756–760 (2008).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    39.
    McKinney, M. L. & Lockwood, J. L. Biotic homogenization: a few winners replacing many losers in the next mass extinction. Trends Ecol. Evol. 14, 450–453 (1999).
    CAS  PubMed  Article  Google Scholar 

    40.
    Sol, D., González-Lagos, C., Moreira, D., Maspons, J. & Lapiedra, O. Urbanisation tolerance and the loss of avian diversity. Ecol. Lett. 17, 942–950 (2014).
    PubMed  Article  PubMed Central  Google Scholar 

    41.
    Carrete, M. & Tella, J. L. Inter-individual variability in fear of humans and relative brain size of the species are related to contemporary urban invasion in birds. PLoS ONE 6, e18859 (2011).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    42.
    Díaz, M. et al. The geography of fear: a latitudinal gradient in anti-predator escape distances of birds across Europe. PLoS ONE 8, e64634 (2013).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    43.
    Oro, D., Genovart, M., Tavecchia, G., Fowler, M. S. & Martínez-Abraín, A. Ecological and evolutionary implications of food subsidies from humans. Ecol. Lett. 16, 1501–1514 (2013).
    PubMed  Article  PubMed Central  Google Scholar 

    44.
    Marzluff, J. M. Worldwide urbanization and its effects on birds. Avian ecology and conservation in an urbanizing world (pp. 19–47) (Springer, Boston, MA. 2001).

    45.
    Haskell, D. G., Knupp, A. M. & Schneider, M. C. Nest predator abundance and urbanization. Avian ecology and conservation in an urbanizing world (pp. 243–258). Springer, Boston, MA. (2001).

    46.
    Luna, Á., Romero-Vidal, P., Hiraldo, F. & Tella, J. L. Cities may save some threatened species but not their ecological functions. PeerJ 6, e4908 (2018).
    PubMed  PubMed Central  Article  Google Scholar 

    47.
    Muhly, T. B., Semeniuk, C., Massolo, A., Hickman, L. & Musiani, M. Human activity helps prey win the predator–prey space race. PLoS ONE 6, e17050 (2011).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    48.
    Rebolo-Ifrán, N., Tella, J. L. & Carrete, M. Urban conservation hotspots: predation release allows the grassland-specialist burrowing owl to perform better in the city. Sci. Rep. 7, 3527 (2017).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    49.
    Tuomainen, U. & Candolin, U. Behavioural responses to human-induced environmental change. Biol. Rev. 86, 640–657 (2011).
    PubMed  Article  PubMed Central  Google Scholar 

    50.
    Rodewald, A. D., Kearns, L. J. & Shustack, D. P. Anthropogenic resource subsidies decouple predator-prey relationships. Ecol. Appl. 21, 936–943 (2011).
    PubMed  Article  PubMed Central  Google Scholar 

    51.
    Sih, A., Ferrari, M. C. & Harris, D. J. Evolution and behavioural responses to human-induced rapid environmental change. Evol. Appl. 4, 367–387 (2011).
    PubMed  PubMed Central  Article  Google Scholar 

    52.
    Carrete, M. & Tella, J. L. Behavioural correlations associated with fear of humans differ between rural and urban burrowing owls. Front. Ecol. Evol. 5, 54 (2017).
    Article  Google Scholar 

    53.
    Moore, J. A., Kamarainen, A. M., Scribner, K. T., Mykut, C. & Prince, H. H. The effects of anthropogenic alteration of nesting habitat on rates of extra-pair fertilization and intraspecific brood parasitism in Canada geese branta Canadensis. Ibis 154, 354–362 (2012).
    Article  Google Scholar 

    54.
    Ryder, T. B., Fleischer, R. C., Shriver, W. G. & Marra, P. P. The ecological–evolutionary interplay: density-dependent sexual selection in a migratory songbird. Ecol. Evol. 2, 976–987 (2012).
    PubMed  PubMed Central  Article  Google Scholar 

    55.
    Luna, Á., Palma, A., Sánz-Aguilar, A., Tella, J. L. & Carrete, M. Personality-dependent breeding dispersal in rural but not urban burrowing owls. Sci. Rep. 9, 2886 (2019).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    56.
    Luna, Á., Palma, A., Sánz-Aguilar, A., Tella, J. L. & Carrete, M. Sex, personality and conspecific density influence natal dispersal with lifetime fitness consequences in urban and rural burrowing owls. PLoS ONE 15, e0226089 (2020).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    57.
    Mueller, J. C. et al. Evolution of genomic variation in the burrowing owl in response to recent colonization of urban areas. Proc. R. Soc. B 285, 20180206 (2018).
    PubMed  Article  PubMed Central  Google Scholar 

    58.
    Miles, L. S., Rivkin, L. R., Johnson, M. T., Munshi-South, J. & Verrelli, B. C. Gene flow and genetic drift in urban environments. Mol. Ecol. 28, 4138–4151 (2019).
    PubMed  Article  PubMed Central  Google Scholar 

    59.
    Henger, C. S. et al. Genetic diversity and relatedness of a recently established population of eastern coyotes (Canis latrans) in New York City. Urban Ecosyst. 23, 1–12 (2019).
    Google Scholar 

    60.
    Carrete, M. & Tella, J. L. Individual consistency in flight initiation distances in burrowing owls: a new hypothesis on disturbance-induced habitat selection. Biol. Lett. 6, 167–170 (2010).
    PubMed  Article  PubMed Central  Google Scholar 

    61.
    Cockburn, A. Evolution of helping behavior in cooperatively breeding birds. Annu. Rev. Ecol. Syst. 29, 141–177 (1998).
    Article  Google Scholar 

    62.
    Clutton-Brock, T. Breeding together: kin selection and mutualism in cooperative vertebrates. Science 296, 69–72 (2002).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    63.
    Browning, L. E. et al. Career provisioning rules in an obligate cooperative breeder: prey type, size and delivery rate. Behav. Ecol. Sociobiol. 66, 1639–1649 (2012).
    Article  Google Scholar 

    64.
    Richardson, D. S., Burke, T. & Komdeur, J. Direct benefits and the evolution of female-biased cooperative breeding in Seychelles warblers. Evolution 56, 2313–2321 (2002).
    PubMed  Article  PubMed Central  Google Scholar 

    65.
    Gamero, A., Székely, T. & Kappeler, P. M. Delayed juvenile dispersal and monogamy, but no cooperative breeding in white-breasted mesites (Mesitornis variegata). Behav. Ecol. Sociobiol. 68, 73–83 (2014).
    Article  Google Scholar 

    66.
    Brown, J. L. Territorial behavior and population regulation in birds: a review and re-evaluation. Wilson Bull. 81, 293–329 (1969).
    Google Scholar 

    67.
    Emlen, S. T. The evolution of helping. I. An ecological constraints model. Am. Nat. 119, 29–39 (1982).
    Article  Google Scholar 

    68.
    Emlen, S. T. An evolutionary theory of the family. Proc. Natl. Acad. Sci. USA 92, 8092–8099 (1995).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    69.
    Ekman, J., Eggers, S., Griesser, M. & Tegelström, H. Queuing for preferred territories: delayed dispersal of Siberian jays. J. Anim. Ecol. 70(2), 317–324. https://doi.org/10.1111/j.1365-2656,2001.00490.x (2001).
    Article  Google Scholar 

    70.
    Baglione, V., Canestrari, D., Marcos, J. M., Griesser, M. & Ekman, J. History, environment and social behaviour: experimentally induced cooperative breeding in the carrion crow. Proc. R. Soc. B 269, 1247–1251 (2002).
    PubMed  Article  PubMed Central  Google Scholar 

    71.
    Gardner, J. L., Magrath, R. D. & Kokko, H. Stepping stones of life: natal dispersal in the group-living but noncooperative speckled warbler. Anim. Behav. 66, 521–530 (2003).
    Article  Google Scholar 

    72.
    Russell, E. M., Yom-Tov, Y. & Geffen, E. Extended parental care and delayed dispersal: northern, tropical, and southern passerines compared. Behav. Ecol. 15, 831–838 (2004).
    Article  Google Scholar 

    73.
    Lucia, K. E. et al. Philopatry in prairie voles: an evaluation of the habitat saturation hypothesis. Behav. Ecol. 19, 774–783 (2008).
    Article  Google Scholar 

    74.
    Clobert, J., Le Galliard, J. F., Cote, J., Meylan, S. & Massot, M. Informed dispersal, heterogeneity in animal dispersal syndromes and the dynamics of spatially structured populations. Ecol. Lett. 12, 197–220 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    75.
    Bocedi, G., Heinonen, J. & Travis, J. M. Uncertainty and the role of information acquisition in the evolution of context-dependent emigration. Am. Nat. 179, 606–620 (2012).
    PubMed  Article  PubMed Central  Google Scholar 

    76.
    Bergmüller, R., Taborsky, M., Peer, K. & Heg, D. Extended safe havens and between-group dispersal of helpers in a cooperatively breeding cichlid. Behaviour 142, 1643–1667 (2005).
    Article  Google Scholar 

    77.
    Tanaka, H., Frommen, J. G., Engqvist, L. & Kohda, M. Task-dependent workload adjustment of female breeders in a cooperatively breeding fish. Behav. Ecol. 29, 221–229 (2018).
    Article  Google Scholar 

    78.
    Suárez-Seoane, S., Osborne, P. E. & Alonso López, J. C. Large-scale habitat selection by agricultural steppe birds in Spain: identifying species–habitat responses using generalized additive models. J. Appl. Ecol. 39, 755–771 (2002).
    Article  Google Scholar 

    79.
    Boyce, M. S. et al. Can habitat selection predict abundance?. J. Anim. Ecol. 85, 11–20 (2016).
    PubMed  Article  PubMed Central  Google Scholar 

    80.
    Muller, K. L., Stamps, J. A., Krishnan, V. V. & Willits, N. H. The effects of conspecific attraction and habitat quality on habitat selection in territorial birds (Troglodytes aedon). Am. Nat. 150, 650–661 (1997).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    81.
    Danchin, E., Boulinier, T. & Massot, M. Conspecific reproductive success and breeding habitat selection: implications for the study of coloniality. Ecology 79, 2415–2428 (1998).
    Article  Google Scholar 

    82.
    Brown, C. R., Brown, M. B. & Danchin, E. Breeding habitat selection in cliff swallows: The effect of conspecific reproductive success on colony choice. J. Anim. Ecol. 69, 133–142 (2000).
    Article  Google Scholar 

    83.
    Danchin, E., Giraldeau, L. A., Valone, T. J. & Wagner, R. H. Public information: from nosy neighbors to cultural evolution. Science 305, 487–491 (2004).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    84.
    Serrano, D., Forero, M. G., Donázar, J. A. & Tella, J. L. Dispersal and social attraction affect colony selection and dynamics of lesser kestrels. Ecology 85, 3438–3447 (2004).
    Article  Google Scholar 

    85.
    Farrell, S. L., Morrison, M. L., Campomizzi, A. J. & Wilkins, R. N. Conspecific cues and breeding habitat selection in an endangered woodland warbler. J. Anim. Ecol. J. Anim. Ecol. 81, 1056–1064 (2012).
    PubMed  Article  PubMed Central  Google Scholar 

    86.
    Pärt, T. The importance of local familiarity and search costs for age-and sex-biased philopatry in the collared flycatcher. Anim. Behav. 49, 1029–1038 (1995).
    Article  Google Scholar 

    87.
    Clarke, A. L., Sæther, B. E. & Røskaft, E. Sex biases in avian dispersal: a reappraisal. Oikos 79, 429–438 (1997).
    Article  Google Scholar 

    88.
    Piper, W. H., Walcott, C., Mager, J. N. & Spilker, F. J. Nestsite selection by male loons leads to sex-biased site familiarity. J. Anim. Ecol. J. Anim. Ecol. 77, 205–210 (2008).
    PubMed  Article  PubMed Central  Google Scholar 

    89.
    Stacey, P. B. & Ligon, J. D. The benefits-of-philopatry hypothesis for the evolution of cooperative breeding: variation in territory quality and group size effects. Am. Nat. 137, 831–846 (1991).
    Article  Google Scholar 

    90.
    Goldstein, J. M., Woolfenden, G. E. & Hailman, J. P. A same-sex stepparent shortens a prebreeder’s duration on the natal territory: tests of two hypotheses in Florida scrub-jays. Behav. Ecol. Sociobiol. 44, 15–22 (1998).
    Article  Google Scholar 

    91.
    Ekman, J. & Griesser, M. Why offspring delay dispersal: experimental evidence for a role of parental tolerance. Proc. R. Soc. B 269, 1709–1713 (2002).
    PubMed  Article  PubMed Central  Google Scholar 

    92.
    Griesser, M. & Ekman, A. Nepotistic alarm calling in the Siberian jay, Perisoreus infaustus. Anim. Behav. 67, 933–939 (2004).
    Article  Google Scholar 

    93.
    Griesser, M. Referential calls signal predator behavior in a group-living bird species. Curr. Biol. 18, 69–73 (2008).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    94.
    Emlen, S. T. & Wrege, P. H. Breeding biology of white-fronted bee-eaters at Nakuru: the influence of helpers on breeder fitness. J. Anim. Ecol. 60, 309–326 (1991).
    Article  Google Scholar 

    95.
    Cockburn, A. et al. Can we measure the benefits of help in cooperatively breeding birds: the case of superb fairy-wrens Malurus cyaneus?. J. Anim. Ecol. J. Anim. Ecol. 77, 430–438 (2008).
    PubMed  Article  PubMed Central  Google Scholar 

    96.
    Kingma, S. A., Hall, M. L., Arriero, E. & Peters, A. Multiple benefits of cooperative breeding in purple-crowned fairy-wrens: a consequence of fidelity?. J. Anim. Ecol. 79, 757–768 (2010).
    PubMed  PubMed Central  Google Scholar 

    97.
    Fitch, M. A. & Shugart, G. W. Comparative biology and behavior of monogamous pairs and one male-two female trios of Herring Gulls. Behav. Ecol. Sociobiol. 14, 1–7 (1983).
    Article  Google Scholar 

    98.
    Del Hoyo, J., Elliot, A. & Sargatal, J. Handbook of the Birds of the World. Barn Owls to Hummingbirds. Vol. 5.Lynx Edicions, Barcelona (1999).

    99.
    Clayton, K. M. & Schmutz, J. K. Is the decline of Burrowing Owls Speotyto cunicularia in prairie Canada linked to changes in Great Plains ecosystems?. Bird Conserv. Int. 9, 163–185 (1999).
    Article  Google Scholar 

    100.
    Haug, E. A., Millsap, B. A. & Martell, M. S. The burrowing owl (Speotyto cunicularia). In Poole, A. & F. Gill (editors). The birds of North America. Philadelphia: The Academy of Natural Sciences; Washington, D. C. The American Ornithologists’ Union. Washington, D. C. The American Ornithologists’ Union (1993).

    101.
    Carrete, M. & Tella, J. L. High individual consistency in fear of humans throughout the adult lifespan of rural and urban burrowing owls. Sci. Rep. 3, 3524 (2013).
    ADS  PubMed  PubMed Central  Article  Google Scholar 

    102.
    Rebolo-Ifrán, N. et al. Links between fear of humans, stress and survival support a non-random distribution of birds among urban and rural habitats. Sci. Rep. 5, 13723 (2015).
    ADS  PubMed  PubMed Central  Article  Google Scholar 

    103.
    Kalinowski, S. T., Wagner, A. P. & Taper, M. L. ML-Relate: Software for estimating relatedness and relationship from multilocus genotypes. Mol. Ecol. Notes 6, 576–579 (2006).
    CAS  Article  Google Scholar 

    104.
    Kalinowski, S. T., Taper, M. L. & Marshall, T. C. Revising how the computer program CERVUS accommodates genotyping error increases success in paternity assignment. Mol. Ecol. 16, 1099–1106 (2007).
    PubMed  Article  PubMed Central  Google Scholar 

    105.
    Grueber, C. E., Nakagawa, S., Laws, R. J. & Jamieson, I. G. Multimodel inference in ecology and evolution: challenges and solutions. J. Evol. Biol. 24, 699–711 (2011).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    106.
    Oliehoek, P. A., Windig, J. J., Van Arendonk, J. A. & Bijma, P. Estimating relatedness between individuals in general populations with a focus on their use in conservation programs. Genetics 173, 483–496 (2006).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    107.
    Laake, J. L. RMark: an R interface for analysis of capture-recapture data with MARK (2013).

    108.
    Zar, J. H. Statistical significance of mutation frequencies, and the power of statistical testing, using the Poisson distribution. Biometr. J. 26, 83–88 (1984).
    Article  Google Scholar 

    109.
    Labocha, M. K. & Hayes, J. P. Morphometric indices of body condition in birds: a review. J. Ornithol. 153, 1–22 (2012).
    Article  Google Scholar 

    110.
    Choquet, R., Lebreton, J. D., Gimenez, O., Reboulet, A. M. & Pradel, R. U-CARE: Utilities for performing goodness of fit tests and manipulating CApture–REcapture data. Ecography 32, 1071–1074 (2009).
    Article  Google Scholar 

    111.
    Choquet, R., Rouan, L., Pradel, R. Program E-SURGE: a software application for fitting multievent models. In Modeling demographic processes in marked populations. pp. 845–865. Springer, Boston, MA, 2009.

    112.
    White, G. C. & Burnham, K. P. Program MARK: survival estimation from populations of marked animals. Bird Study 46, S120–S139 (1999).
    Article  Google Scholar 

    113.
    Hadfield, J. D. MCMC methods for multi-response generalized linear mixed models: the MCMCglmm R package. J. Stat. Softw. 33, 1–22 (2010).
    Article  Google Scholar 

    114.
    Mollie, E. B. et al. glmmTMB balances speed and flexibility among Packages for zeroinflated generalized linear mixed modeling. R J. 9, 378–400 (2017).
    Article  Google Scholar 

    115.
    Hartig F. DHARMa: Residual Diagnostics for Hierarchical (Multi-Level/Mixed) Regression Models. R Package. 2018. https://www.cran.r-project.org/package=DHARMa, Version 0. 2. 0.

    116.
    Burnham, K. P. & Anderson, D. R. Model Selection and Multimodel Inference. A Practical Information-Theoretic Approach (Springer, New York, 2002).
    Google Scholar 

    117.
    Barton, K. MuMIn: Multi-Model Inference. R package version 1. 40. 0. https://CRANhttps://CRAN.R-project.org/package=MuMIn (2017). More

  • in

    The spreading of the invasive sacred ibis in Italy

    1.
    Wittenberg, R. & Cock, M. J. W. Invasive Alien Species: A Toolkit of Best Prevention and Management Practices (CABI, Wallingford, 2001).
    Google Scholar 
    2.
    Genovesi, P. Eradications of invasive alien species in Europe: a review. Biol. Invas. 7, 127–133 (2005).
    Article  Google Scholar 

    3.
    Butchart, S. H. M. et al. Global biodiversity: indicators of recent declines. Science 328, 1164–1168 (2010).
    ADS  CAS  PubMed  Article  Google Scholar 

    4.
    Pejchar, L. & Mooney, H. A. Invasive species, ecosystem services and human well-being. Trends Ecol. Evol. 24, 497–504 (2009).
    PubMed  Article  Google Scholar 

    5.
    Vilà, M. et al. How well do we understand the impacts of alien species on ecosystem services? A pan-European, cross-taxa assessment. Front. Ecol. Environ. 8, 135–144 (2010).
    Article  Google Scholar 

    6.
    David, P. et al. Impacts of invasive species on food webs: a review of empirical data. In Advances in Ecological Research (eds Bohan, D. A. et al.) 1–60 (Academic Press, Cambridge, 2017).
    Google Scholar 

    7.
    Doherty, T. S., Glen, A. S., Nimmo, D. G., Ritchie, E. G. & Dickman, C. R. Invasive predators and global biodiversity loss. Proc. Natl. Acad. Sci. 113, 11261–11265 (2016).
    CAS  PubMed  Article  Google Scholar 

    8.
    Blackburn, T. M. & Ewen, J. G. Parasites as drivers and passengers of human-mediated biological invasions. EcoHealth 14, 61–73 (2017).
    PubMed  Article  Google Scholar 

    9.
    Volponi, S. & Emiliani, D. Presenza e riproduzione di due specie di Ciconiiformi esotici con formazione di ibridi nelle zone umide ravennati del Parco del Delta del Po. Convegno Inanellatori Ital. 10, 1 (2008).
    Google Scholar 

    10.
    Negri, A. et al. Mitochondrial DNA and microsatellite markers evidence a different pattern of hybridization in red-legged partridge (Alectoris rufa) populations from NW Italy. Eur. J. Wildl. Res. 59, 407–419 (2013).
    Article  Google Scholar 

    11.
    Raffini, F. et al. From nucleotides to satellite imagery: approaches to identify and manage the invasive pathogen Xylella fastidiosa and its insect vectors in Europe. Sustainability 12, 4508 (2020).
    CAS  Article  Google Scholar 

    12.
    Little, M. N. et al. The influence of riparian invasion by the terrestrial shrub Lonicera maackii on aquatic macroinvertebrates in temperate forest headwater streams. Biol. Invas. https://doi.org/10.1007/s10530-020-02349-8 (2020).
    Article  Google Scholar 

    13.
    Falk-Petersen, J., Bøhn, T. & Sandlund, O. T. On the numerous concepts in invasion biology. Biol. Invas. 8, 1409–1424 (2006).
    Article  Google Scholar 

    14.
    Helmstedt, K. J. et al. Prioritizing eradication actions on islands: it’s not all or nothing. J. Appl. Ecol. 53, 733–741 (2016).
    Article  Google Scholar 

    15.
    Moon, K., Blackman, D. A. & Brewer, T. D. Understanding and integrating knowledge to improve invasive species management. Biol. Invas. 17, 2675–2689 (2015).
    Article  Google Scholar 

    16.
    European Commission DG ENV. More rigorous studies needed to evaluate impact of invasive birds. Sci. Environ. Policy 1 (2011).

    17.
    Strubbe, D., Shwartz, A. & Chiron, F. Concerns regarding the scientific evidence informing impact risk assessment and management recommendations for invasive birds. Biol. Conserv. 144, 2112–2118 (2011).
    Article  Google Scholar 

    18.
    Bertolino, S. et al. Spatially explicit models as tools for implementing effective management strategies for invasive alien mammals. Mamm. Rev. 50, 187–199 (2020).
    Article  Google Scholar 

    19.
    Graham, S. et al. Opportunities for better use of collective action theory in research and governance for invasive species management. Conserv. Biol. 33, 275–287 (2019).
    PubMed  PubMed Central  Article  Google Scholar 

    20.
    del Hoyo, J., Elliot, A. & Sargatal, J. Handbook of the Birds of the World (Lynx Edicions, Barcelona, 1992).
    Google Scholar 

    21.
    Kopij, G. Breeding ecology of the Sacred Ibis Threskiornis aethiopicus in the Free State, South Africa. S. Afr. J. Wildl. Res. 29, 25–30 (1999).
    Google Scholar 

    22.
    del Hoyo, J. & Collar, N. HBW and Birdlife International Illustrated Checklist of the Birds of the World. Vol: Non-passerines (2014).

    23.
    Robert, H., Lafontaine, R., Delsinne, T. & Beudels-Jamar, R. Risk analysis of the Sacred ibis Threskiornis aethiopicus (Royal Belgian Institute of Natural Sciences, Brussels, 2013).
    Google Scholar 

    24.
    Yésou, P. & Clergeau, P. Sacred Ibis: a new invasive species in Europe. Bird. World 18, 517–526 (2006).
    Google Scholar 

    25.
    Clergeau, P. & Yésou, P. Behavioural flexibility and numerous potential sources of introduction for the Sacred Ibis: causes of concern in Western Europe?. Biol. Invas. 8, 1381–1388 (2006).
    Article  Google Scholar 

    26.
    Nentwig, W., Bacher, S., Kumschick, S., Pyšek, P. & Vilà, M. More than “100 worst” alien species in Europe. Biol. Invas. 20, 1611–1621 (2018).
    Article  Google Scholar 

    27.
    Carpegna, F., Della Toffola, M., Alessandria, G. & Re, A. L’Ibis sacro Threskiornis aethiopicus nel Parco Naturale Lame del Sesia e sua presenza in Piemonte. Avocetta 23, 82 (1999).
    Google Scholar 

    28.
    Marion, L. Is the Sacred ibis a real threat to biodiversity? Long-term study of its diet in non-native areas compared to native areas. C. R. Biol. 336, 207–220 (2013).
    PubMed  Article  Google Scholar 

    29.
    Novarini, N. & Stival, E. Wading birds predation on Bufotes viridis (Laurenti, 1768) in the Ca’ Vallesina wetland (Ca’ Noghera, Venice, Italy). Boll. Mus. Storia Nat. Venezia 67, 71–75 (2017).
    Google Scholar 

    30.
    Clergeau, P., Reeber, S., Bastian, S. & Yésou, P. L. profil alimentaire de l’Ibis sacré Threskiornis aethiopicus introduit en France métropolitaine: espèce généraliste ou spécialiste. Rev. Ecol. Terre Vie 65, 331–342 (2010).
    Google Scholar 

    31.
    Fick, S. E. & Hijmans, R. J. WorldClim 2: new 1-km spatial resolution climate surfaces for global land areas. Int. J. Climatol. 37, 4302–4315 (2017).
    Article  Google Scholar 

    32.
    Pavia, M. L. nuova Banca Dati del Gruppo Piemontese Studi Ornitologici. Tichodroma 6, 152 (2017).
    Google Scholar 

    33.
    Saporetti, F. Abundance, phenology and geographical distribution in relation to habitat of Tringa species in N Italy: a summary of data from the Italian online portal www.ornitho.it. Wader Study 122, 60–70 (2015).
    Article  Google Scholar 

    34.
    iNaturalist. iNaturalist (2020) (accessed 29 April 2020). https://www.inaturalist.org.

    35.
    Bibby, C., Burgess, N., Hill, D. & Mustoe, S. Bird Census Techniques 2nd edn. (Academic Press, Cambridge, 2000).
    Google Scholar 

    36.
    Buckland, S. T. et al. Introduction to Distance Sampling: Estimating Abundance of Biological Populations (Oxford University Press, Oxford, 2001).
    Google Scholar 

    37.
    Thomas, L. et al. Distance software: design and analysis of distance sampling surveys for estimating population size. J. Appl. Ecol. 47, 5–14 (2010).
    PubMed  Article  Google Scholar 

    38.
    Fasola, M. & Ruiz, X. The value of rice fields as substitutes for natural wetlands for Waterbirds in the Mediterranean region. Colon. Waterbirds 19, 122–128 (1996).
    Article  Google Scholar 

    39.
    Fasola, M., Rubolini, D., Merli, E., Boncompagni, E. & Bressan, U. Long-term trends of heron and egret populations in Italy, and the effects of climate, human-induced mortality, and habitat on population dynamics. Popul. Ecol. 52, 59–72 (2010).
    Article  Google Scholar 

    40.
    Fasola, M. & Cardarelli, E. Long-term changes in the food resources of a guild of breeding Ardeinae (Aves) in Italy. Ital. J. Zool. 82, 238–250 (2014).
    Google Scholar 

    41.
    Fasola, M., Merli, E., Boncompagni, E. & Rampa, A. Monitoring heron populations in Italy, 1972–2010. J. Heron Biol. Conserv. 1, 1–10 (2011).
    Google Scholar 

    42.
    Ter Braak, C., Van Strien, A., Meijer, R. & Verstrael, T. Analysis of monitorng data with many missing values: which method? In Proc. 12th Int. Conf. IBCC EOAC Noordwijkerhout Neth. 663–673 (1994).

    43.
    Gregory, R. D. et al. Developing indicators for European birds. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 360, 269–288 (2005).
    PubMed  PubMed Central  Article  Google Scholar 

    44.
    Pannekoek, J. & van Strien, A. TRIM 3 Manual. TRends and Indices for Monitoring data. Stat. Neth. Voorbg. Neth. Research Paper No. 0102, 1–57 (2001).

    45.
    Sokal, R. R. & Rohlf, F. J. Biometry: The Principles and Practice of Statistics in Biological Research (W.H. Freeman, New York, 1995).
    Google Scholar 

    46.
    La Sorte, F. A. et al. Opportunities and challenges for big data ornithology. The Condor 120, 414–426 (2018).
    Article  Google Scholar 

    47.
    Camerano, P., Giannetti, F., Terzuolo, P. G. & Guiot, E. La Carta Forestale del Piemonte—Aggiornamento 2016 (IPLA S.p.A.—Regione Piemonte, Turin, 2017).
    Google Scholar 

    48.
    Guisan, A., Thuille, W. & Zimmerman, N. Habitat Suitability and Distribution Models: With Applications in R (Cambridge University Press, Cambridge, 2017).
    Google Scholar 

    49.
    Naimi, B. & Araújo, M. B. sdm: a reproducible and extensible R platform for species distribution modelling. Ecography 39, 368–375 (2016).
    Article  Google Scholar 

    50.
    Naimi, B., Hamm, N. A. S., Groen, T. A., Skidmore, A. K. & Toxopeus, A. G. Where is positional uncertainty a problem for species distribution modelling?. Ecography 37, 191–203 (2014).
    Article  Google Scholar 

    51.
    Chatterjee, S. & Hadi, A. S. Regression Analysis by Example (Wiley, Hoboken, 2006).
    Google Scholar 

    52.
    Elith, J. et al. A statistical explanation of MaxEnt for ecologists: statistical explanation of MaxEnt. Divers. Distrib. 17, 43–57 (2011).
    Article  Google Scholar 

    53.
    Phillips, S. J., Anderson, R. P., Dudík, M., Schapire, R. E. & Blair, M. E. Opening the black box: an open-source release of Maxent. Ecography 40, 887–893 (2017).
    Article  Google Scholar 

    54.
    Treggiari, A. A., Gagliardone, M., Pellegrino, I. & Cucco, M. Habitat selection in a changing environment: the relationship between habitat alteration and Scops Owl (Aves: Strigidae) territory occupancy. Ital. J. Zool. 80, 574–585 (2013).
    Article  Google Scholar 

    55.
    Allouche, O., Tsoar, A. & Kadmon, R. Assessing the accuracy of species distribution models: prevalence, kappa and the true skill statistic (TSS): assessing the accuracy of distribution models. J. Appl. Ecol. 43, 1223–1232 (2006).
    Article  Google Scholar 

    56.
    Heckmann, M. & Burk, L. Gridsampler—a simulation tool to determine the required sample size for repertory grid studies. J. Open Res. Softw. 5, 2 (2017).
    Article  Google Scholar 

    57.
    Guisan, A., Graham, C. H., Elith, J. & Huettmann, F. Sensitivity of predictive species distribution models to change in grain size. Divers. Distrib. 13, 332–340 (2007).
    Article  Google Scholar 

    58.
    Marion, L. & Marion, P. Première installation spontanée d’une colonie d’Ibis Sacré Threskiornis aethiopicus, au Lac de Grand-Lieu. Données préliminaires sur la production en jeunes et sur le régime alimentaire. Alauda 62, 275–280 (1994).
    Google Scholar 

    59.
    ONCFS. L’Ibis sacré (2020). http://www.oncfs.gouv.fr/La-lutte-contre-les-especes-exotiques-envahissantes-ru152/LIbis-sacre-ar282. Accessed 31 July 2020.

    60.
    Yésou, P., Clergeau, P., Bastian, S., Reeber, S. & Maillard, J.-F. The Sacred Ibis in Europe: ecology and management. Br. Birds 110, 197–212 (2017).
    Google Scholar 

    61.
    Herring, G. & Gawlik, D. E. Potential for successful population establishment of the nonindigenous sacred ibis in the Florida Everglades. Biol. Invas. 10, 969–976 (2008).
    Article  Google Scholar 

    62.
    Chen, K. & Hetherington, W. Sacred ibis nest control ineffective: environmentalist. Taipei Times 4 (2018).

    63.
    Bird Ecology Study Group. African Sacred Ibis in Taiwan (2020). https://besgroup.org/2019/01/28/african-sacred-ibis-in-taiwan/. Accessed 31 July 2020.

    64.
    Grussu, M. Checklist of the birds of Sardinia. Aves Ichnusae 4, 3–55 (2001).
    Google Scholar 

    65.
    Chane, M. & Balakrishnan, M. Population structure, feeding habits and activity patterns of the African Sacred ibis (Threskiornis aethiopicus) in Dilla Kera area, southern Ethiopia. Ethiop. J. Biol. Sci. 15, 93–105 (2016).
    Google Scholar 

    66.
    Smeraldo, S. et al. Modelling risks posed by wind turbines and power lines to soaring birds: the black stork (Ciconia nigra) in Italy as a case study. Biodivers. Conserv. 29, 1959–1976 (2020).
    Article  Google Scholar 

    67.
    Grussu, M. Gli uccelli alloctoni in Sardegna: una checklist aggiornata. Mem. Della Soc. Ital. Sci. Nat. E Mus. Civ. Storia Nat. Milano 36, 65 (2008).
    Google Scholar 

    68.
    Ranghetti, L. et al. Testing estimation of water surface in Italian rice district from MODIS satellite data. Int. J. Appl. Earth Observ. Geoinform. 52, 284–295 (2016).
    ADS  Article  Google Scholar 

    69.
    Fasola, M., Cardarelli, E., Pellitteri-Rosa, D. & Ranghetti, L. The recent decline of heron populations in Italy and the changes in rice cultivation practice. In 25th International Ornithological Congress. Ornithological Science (The Ornithological Society of Japan, 2014).

    70.
    Delmastro, G. B. Il gambero della Louisiana Procambarus clarkii (Girard, 1852) in Piemonte: nuove osservazioni su distribuzione, biologia, impatto e utilizzo (Crustacea: Decapoda: Cambaridae). Riv. Piemontese Storia Nat. 38, 61–129 (2017).
    Google Scholar 

    71.
    Bernini, G. et al. Complexity of biogeographic pattern in the endangered crayfish Austropotamobius italicus in northern Italy: molecular insights of conservation concern. Conserv. Genet. 17, 141–154 (2016).
    Article  Google Scholar 

    72.
    Fàbregas, M. C., Guillén-Salazar, F. & Garcés-Narro, C. The risk of zoological parks as potential pathways for the introduction of non-indigenous species. Biol. Invas. 12, 3627–3636 (2010).
    Article  Google Scholar 

    73.
    Brichetti, P. & Fracasso, G. Ornitologia Italiana. 1 Gaviidae-Falconidae (Alberto Perdisa editore, Bologna, 2003).
    Google Scholar 

    74.
    Wasef, S. Mitogenomic diversity in Sacred Ibis Mummies sheds light on early Egyptian practices. PLoS ONE 14, e0223964 (2019).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    75.
    Fasola, M. & Alieri, R. Conservation of heronry Ardeidae sites in North Italian agricultural landscapes. Biol. Conserv. 62, 219–228 (1992).
    Article  Google Scholar 

    76.
    Cocchi, R. & Volponi, S. Information on measures and related costs in relation to species included on the Union list—Threskiornis aethiopicus. Tech. Note Prep. IUCN Eur. Comm. ISPRA Ozzano Emilia Italy (2020).

    77.
    Smits, R., Van Horssen, P. & Van Der Winden, J. A Risk Analysis of the Sacred Ibis in the Netherlands: Including Biology and Management Options of this Invasive Species (Bureau Waardenburg Consultants for Environment & Ecology, Culemborg, 2010).
    Google Scholar 

    78.
    Byers, J. E. et al. Directing research to reduce the impacts of nonindigenous species. Conserv. Biol. 16, 630–640 (2002).
    Article  Google Scholar 

    79.
    Heath, J. A. & Frederick, P. C. Trapping White Ibises with rocket nets and mist nets in the Florida Everglades. J. Field Ornithol. 74, 187–192 (2003).
    Article  Google Scholar 

    80.
    Gurevitch, J. & Padilla, D. Are invasive species a major cause of extinctions?. Trends Ecol. Evol. 19, 470–474 (2004).
    PubMed  Article  Google Scholar 

    81.
    Didham, R. K., Tylianakis, J. M., Hutchison, M. A., Ewers, R. M. & Gemmell, N. J. Are invasive species the drivers of ecological change?. Trends Ecol. Evol. 20, 470–474 (2005).
    PubMed  Article  Google Scholar 

    82.
    Vaslin, M. Predaction de l’Ibis sacré sur des colonies des sternes et des guifettes. Ornithos 12, 106–109 (2005).
    Google Scholar 

    83.
    Evans, T. & Blackburn, T. M. Global variation in the availability of data on the environmental impacts of alien birds. Biol. Invas. 22, 1027–1036 (2020).
    Article  Google Scholar 

    84.
    Clergeau, P., Fourcy, D., Reeber, S. & Yésou, P. New but nice? Do alien sacred ibises Threskiornis aethiopicus stabilize nesting colonies of native spoonbills Platalea leucorodia at Grand-Lieu Lake, France?. Oryx 44, 533–538 (2010).
    Article  Google Scholar 

    85.
    Bremner, A. & Park, K. Public attitudes to the management of invasive non-native species in Scotland. Biol. Conserv. 139, 306–314 (2007).
    Article  Google Scholar  More