More stories

  • in

    Ecology-guided prediction of cross-feeding interactions in the human gut microbiome

    Overview of the GutCP algorithm
    Our approach uses the idea that we can leverage cross-feeding interactions—which comprise knowing the metabolites that each microbial species is capable of consuming and producing—to mechanistically connect the levels of microbes and metabolites in the human gut. Several different mechanistic models in past studies have shown that this is indeed possible18,20,29,36,37. While GutCP is generalizable and can be used with any of these models, in this paper, we use a previously published consumer-resource model20. We use this model because of its context and performance: it is built specifically for the human gut and is best able to explain the experimentally measured species composition of the gut microbiome with its resulting metabolic environment, or fecal metabolome (compared with other state-of-the-art methods, such as ref. 29). To predict the metabolome from the microbiome, it relies on a manually curated set of known cross-feeding interactions9. It then uses these known interactions to follow the stepwise flow of metabolites through the gut. At each step (ecologically, at each trophic level), the metabolites available to the gut are utilized by microbial species that are capable of consuming them, and a fraction of these metabolites are secreted as metabolic byproducts. These byproducts are then available for consumption by another set of species in the next trophic level. After several such steps, the metabolites that are left unconsumed constitute the fecal metabolome.
    We hypothesized that adding new, yet-undiscovered cross-feeding interactions would improve our ability to predict the levels of metabolites with our mechanistic and causal model. Specifically, we predict that the set of undiscovered interactions resulting in the most accurate and optimal improvement in predictions would be the most likely candidates for true cross-feeding interactions. Inferring such an optimal set of new cross-feeding interactions or reactions is the main logic driving GutCP. In what follows, we sometimes refer to cross-feeding reactions (i.e., metabolite consumption or production by microbes) as “links” in an overall cross-feeding network of the gut microbiome, whose nodes are microbes and metabolites (Fig. 1a; metabolites in blue, microbes in orange); the links themselves are directed edges connecting the nodes. Links can be of two types: consumption or nutrient uptake reactions (from nutrients to microbes) and production or nutrient secretion reactions (from microbes to their metabolic byproducts).
    Fig. 1: Overview of the GutCP algorithm.

    a Schematic of the original set of known cross-feeding interactions (top) and bar plot of the prediction error for each metabolite and microbe (bottom). The cross-feeding interactions are represented as a network, whose nodes are either metabolites (cyan circles) or microbial species (orange ellipses), and directed links represent the abilities of different species to consume (red arrows) and produce (blue arrows) individual metabolites. b GutCP adds a new consumption link (red) and production link (blue) as added links reduce the prediction errors for metabolites and microbes.

    Full size image

    The salient aspects of our method are outlined in Fig. 1. We start with the known set of consumption and production links that were originally used by the model; these links are known from direct experiments and represent a ground-truth dataset or original cross-feeding network9. These are shown in Fig. 1a through the pink and blue arrows connecting nutrients 1 through 6 with microbes (a) through (c). For each sample, using only the species abundance from the microbiome, we use the model to quantitatively estimate the microbiome’s species and metabolomic composition. Briefly, we assume that a defined set of polysaccharides, common to human diets, are available as the nutrient intake to the gut (nutrients 1 and 4 in Fig. 1a). We calculate the microbiome and metabolome profiles separately for each individual, which contain a different set of microbial species in their guts. At the first trophic level, all microbial species that are capable of using the polysaccharides (indicated by the pink arrows in Fig. 1a) consume each of them in proportion to their abundances (microbes a, b, and c in Fig. 1a). They subsequently secrete a fixed fraction of the consumed nutrients as metabolic byproducts; every species at this trophic level secretes all the metabolic byproducts it is known to secrete (blue arrows in Fig. 1a) in equal proportion (nutrients 2–6 in Fig. 1a). At the next trophic level, all species detected in the individual’s gut which can consume the newly secreted byproducts consume them as nutrients, secreting a new set of byproducts, and this continues for four trophic levels (not shown in Fig. 1a for simplicity). At the end of this process, all metabolites which remain unconsumed by the community comprise the metabolome of the individual and the microbial species which consume nutrients and grow comprise the microbiome of the individual (for a complete description, see “Methods” and previous work20).
    For each metabolite and microbial species, there can be two kinds of prediction errors, or biases: individual (the sample-specific difference between predicted and measured levels) and systematic (average difference across all samples). We focused on the “systematic bias” for each metabolite and microbial species: the average deviation of the predicted levels from the measured levels across all samples in our dataset (Fig. 1a, bottom). The systematic bias for each metabolite and microbe tells us whether our model generally tends to predict their level to be greater than observed (overpredicted), less than observed (underpredicted), or neither (well-predicted). We assume that metabolites and microbes with a large systematic bias are most likely to harbor missing consumption or production links that are relevant across many samples. We prioritize adding links to them in proportion to their systematic biases.
    After measuring the systematic bias for each metabolite and microbe, GutCP proceeds in discrete steps (Fig. 1a, b). At each step, we attempt to add a new link to the current cross-feeding network. This new link is chosen randomly from the entire set of combinatorially possible links (see “Methods”; for S species, M metabolites, and two kinds of links (consumption and production), there are a total of 2SM combinatorially possible links). We accept this link—keeping it in the current network—if it leads to an overall improvement in the agreement between the predicted and measured levels of microbes and metabolites. We repeat the process of adding new links—accepting or rejecting them—until the improvements in the levels of metabolites and microbes became insignificant. Overall, GutCP can add several links to improve the agreement between the predicted and measured levels of microbes and metabolites (in Fig. 1a, b, bottom, adding the extra red and blue link at the top results in improved predictions for metabolite (1), metabolite (3), and microbe (b). Figure 2a shows how the cross-feeding network improves over a typical GutCP run via the red trajectory, starting from the original network (Fig. 2a, top left) to the final network state (Fig. 2a, bottom right). Trajectories from 100 other runs are shown in gray. GutCP repeatably reduces both the error of the metabolome predictions (y axis; measured as ({text{log}}_{10}(frac{,text{pred}-text{meas}}{text{measurement},}))) and improves the correlation between the predicted and measured metabolomes (x axis).
    Fig. 2: Improvement in predictions using GutCP.

    a Improvement in log error (({text{log}}_{10}(frac{,text{pred}-text{meas}}{text{measurement},}))) and the correlation between the prediction and measured fecal metabolome during 100 typical runs of the GutCP algorithm. The gray point at the top left indicates the performance of the original cross-feeding network of Ref. 9, and the black points at the bottom right, that of improved networks predicted using GutCP. A trajectory example, highlighting how performance improves over a GutCP run, is shown in red, and others are shown in gray. b Rarefaction curve showing the number of unique cross-feeding interactions discovered by GutCP over 100 runs of the algorithm. c Prevalence of links, i.e., the number of GutCP runs in which they repeatedly appeared (red dots; total 100 runs) and for comparison, a corresponding binomial distribution with the same mean (black dotted line). P values for different prevalences are estimated using the one-sided binomial test.

    Full size image

    Cross-validating the newly predicted interactions
    To test if the cross-feeding interactions predicted by GutCP are generalizable to unknown datasets, we performed fourfold cross-validation. We used a sample -omics dataset of the gut microbiome and metabolome sampled from 41 human individuals, comprising 221 metabolites and 72 microbial species (data from ref. 38). We split our -omics dataset into two subsets: training (three-fourths of the individuals) and test (one-fourth of the individuals) subsets. We then ran GutCP on the training subset to discover new interactions and added them to the ground-truth interactions taken from ref. 9. Doing so resulted in a network of cross-feeding interactions learned only from the training subset of the data. Finally, we evaluated the improvement in accuracy of metabolome predictions resulting from the trained network on the unseen, test subset of the data. We repeated this process three times, each time splitting the full dataset into a training subset (with a randomly chosen three-fourths of the individuals) and test subset (with the remaining one-fourth of the individuals); finally, we calculated the average improvement in prediction accuracy over all four splits.
    We found that both the training and test set performances after using the links predicted by GutCP were significantly better than the baseline given by the original cross-feeding network (Table 1). Specifically, both measures of model performance, namely the logarithmic error and the average correlation, improved by 64% and 20%, respectively, after adding GutCP’s discovered interactions. In addition, the test set performance was comparable to the training set performance (6% difference; Table 1). This suggests that the cross-feeding interactions inferred by GutCP are not likely to be a result of over-fitting.
    Table 1 Cross-validating the newly predicted interactions.
    Full size table

    Building a consensus-based atlas of predicted cross-feeding interactions
    Having confirmed that GutCP is unlikely to over-fit data, we pooled the entire sample dataset of 41 individuals and ran 100 independent instances of our prediction algorithm on it; we verified that incorporating more instances did not qualitatively affect our results (Fig. 2b shows a rarefaction curve, which highlights the number of new links discovered by GutCP as we perform more runs the algorithm). Each run of the algorithm resulted in an average of 140 newly predicted cross-feeding interactions. Then, based on consensus from many runs, we assigned a confidence level to each predicted interaction, namely what fraction of GutCP runs it was discovered in. By calculating a null distribution (Fig. 2c, black), which predicts the fraction of GutCP runs where a random link would be discovered by chance, we assigned a P value to each link and set a threshold at P = 10−3 (Fig. 3c, red; see “Methods” for details). Doing so finally resulted in a complete consensus-based atlas of 293 predicted cross-feeding interactions, which we have provided as a resource for experimental verification in Supplementary Table 1. Figure 3a shows a condensed version of these interactions obtained from the simulation with the best performance (the trajectory example in Fig. 2a with the lowest log error and highest correlation coefficient) in the form of a matrix; specifically, newly added interactions are in dark colors, and old interactions in faded colors. Supplementary Fig. 3 shows a complete version of this matrix. Note that some of the predicted interactions in Fig. 3a are unrealistic, e.g., the production of certain sugars like D-Fructose and D-Sorbitol. Such interactions are unlikely to be predicted in repeated simulations, and thus will not be part of the final consensus set. This illustrates the power of pooling results from several simulations to arrive at a set of highly probable predictions.
    Fig. 3: New cross-feeding interactions predicted by GutCP.

    a Concise matrix representation of the improved cross-feeding network of the gut microbiome predicted by GutCP (the trajectory example in Fig. 2a with the best performance). The rows are metabolites, and columns, microbial species. Faded cells represent the original, known set of cross-feeding interactions, both production (light blue), consumption (light red), and bidirectional links (gray). The new cross-feeding interactions predicted by GutCP are shown in dark colors: production links in dark blue, consumption links in dark red, and bidirectional links in black. b Network of 293 new links predicted by GutCP (with a P value  More

  • in

    Disentangling the role of environment in cross-taxon congruence of species richness along elevational gradients

    1.
    Brown, J. H. Why are there so many species in the tropics? J. Biogeogr. 41, 8–22 (2014).
    PubMed  Article  PubMed Central  Google Scholar 
    2.
    Classen, A. et al. Temperature versus resource constraints: Which factors determine bee diversity on Mount Kilimanjaro, Tanzania? Glob. Ecol. Biogeogr. 24, 642–652 (2015).
    Article  Google Scholar 

    3.
    Rahbek, C. et al. Humboldt’s enigma: What causes global patterns of mountain biodiversity? Science 365, 1108–1113 (2019).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    4.
    Toranza, C. & Arim, M. Cross-taxon congruence and environmental conditions. BMC Ecol. 10, 18 (2010).
    PubMed  PubMed Central  Article  Google Scholar 

    5.
    Gioria, M., Bacaro, G. & Feehan, J. Evaluating and interpreting cross-taxon congruence: Potential pitfalls and solutions. Acta Oecol. 37, 187–194 (2011).
    ADS  Article  Google Scholar 

    6.
    Graham, C. H. et al. The origin and maintenance of montane diversity: Integrating evolutionary and ecological processes. Ecography (Cop.) 37, 711–719 (2014).
    Article  Google Scholar 

    7.
    Westgate, M. J., Tulloch, A. I. T., Barton, P. S., Pierson, J. C. & Lindenmayer, D. B. Optimal taxonomic groups for biodiversity assessment: A meta-analytic approach. Ecography (Cop.) 40, 539–548 (2017).
    Article  Google Scholar 

    8.
    Lomolino, M. V. Elevation gradients of species-density: Historical and prospective views. Glob. Ecol. Biogeogr. 8, 1–2 (2001).
    Google Scholar 

    9.
    McCain, C. M. Global analysis of bird elevational diversity. Glob. Ecol. Biogeogr. 18, 346–360 (2009).
    Article  Google Scholar 

    10.
    Peters, M. K. et al. Predictors of elevational biodiversity gradients change from single taxa to the multi-taxa community level. Nat. Commun. 7, 13736 (2016).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    11.
    Sundqvist, M. K., Sanders, N. J. & Wardle, D. A. Community and ecosystem responses to elevational gradients: Processes, mechanisms, and insights for global change. Annu. Rev. Ecol. Evol. Syst. 44, 261–280 (2013).
    Article  Google Scholar 

    12.
    Ruggiero, A. & Hawkins, B. A. Why do mountains support so many species of birds? Ecography (Cop.) 31, 306–315 (2008).
    Article  Google Scholar 

    13.
    Mccain, C. M. & Colwell, R. K. Assessing the threat to montane biodiversity from discordant shifts in temperature and precipitation in a changing climate. Ecol. Lett. 14, 1236–1245 (2011).
    PubMed  Article  PubMed Central  Google Scholar 

    14.
    Hawkins, B. A. et al. Energy, water, and broad-scale geographic patterns of species richness. Ecology 84, 3105–3117 (2003).
    Article  Google Scholar 

    15.
    Currie, D. J. Energy and large-scale patterns of animal and plant species richness. Am. Nat. 137, 27–49 (1991).
    Article  Google Scholar 

    16.
    Stein, A., Gerstner, K. & Kreft, H. Environmental heterogeneity as a universal driver of species richness across taxa, biomes and spatial scales. Ecol. Lett. 17, 866–880 (2014).
    PubMed  Article  PubMed Central  Google Scholar 

    17.
    Costanza, J. K., Moody, A. & Peet, R. K. Multi-scale environmental heterogeneity as a predictor of plant species richness. Landsc. Ecol. 26, 851–864 (2011).
    Article  Google Scholar 

    18.
    Vetaas, O. R., Paudel, K. P. & Christensen, M. Principal factors controlling biodiversity along an elevation gradient: Water, energy and their interaction. J. Biogeogr. https://doi.org/10.1111/jbi.13564 (2019).
    Article  Google Scholar 

    19.
    Lande, R. Risks of population extinction from demographic and environmental stochasticity and random catastrophes. Am. Nat. 142, 911–927 (1993).
    PubMed  Article  PubMed Central  Google Scholar 

    20.
    Kaspari, M., Alonso, L. & O’Donnell, S. Three energy variables predict ant abundance at a geographical scale. Proc. R. Soc. B Biol. Sci. 267, 485–489 (2000).
    CAS  Article  Google Scholar 

    21.
    Pianka, E. R. Latitudinal gradients in species diversity: A review of concepts. Am. Nat. 100, 33–46 (1966).
    Article  Google Scholar 

    22.
    Werenkraut, V. & Ruggiero, A. The richness and abundance of epigaeic mountain beetles in north-western Patagonia, Argentina: Assessment of patterns and environmental correlates. J. Biogeogr. 41, 561–573 (2014).
    Article  Google Scholar 

    23.
    R Core Team. R version 3.6.2 ‘Dark and Stormy Night’ (2019). (Accessed 12 December 2019). https://www.r-project.org. 

    24.
    Blanchet, F. G., Cazelles, K. & Gravel, D. Co-occurrence is not evidence of ecological interactions. Ecol. Lett. 23, 1050–1063 (2020).
    PubMed  Article  Google Scholar 

    25.
    Hodkinson, I. D. Terrestrial insects along elevation gradients: Species and community responses to altitude. Biol. Rev. 80, 489–513 (2005).
    PubMed  Article  Google Scholar 

    26.
    Kampmann, D. et al. Mountain grassland biodiversity: Impact of site conditions versus management type. J. Nat. Conserv. 16, 12–25 (2008).
    Article  Google Scholar 

    27.
    Janzen, D. H. et al. Changes in the arthropod community along an elevational transect in the Venezuelan Andes. Biotropica 8, 193–203 (1976).
    Article  Google Scholar 

    28.
    Sirin, D., Eren, O. & Ciplak, B. Grasshopper diversity and abundance in relation to elevation and vegetation from a snapshot in Mediterranean Anatolia: Role of latitudinal position in altitudinal differences. J. Nat. Hist. 44, 1343–1363 (2010).
    Article  Google Scholar 

    29.
    Alexander, G. & Hilliard, J. R. Altitudinal and seasonal distribution of Orthoptera in the Rocky Mountains of northern Colorado. Ecol. Monogr. 39, 385–432 (1969).
    Article  Google Scholar 

    30.
    Mojica, A. S. & Fagua, G. Estructura de las comunidades de orthoptera (insecta) en un gradiente altitudinal de un bosque andino. Rev. Colomb. Entomol. 32, 200–213 (2006).
    Google Scholar 

    31.
    Grytnes, J. A. Species-richness patterns of vascular plants along seven altitudinal transects in Norway. Ecography (Cop.) 26, 291–300 (2003).
    Article  Google Scholar 

    32.
    McCain, C. M. & Grytnes, J.-A. Elevational gradients in species richness. Encyl. Life Sci. https://doi.org/10.1002/9780470015902.a0022548 (2010).
    Article  Google Scholar 

    33.
    Xu, X. et al. Altitudinal patterns of plant species richness in the Honghe region of China. Pak. J. Bot. 49, 1039–1048 (2017).
    Google Scholar 

    34.
    Kerr, J. T. & Packer, L. Habitat heterogeneity as a determinant of mammal species richness in high-energy regions. Nature 385, 252–254 (1997).
    ADS  CAS  Article  Google Scholar 

    35.
    Röder, J. et al. Heterogeneous patterns of abundance of epigeic arthropod taxa along a major elevation gradient. Biotropica 49, 217–228 (2017).
    Article  Google Scholar 

    36.
    Evans, K. L., Warren, P. H. & Gaston, K. J. Species-energy relationships at the macroecological scale: A review of the mechanisms. Biol. Rev. Camb. Philos. Soc. 80, 1–25 (2005).
    PubMed  Article  Google Scholar 

    37.
    Kissling, W. D., Rahbek, C. & Böhning-Gaese, K. Food plant diversity as broad-scale determinant of avian frugivore richness. Proc. R. Soc. Biol. Sci. 274, 799–808 (2007).
    Article  Google Scholar 

    38.
    Kissling, W. D., Field, R. & Böhning-Gaese, K. Spatial patterns of woody plant and bird diversity: Functional relationships or environmental effects? Glob. Ecol. Biogeogr. 17, 327–339 (2008).
    Article  Google Scholar 

    39.
    Chown, S. L. & Gaston, K. J. Exploring links between physiology and ecology at macro scales: The role of respiratory metabolism in insects. Biol. Rev. 74, 87–120 (1999).
    Article  Google Scholar 

    40.
    de Araújo, W. S. Different relationships between galling and non-galling herbivore richness and plant species richness: A meta-analysis. Arthropod. Plant. Interact. 7, 373–377 (2013).
    Article  Google Scholar 

    41.
    Qian, H. & Kissling, W. D. Spatial scale and cross-taxon congruence of terrestrial vertebrate and vascular plant species richness in China. Ecology 91, 1172–1183 (2010).
    PubMed  Article  PubMed Central  Google Scholar 

    42.
    Burrascano, S. et al. Congruence across taxa and spatial scales: Are we asking too much of species data? Glob. Ecol. Biogeogr. 27, 980–990 (2018).
    Article  Google Scholar 

    43.
    Field, R. et al. Spatial species-richness gradients across scales: A meta-analysis. J. Biogeogr. 36, 132–147 (2009).
    Article  Google Scholar 

    44.
    Giorgis, M. A. et al. Composición florística del Bosque Chaqueño Serrano de la provincia de Córdoba, Argentina. Kurtziana 36, 9–43 (2011).
    Google Scholar 

    45.
    Cabido, M., Funes, G., Pucheta, E., Vendramani, F. & Díaz, S. A chorological analysis of the mountains from Central Argentina. Is all what we call Sierra Chaco really Chaco? Contribution to the study of the flora and vegetation of the Chaco: 12. Candollea 53, 321–331 (1998).
    Google Scholar 

    46.
    Giorgis, M. A. et al. Changes in floristic composition and physiognomy are decoupled along elevation gradients in central Argentina. Appl. Veg. Sci. 20, 553–571 (2017).
    Article  Google Scholar 

    47.
    Cabrera, A. L. Fitogeografia de la República Argentina. In Enciclopedia Argentina de Agricultura y Jardinería Vol. 14 (ed. Kugler, W. F.) 1–42 (ACME, New York, 1976).
    Google Scholar 

    48.
    Giorgis, M. A. et al. Diferencias en la estructura de la vegetación del sotobosque entre una plantación de Pinus taedaL. (Pinaceae) y un matorral serrano (Cuesta Blanca, Córdoba). Kurtziana 31, 39–49 (2005).
    Google Scholar 

    49.
    Martínez, G. A., Arana, M. D., Oggero, A. J. & Natale, E. S. Biogeographical relationships and new regionalisation of high-altitude grasslands and woodlands of the central Pampean Ranges (Argentina), based on vascular plants and vertebrates. Aust. Syst. Bot. 29, 473–488 (2016).
    Article  Google Scholar 

    50.
    QGIS Development Team. QGIS Geographic Information System (Accessed 19 April 2019). (2019).

    51.
    Kent, M. The description of vegetation in the field. In Vegetation Description and Data Analysis: A Practical Approach (ed. Kent, M.) 65–116 (Wiley-Blackwell, Hoboken, 2012).
    Google Scholar 

    52.
    Catálogo de las plantas vasculares del Cono Sur : (Argentina, Sur de Brasil, Chile, Paraguay y Uruguay). (Missouri Botanical Garden Press, 2008).

    53.
    Haddad, N., Tilman, D., Haarstad, J., Ritchie, M. & Knops, J. M. N. Contrasting effects of plant richness and composition on insect communities: a field experiment. Am. Nat. 158, 17–35 (2001).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    54.
    Braun, H. & Zubarán, G. Tettigoniidae (Orthoptera) Species from Argentina and Uruguay (2019).

    55.
    Carbonell, C. S., Cigliano, M. M. & Lange, C. E. Acridomorph (Orthoptera) Species of Argentina and Uruguay. Version II [2019]. https://biodar.unlp.edu.ar/acridomorph/.

    56.
    Cigliano, M. M., Braun, H., Eades, D. C. & Otte, D. Orthoptera Species File. Version 5.0/5.0 (2018). http://orthoptera.speciesfile.org.

    57.
    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 

    58.
    Hijmans, R. J., Cameron, S. E., Parra, J. L., Jones, P. G. & Jarvis, A. Very high resolution interpolated climate surfaces for global land areas. Int. J. Climatol. 25, 1965–1978 (2005).
    Article  Google Scholar 

    59.
    Carrara, R., Silvestro, V. A., Cheli, G. H., Campón, F. F. & Flores, G. E. Disentangling the effect of climate and human influence on distribution patterns of the darkling beetle Scotobius pilularius Germar, 1823 (Coleoptera: Tenebrionidae). Ann. Zool. 66, 693–701 (2016).
    Article  Google Scholar 

    60.
    Aisen, S., Werenkraut, V., Márquez, M. E. G., Ramírez, M. J. & Ruggiero, A. Environmental heterogeneity, not distance, structures montane epigaeic spider assemblages in north-western Patagonia (Argentina). J. Insect Conserv. 21, 1–12 (2017).
    Article  Google Scholar 

    61.
    Bilskie, J. Soil Water Status: Content and Potential (Campbell Scientific Inc., Logan, 2001).
    Google Scholar 

    62.
    Tucker, C. Red and photographic infrared linear combinations for monitoring vegetation. Remote Sens. Environ. 8, 127–150 (1979).
    ADS  Article  Google Scholar 

    63.
    Wang, J., Rich, P. M., Price, K. P. & Dean-Kettle, W. Relations between NDVI, grassland production, and crop yield in the central great plains. Geocarto Int. 20, 5–11 (2005).
    Article  Google Scholar 

    64.
    Oindo, B. O., de By, R. A. & Skidmore, A. K. Interannual variability of NDVI and bird species diversity in Kenya. Int. J. Appl. Earth Obs. Geoinf. 2, 172–180 (2000).
    ADS  Article  Google Scholar 

    65.
    IGN. Modelo Digital de Elevaciones de la República Argentina. (Instituto Geográfico Nacional—Dirección General de Servicios Geográficos—Dirección de Geodesia, 2016).

    66.
    Riley, S. J., DeGloria, S. D. & Elliot, R. A terrain ruggedness index that quantifies topographic heterogeneity. Intermt. J. Sci. 5, 23–27 (1999).
    Google Scholar 

    67.
    Stein, A. & Kreft, H. Terminology and quantification of environmental heterogeneity in species-richness research. Biol. Rev. 90, 815–836 (2015).
    PubMed  Article  Google Scholar 

    68.
    Tilman, D. & Pacala, S. W. The maintenance of species richness in plant communities. In Species Diversity in Ecological Communities (eds Ricklefs, R. E. & Schulter, D.) 13–25 (University of Chicago Press, Chicago, 1993).
    Google Scholar 

    69.
    Cleveland, W. S., Grosse, E. & Shyu, W. M. Local regresion models. In Statistical Models in S (eds Chambers, J. M. & Hastie, T. J.) 227 (Chapman and Hall, London, 1993).
    Google Scholar 

    70.
    Szewczyk, T. & Mccain, C. M. A systematic review of global drivers of ant elevational diversity. PLoS ONE 11, e0155404 (2016).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    71.
    Beck, J. et al. Elevational species richness gradients in a hyperdiverse insect taxon: A global meta-study on geometrid moths. Glob. Ecol. Biogeogr. 26, 412–424 (2017).
    Article  Google Scholar 

    72.
    Bolker, B. M. Ecological Statistics: Contemporary Theory and Application (Oxford University Press Inc., Oxford, 2015).
    Google Scholar 

    73.
    Grace, J. B. Structural Equation Modeling and Natural Systems (Cambridge University Press, Cambridge, 2006).
    Google Scholar 

    74.
    Lefcheck, J. S. piecewiseSEM: Piecewise structural equation modelling in r for ecology, evolution, and systematics. Methods Ecol. Evol. 7, 573–579 (2016).
    Article  Google Scholar 

    75.
    Jiménez-Alfaro, B., Chytrý, M., Mucina, L., Grace, J. B. & Rejmánek, M. Disentangling vegetation diversity from climate-energy and habitat heterogeneity for explaining animal geographic patterns. Ecol. Evol. 6, 1515–1526 (2016).
    PubMed  PubMed Central  Article  Google Scholar  More

  • in

    The concerted emergence of well-known spatial and temporal ecological patterns in an evolutionary food web model in space

    We have introduced and investigated a spatially explicit evolutionary food web model that allows us to explore the distribution of species in space and time, as well as the waxing and waning of species ranges with time. This is the first model that makes it possible to explore these features in the context of trophic networks, showing how they are influenced by competition as well as by predators and prey. Indeed, our model produces empirically well-known patterns in space and time, such as lifetime distributions, species-area relationships, distance decay of similarity, and temporal change of geographic range. On top, we obtain a variety of additional result and gain insights into the mechanisms that generate these patterns. While one might argue that some patterns must emerge trivially in a model like ours, the multitude of patterns emerging together is remarkable.
    We find that most species only appear for short times and have small ranges. For species that conquer larger portions of the web we analysed the shape of the range evolution and found that basal species show the empirically observed “hat” pattern more often than species in higher trophic levels. This indicates that the trophic position of a species plays a major role for its range expansion success as well as the shape of its range expansion trajectory over time. To our knowledge, this has not been discussed in the literature so far.
    Recently Zliobaite et al.11 analyzed which factors are more correlated to the rise and fall of the range expansion trajectory in fossil data sets of basal mammals. The range expansion follows the so called “hat pattern” that consists of five phases in a species lifetime: origination, expansion, peak, decline and extinction. They found that the temporal location of the peak of the hat pattern is more impacted by competition while the brims are more influenced by abiotic environmental factors. They also compared the range curves of different random walk models with the shape of empirical data and found that a random walk model with competition and environmental factor provides the most realistic looking curves.
    Our results for basal species provide an illustration of the mechanism that might lead to these findings. There is one major difference in assumptions: the “environment” in our case is the trophic environment (network structure and abundance distribution). We do not model an abiotic surrounding, yet basal range curves look strikingly hat shaped. A species needs to fit into this trophic environment (network) to first establish a viable population (origination). To successfully increase its range it needs to disperse to neighbouring habitats and be a viable competitor there as well. This leads to the extinction of another species as the dispersing competitor takes its place. As neighbouring basal communities are similar in our systems, the chances are high that the species can spread on a large portion on the grid replacing other species (expansion). This continues until the species has reached the maximum range (peak). It is only a matter of time then until this process is repeated with the species having become the inferior competitor. The species is then successively replaced by a better adapted species (decline). The species thus ages as the network structure changes. At some point the species has vanished on all habitats (extinction). This means that we observe the same dynamics as suggested by Zliobaite et al., but with the trophic and not the abiotic environment as the main driver of the initial increase and later decrease of the range. The truth is probably that both the trophic and abiotic environment are important, as both play an important role in real ecosystems.
    Regarding higher trophic species in our system the range expansion curves look more diverse and often do not resemble the hat shape. These species depend on the composition on the layer below. As this layer changes the fate of the higher species changes as well. The emergence of a new prey species that spreads over the network can save a consumer species from extinction. This cannot happen for basal species as these are more or less completely controlled by competition. The studies that we know often deal with basal species, so we are not convinced that the hat pattern is ubiquitous for all species. Future empirical work could focus on predator range expansion and try to find a case where a predator species could regain its range after the emergence of a new prey.
    A qualitative difference between the basic trophic layer and higher layers occurred in our data also with respect to the similarity of networks in nearby habitats. The similarity index decays particularly slowly for the basal layer. Theory on distance decay suggests that spatial heterogeneity is a main driver in community turnover in two ways: (1) competitive species sorting along environmental gradients and (2) topological influences that let species with different dispersal abilities experience different landscapes6. As we use a homogeneous landscape we expect the first driver to be non-existent for the basal layer, as all habitats hold the same type and amount of resource. As species do not fundamentally differ in their dispersal abilities and we do not have a heterogeneous spatial topology, the second point is only weakly relevant for the basal species. They have slightly different chances of being chosen for dispersal as we choose the next disperser depending on the biomass density. As we have seen, biomass densities are quite similar for basal species. What remains is a temporal aspect: species that are older can reside on more habitats and have thus a higher chance of being chosen to disperse. To put the cart before the horse, this confirms the theory on distance decay: we expect a much faster decay in a heterogeneous environment, and this is exactly what we observe for higher trophic layers, which experience heterogeneity due to the spatial turnover in basal species composition. A trend to faster decay rates in higher trophic levels was also found in a meta-study7.
    A model is always a simplification of reality. Some of the assumptions underlying our model are worth discussing. Species in nature are not restricted to the one-dimensional niche space that we assume. In fact, the original Webworld model characterised species by a large vector of traits31. We, in contrast, characterized species by three traits, all of which are based on body mass. We think that this is the reason why we do not observe super abundant species, but species densities are all of a similar order of magnitude. With a higher-dimensional trait space there must exist more diverse species types and probably also super abundant species, which have a globally optimal trait vector. Nevertheless, our simplification leads to an overall shape of the rank abundance curves that is realistic, as one would expect from a niche apportionment model40. Harpole and Tilman showed that diversity and evenness of grassland communities decreased when niche dimensionality was reduced by adding limiting nutrients to plot experiments41. In turn this indicates that rank abundance curves for less dimensional communities will be flatter. This is in line with the shape of our rank abundance curves.
    The probably most interesting trait to add to each species would be its dispersal rate, and to let the dispersal rate evolve. We would expect that this would lead to higher-level species dispersing faster than lower-level species, thus making the differences in range between the different trophic levels smaller. In fact, a previous predator-prey model showed that the predator’s dispersal ability evolved in accordance to spatio-temporal fluctuations of the prey; with higher dispersal rates evolving for larger fluctuations in prey42. However, in the context of food webs the evolution of dispersal is poorly understood and a model like ours would be a good starting point.
    Our choice of parameters is guided by the aim to make the model feasible. To be able to perform computer simulations on a large number of habitats we chose a relatively small value for the amount of resource R, so that the number of species of a local food web remained below 25. As we wanted to simulate food webs and not only basal communities we needed to choose a value of the efficiency (lambda ) that allows for the emergence of several trophic layers. In the original Webworld Model, (lambda ) was identified with the proportion of biomass that is passed from one trophic layer to the next. The value of 0.65 that we use here is much larger than the empirically established value of 0.143. However, in the original Webworld model no distinction was made between resident biomass and biomass fluxes. Therefore the variable B was identified with biomass, while its occurrence in Eq. (4) in fact suggests that it represents biomass flux. A further reason for the difference between the model value and the empirical value of (lambda ) is that the model does not take into account energy input due to the below-ground ecological processes. Due to all these simplifications of the model, we do not consider it important to provide an empirical justification of the precise value of the parameter (lambda ), but base its value on the condition that the model yields food webs with several trophic levels.
    In contrast to other models, the model used here does not rely on an extrinsic extinction rate that randomly extirpates species that might be well adapted to the network. All extinction events are driven by the trophic dynamics, yet we observe an ongoing species turn over. We thus study the pure food web dynamics without a heterogeneous or fluctuating environment and still observe ecological reasonable species distributions. This indicates that incorporating abiotic environments and their fluctuations is not necessarily needed to study food web dynamics.
    Rogge et al. analysed lifetime distributions and SAR curves in a model that is simpler than ours as it does not include population sizes22. The lifetime distributions that we find are considerably steeper (slope (-2.4)) compared to their value around (-1.7). It is also larger than the values reported for empirical findings, which lie between 1 and 2; Newman and Palmer pin them down to (1.7pm 0.3)10. Data for contemporary lifetime distributions show a power-law like shape23,44 with exponents that are in agreement with the exponent of paleological data10. It is noteworthy that there is no consensus whether lifetime distributions follow a power-law or an exponential law, as data often allow for both types of fit, due to (large) uncertainties in fossil data10,45. Exponentials, of course, have a changing slope in a double-logarithmic plot and can thus also be compatible with the exponent observed by us.
    Curiously, our model shows steep lifetime distributions even though there is no external random extinction implemented as in other models22. One implication of our value (alpha > 2) is that our distribution has a well-defined mean. This features is shared with exponential distributions.
    McPeek argues that lifetime distributions depend on the number and survival time of “transient” species, i.e. species that are on their way to extinction25. He reasons that the time to extinction is elongated for species that are similar, because the inferior competitor holds out longer when it competes with more similar species. If this applies to our type of model, this indicates that species in our system are, despite the one-dimensional niche axis, not as similar as species in the model of Rogge et al.22 that uses the same niche axis, as we observe shorter lifetimes. The difference is that interaction links in22 are binary (presence versus absence), whilst we use Gaussian feeding kernels. The fact that this difference affects the lifetime distributions emphasises the importance of considering details of the trophic interactions. The SAR curves on the contrary are flatter in our model than in the model by Rogge et al.  and in better agreement with empirical data.
    O’Sullivan et al.18 found in a competitive metacommunity assembly models a similar collection of macroecological patterns (SAD, range size distribution RSD and SAR) as we did, when regional diversity was near equilibrium. They refer to the work of McGill16 who analysed the assumptions underlying models of macroecological patterns and found that three key ingredients seem to be sufficient for such patterns to emerge. Those are a left skewed SAD, clumping of populations in space, and species distributions in space that are uncorrelated from other species spatial distributions. O’Sullivan et al.18 report that all three ingredients occur in their model and are shaped by regional diversity equilibrium. The closer the system to regional equilibrium the stronger are the observed key patterns (SAD, RSD, Spatial non-correlation). They relate their finding with the theory of ecological structural stability, which revolves around the dynamics on a regional scale. Our communities, in contrast, are trophic communities, operate always near local and regional species equilibrium, i.e. in the regime where O’Sullivan and coauthors18 find the most prominent form of the basic patterns. Comparing the patterns we observe, we also see SADs that are left skewed, and a local clumping of species. We did not analyse the spatial correlation between species. As we have trophic layers of species there will be some correlation between predators and their prey as they can only persist in a habitat if prey is present. In addition to the results obtained by O’Sullivan et al.18, we also derive liefetime distributions, i.e., a paleoecological pattern that also seem to be connected to the metacommunity dynamics. This might indicate that spatial non-correlation is not the most important factor in the mechanisms producing macroecological patterns.
    To conclude, our evolutionary food web model produces empirically well studied ecological and paleological patterns. We thus are armed with a valuable tool to broaden our understanding of the mechanisms behind those patterns. Our findings that trophic position influences geographic range and lifetime of a species might motivate further work regarding the interplay of abiotic and trophic factors on range expansion on evolutionary time scales.
    More generally, evolutionary models can assist us in forming a deeper knowledge of the processes that lead to what is remnant in fossils. As recently pointed out by Marshall46 in his fifth law of paleobiology, extinction erases information. It is a strength of evolutionary food web models that they allow us to study processes whose extent eludes direct observations. More

  • in

    Capital-income breeding in wild boar: a comparison between two sexes

    1.
    Bednekoff, P. A. Life histories and Predation risk. In Encyclopedia of Animal Behavior 283–287 (Elsevier, Amsterdam, 2010).
    Google Scholar 
    2.
    Jönsson, K. I. Capital and income breeding as alternative tactics of resource use in reproduction. Oikos 78, 57 (1997).
    Article  Google Scholar 

    3.
    Stephens, P. A., Boyd, I. L., McNamara, J. M. & Houston, A. I. Capital breeding and income breeding: their meaning, measurement, and worth. Ecology 90, 2057–2067 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    4.
    Kerby, J. & Post, E. Capital and income breeding traits differentiate trophic match–mismatch dynamics in large herbivores. Philos. Trans. R. Soc. B 368, 20120484 (2013).
    Article  Google Scholar 

    5.
    Williams, C. T. et al. Seasonal reproductive tactics: annual timing and the capital-to-income breeder continuum. Philos. Trans. R. Soc. B 372, 20160250 (2017).
    Article  Google Scholar 

    6.
    Apollonio, M. et al. Capital-income breeding in male ungulates: Causes and consequences of strategy differences among species. Front. Ecol. Evol. 8, 308 (2020).
    Article  Google Scholar 

    7.
    Brivio, F., Grignolio, S. & Apollonio, M. To feed or not to feed? Testing different hypotheses on rut-induced hypophagia in a mountain ungulate. Ethology 116, 406–415 (2010).
    Article  Google Scholar 

    8.
    Corlatti, L. & Bassano, B. Contrasting alternative hypotheses to explain rut-induced hypophagia in territorial male chamois. Ethology 120, 32–41 (2014).
    Article  Google Scholar 

    9.
    Miquelle, D. G. Why don’t bull moose eat during the rut?. Behav. Ecol. Sociobiol. 27, 145–151 (1990).
    Article  Google Scholar 

    10.
    Apollonio, M. & Di Vittorio, I. Feeding and reproductive behaviour in fallow bucks (Dama dama). Naturwissenschaften 91, 579–584 (2004).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    11.
    Mysterud, A., Langvatn, R. & Stenseth, N. C. Patterns of reproductive effort in male ungulates. J. Zool. 264, 209–215 (2004).
    Article  Google Scholar 

    12.
    Coltman, D. W., Festa-Bianchet, M., Jorgenson, J. T. & Strobeck, C. Age-dependent sexual selection in bighorn rams. Proc. R. Soc. Lond. B 269, 165–172 (2002).
    CAS  Article  Google Scholar 

    13.
    Apollonio, M., Brivio, F., Rossi, I., Bassano, B. & Grignolio, S. Consequences of snowy winters on male mating strategies and reproduction in a mountain ungulate. Behav. Process. 98, 44–50 (2013).
    Article  Google Scholar 

    14.
    Mysterud, A., Solberg, E. J. & Yoccoz, N. G. Ageing and reproductive effort in male moose under variable levels of intrasexual competition. J. Anim. Ecol. 74, 742–754 (2005).
    Article  Google Scholar 

    15.
    Garel, M. et al. Sex-specific growth in Alpine Chamois. J. Mammal. 90, 954–960 (2009).
    Article  Google Scholar 

    16.
    Mason, T. H. E. et al. Contrasting life histories in neighbouring populations of a large mammal. PLoS ONE 6, e28002 (2011).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    17.
    Dardaillon, M. Le sanglier et le milieu Camarguais: Dynamique Coadaptative. (1984).

    18.
    Spitz, F., Valet, G. & Lehr Brisbin, I. Variation in body mass of wild boars from southern France. J. Mammal. 79, 251–259 (1998).
    Article  Google Scholar 

    19.
    Servanty, S., Gaillard, J., Toïgo, C., Brandt, S. & Baubet, E. Pulsed resources and climate-induced variation in the reproductive traits of wild boar under high hunting pressure. J. Anim. Ecol. 78, 1278–1290 (2009).
    Article  Google Scholar 

    20.
    Gamelon, M. et al. Fluctuating food resources influence developmental plasticity in wild boar. Biol. Lett. 9, 20130419 (2013).
    PubMed  PubMed Central  Article  Google Scholar 

    21.
    Frauendorf, M., Gethöffer, F., Siebert, U. & Keuling, O. The influence of environmental and physiological factors on the litter size of wild boar (Sus scrofa) in an agriculture dominated area in Germany. Sci. Total Environ. 541, 877–882 (2016).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    22.
    Gamelon, M. et al. Reproductive allocation in pulsed-resource environments: a comparative study in two populations of wild boar. Oecologia 183, 1065–1076 (2017).
    ADS  PubMed  Article  PubMed Central  Google Scholar 

    23.
    Massei, G., Genov, P. V. & Staines, B. W. Diet, food availability and reproduction of wild boar in a Mediterranean coastal area. Acta Theriol. (Warsz.) 41, 307–320 (1996).
    Article  Google Scholar 

    24.
    Schley, L. & Roper, T. J. Diet of wild boar Sus scrofa in Western Europe, with particular reference to consumption of agricultural crops. Mamm. Rev. 33, 43–56 (2003).
    Article  Google Scholar 

    25.
    Canu, A. et al. Reproductive phenology and conception synchrony in a natural wild boar population. Hystrix 26, 77–84 (2015).
    Google Scholar 

    26.
    Allen, J. A. The influence of physical conditions in the genesis of species. Radic. Rev. 1, 108–140 (1877).
    Google Scholar 

    27.
    Fernández-Llario, P., Carranza, J. & De Trucios, S. H. Social organization of the wild boar (Sus scrofa) in Doñana National Park. Misc. Zool. 19, 9–18 (1996).
    Google Scholar 

    28.
    Bywater, K. A., Apollonio, M., Cappai, N. & Stephens, P. A. Litter size and latitude in a large mammal: the wild boar Sus scrofa. Mamm. Rev. 40, 212–220 (2010).
    Google Scholar 

    29.
    Merta, D., Mocała, P., Pomykacz, M. & Frąckowiak, W. Autumn-winter diet and fat reserves of wild boars (Sus scrofa) inhabiting forest and forest-farmland environment in south-western Poland. J. Vertebr. Biol. 63, 95–102 (2014).
    Google Scholar 

    30.
    Ježek, M., Štípek, K., Kušta, T., Červený, J. & Vícha, J. Reproductive and morphometric characteristics of wild boar (Sus scrofa) in the Czech Republic. J. For. Sci. 57, 285–292 (2011).
    Article  Google Scholar 

    31.
    Markina, F. A., Cortezo, R. G. & Gómez, C.S.-R. Physical development of wild boar in the Cantabric Mountains, Álava, Nothern Spain. Galemys Bol. Inf Soc. Esp. Para Conserv. Estud. Los Mamíferos 16, 25–34 (2004).
    Google Scholar 

    32.
    Gallo Orsi, U., Macchi, E., Perrone, A. & Durio, P. Biometric data and growth rates of a wild boar population living in the Italian Alps. J. Mt. Ecol. 3, 60–63 (1995).
    Google Scholar 

    33.
    Pedone, P., Mattioli, S. & Mattioli, L. Body size and growth patterns in wild boars of Tuscany, Central Italy. J. Mt. Ecol. 3, 66–68 (1995).
    Google Scholar 

    34.
    Šprem, N. et al. Morphometrical analysis of reproduction traits for the wild boar (Sus scrofa L.) in Croatia. Agric. Conspec. Sci. 76, 263–265 (2011).
    Google Scholar 

    35.
    Merli, E., Grignolio, S., Marcon, A. & Apollonio, M. Wild boar under fire: the effect of spatial behaviour, habitat use and social class on hunting mortality. J. Zool. 303, 155–164 (2017).
    Article  Google Scholar 

    36.
    Poteaux, C. et al. Socio-genetic structure and mating system of a wild boar population. J. Zool. 278, 116–125 (2009).
    Article  Google Scholar 

    37.
    Mauget, R. & Boissin, J. Seasonal changes in testis weight and testosterone concentration in the European wild boar (Sus scrofa L.). Anim. Reprod. Sci. 13, 67–74 (1987).
    CAS  Article  Google Scholar 

    38.
    Bisi, F. et al. Climate, tree masting and spatial behaviour in wild boar (Sus scrofa L.): Insight from a long-term study. Ann. For. Sci. 75, 46 (2018).
    Article  Google Scholar 

    39.
    Keuling, O., Stier, N. & Roth, M. How does hunting influence activity and spatial usage in wild boar Sus scrofa L.?. Eur. J. Wildl. Res. 54, 729–737 (2008).
    Article  Google Scholar 

    40.
    Brivio, F. et al. An analysis of intrinsic and extrinsic factors affecting the activity of a nocturnal species: the wild boar. Mamm. Biol. 84, 73–81 (2017).
    Article  Google Scholar 

    41.
    Singer, F. J., Otto, D. K., Tipton, A. R. & Hable, C. P. Home ranges, movements, and habitat use of European wild boar in Tennessee. J. Wildl. Manag. 45, 343–353 (1981).
    Article  Google Scholar 

    42.
    Dardaillon, M. Wild boar social groupings and their seasonal changes in the Camargue, southern France. Z. Für Säugetierkd. 53, 22–30 (1988).
    Google Scholar 

    43.
    Treyer, D. et al. Influence of sex, age and season on body weight, energy intake and endocrine parameter in wild living wild boars in southern Germany. Eur. J. Wildl. Res. 58, 373–378 (2012).
    Article  Google Scholar 

    44.
    Festa-Bianchet, M. The cost of trying: weak interspecific correlations among life-history components in male ungulates. Can. J. Zool. 90, 1072–1085 (2012).
    Article  Google Scholar 

    45.
    Knott, K. K., Barboza, P. S. & Bowyer, R. T. Growth in arctic ungulates: postnatal development and organ maturation in Rangifer tarandus and Ovibos moschatus. J. Mammal. 86, 121–130 (2005).
    Article  Google Scholar 

    46.
    Briedermann, L. Wild boars. Deutscher Landwirtschaftsverlag (1990).

    47.
    Chianucci, F. et al. Multi-temporal dataset of stand and canopy structural data in temperate and Mediterranean coppice forests. Ann. For. Sci. 76, 80 (2019).
    Article  Google Scholar 

    48.
    Zullinger, E. M., Ricklefs, R. E., Redford, K. H. & Mace, G. M. Fitting sigmoidal equations to mammalian growth curves. J. Mammal. 65, 607–636 (1984).
    Article  Google Scholar 

    49.
    Sand, H., Cederlund, G. & Danell, K. Geographical and latitudinal variation in growth patterns and adult body size of Swedish moose (Alces alces). Oecologia 102, 433–442 (1995).
    ADS  PubMed  Article  PubMed Central  Google Scholar 

    50.
    R Development Core Team. R: A Language and Environment for Statistical Computing (R Foundation for Statistical Computing, Vienna, 2015).
    Google Scholar 

    51.
    Henry, V. G. Length of estrous cycle and gestation in European Wild Hogs. J. Wildl. Manag. 32, 406 (1968).
    Article  Google Scholar 

    52.
    Vericad Corominas, J. R. Estimación de la edad fetal y períodos de concepción y parto del jabalí (Sus scrofa L.) en los Pirineos occidentales. (1981).

    53.
    Zuur, A., Ieno, E. N., Walker, N., Saveliev, A. A. & Smith, G. M. Mixed Effects Models and Extensions in Ecology with R (Springer Science & Business Media, Berlin, 2009).
    Google Scholar 

    54.
    Breiman, L. Random forests. Mach. Learn. 45, 5–32 (2001).
    MATH  Article  Google Scholar 

    55.
    Symonds, M. R. & Moussalli, A. A brief guide to model selection, multimodel inference and model averaging in behavioural ecology using Akaike’s information criterion. Behav. Ecol. Sociobiol. 65, 13–21 (2011).
    Article  Google Scholar  More

  • in

    Primary and secondary aerenchyma oxygen transportation pathways of Syzygium kunstleri (King) Bahadur & R. C. Gaur adventitious roots in hypoxic conditions

    1.
    Boyer, J. S. Plant productivity and environment. Science 218, 443–448 (1982).
    ADS  CAS  PubMed  Article  Google Scholar 
    2.
    Abiko, T. et al. Enhanced formation of aerenchyma and induction of a barrier to radial oxygen loss in adventitious roots of Zea nicaraguensis contribute to its waterlogging tolerance as compared with maize (Zea mays ssp mays). Plant Cell Environ. 35, 1618–1630 (2012).
    CAS  PubMed  Article  Google Scholar 

    3.
    Jackson, M. B. Ethylene and responses of plants to soil waterlogging and submergence. Annu. Rev. Plant Physiol. Plant Mol. Biol. 36, 145–174 (1985).
    CAS  Article  Google Scholar 

    4.
    Colmer, T. D. & Voesenek, L. A. C. J. Flooding tolerance: Suites of plant traits in variable environments. Funct. Plant Biol. 36, 665–681 (2009).
    CAS  PubMed  Article  Google Scholar 

    5.
    Bailey-Serres, J. & Voesenek, L. A. C. J. Flooding stress: Acclimations and genetic diversity. Annu. Rev. Plant Biol. 59, 313–339 (2008).
    CAS  PubMed  Article  Google Scholar 

    6.
    Colmer, T. D. & Greenway, H. Ion transport in seminal and adventitious roots of cereals during O2 deficiency. J. Exp. Bot. 62, 39–57 (2011).
    CAS  PubMed  Article  Google Scholar 

    7.
    Huang, S., Greenway, H. & Colmer, T. D. Responses of coleoptiles of intact rice seedlings to anoxia: K+ net uptake from the external solution and translocation from the caryopses. Ann. Bot. 91, 271–278 (2003).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    8.
    Vartapetian, B. B. et al. Functional electron microscopy in studies of plant response and adaptation to anaerobic stress. Ann. Bot. 91, 155–172 (2003).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    9.
    Visser, E. J. W., Voesenek, L. A. C. J., Vartapetian, B. B. & Jackson, M. B. Flooding and plant growth. Ann. Bot. 91, 107–109 (2003).
    CAS  PubMed Central  Article  PubMed  Google Scholar 

    10.
    Voesenek, L. A. & Bailey-Serres, J. Flood adaptive traits and processes: An overview. New Phytol. 206, 57–73 (2015).
    CAS  PubMed  Article  Google Scholar 

    11.
    Evans, D. E. Aerenchyma formation. New Phytol. 161, 35–49 (2004).
    Article  Google Scholar 

    12.
    Armstrong, W. Aeration in higher plants. In Advances in Botanical Research (ed. Woolhouse, H. W.) (Academic Press, Burlington, 1980).
    Google Scholar 

    13.
    Colmer, T. D. Aerenchyma and an inducible barrier to radial oxygen loss facilitate root aeration in upland, paddy and deep-water rice (Oryza sativa L.). Ann. Bot. 91, 301–309 (2003).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    14.
    Jackson, M. B. & Armstrong, W. Formation of aerenchyma and the processes of plant ventilation in relation to soil flooding and submergence. Plant Biology 1, 274–287 (1999).
    CAS  Article  Google Scholar 

    15.
    Seago, J. L. et al. A re-examination of the root cortex in wetland flowering plants with respect to aerenchyma. Ann. Bot. 96, 565–579 (2005).
    PubMed  Article  Google Scholar 

    16.
    Drew, M. C., He, C. J. & Morgan, P. W. Programmed cell death and aerenchyma formation in roots. Trends Plant Sci. 5, 123–127 (2000).
    CAS  PubMed  Article  Google Scholar 

    17.
    Yamauchi, T., Rajhi, I. & Nakazono, M. Lysigenous aerenchyma formation in maize root is confined to cortical cells by regulation of genes related to generation and scavenging of reactive oxygen species. Plant Signal. Behav. 6, 759–761 (2011).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    18.
    Takahashi, H., Yamauchi, T., Colmer, T. D. & Nakazono, M. Aerenchyma formation in plants. in Low-Oxygen Stress in Plants: Oxygen Sensing and Adaptive Responses to Hypoxia 247–265. (Springer, Wien, 2014).

    19.
    Stevens, K. J., Peterson, R. L. & Reader, R. J. The aerenchymatous phellem of Lythrum salicaria (L.): A pathway for gas transport and its role in flood tolerance. Ann. Bot. 89, 621–625 (2002).
    PubMed  PubMed Central  Article  Google Scholar 

    20.
    Shimamura, S., Mochizuki, T., Nada, Y. & Fukuyama, M. Formation and function of secondary aerenchyma in hypocotyl, roots and nodules of soybean (Glycine max) under flooded conditions. Plant Soil 251, 351–359 (2003).
    CAS  Article  Google Scholar 

    21.
    Shimamura, S., Yamamoto, R., Nakamura, T., Shimada, S. & Komatsu, S. Stem hypertrophic lenticels and secondary aerenchyma enable oxygen transport to roots of soybean in flooded soil. Ann. Bot. 106, 277–284 (2010).
    PubMed  PubMed Central  Article  Google Scholar 

    22.
    De Simone, O. et al. Impact of root morphology on metabolism and oxygen distribution in roots and rhizosphere from two Central Amazon floodplain tree species. Funct. Plant Biol. 29, 1025–1035 (2002).
    PubMed  Article  Google Scholar 

    23.
    Colmer, T. D. & Pedersen, O. Oxygen dynamics in submerged rice (Oryza sativa). New Phytol. 178, 326–334 (2008).
    CAS  PubMed  Article  Google Scholar 

    24.
    Haase, K., De Simone, O., Junk, W. J. & Schmidt, W. Internal oxygen transport in cuttings from flood-adapted várzea tree species. Tree Physiol. 23, 1069–1076 (2003).
    PubMed  Article  Google Scholar 

    25.
    Sou, H. D., Masumori, M., Kurokochi, H. & Tange, T. Histological observation of primary and secondary aerenchyma formation in adventitious roots of Syzygium kunstleri (King) Bahadur and R. C. Gaur grown in hypoxic medium. Forests 10, 137 (2019).
    Article  Google Scholar 

    26.
    Rubinigg, M., Stulen, I., Elzenga, J. T. M. & Colmer, T. D. Spatial patterns of radial oxygen loss and nitrate net flux along adventitious roots of rice raised in aerated or stagnant solution. Funct. Plant Biol. 29, 1475–1481 (2002).
    CAS  PubMed  Article  Google Scholar 

    27.
    Kotula, L., Ranathunge, K., Schreiber, L. & Steudle, E. Functional and chemical comparison of apoplastic barriers to radial oxygen loss in roots of rice (Oryza sativa L.) grown in aerated or deoxygenated solution. J. Exp. Bot. 60, 2155–2167 (2009).
    CAS  PubMed  Article  Google Scholar 

    28.
    Shiono, K. et al. Contrasting dynamics of radial O2-loss barrier induction and aerenchyma formation in rice roots of two lengths. Ann. Bot. 107, 89–99 (2011).
    CAS  PubMed  Article  Google Scholar 

    29.
    Watanabe, K., Nishiuchi, S., Kulichikhin, K. & Nakazono, M. Does suberin accumulation in plant roots contribute to waterlogging tolerance?. Front. Plant Sci. 4, 178 (2013).
    PubMed  PubMed Central  Article  Google Scholar 

    30.
    Khan, N. et al. Root iron plaque on wetland plants as dynamic pool of nutrients and contaminants. In Advances in Agronomy Vol. 138 (ed. Sparks, D. L.) 1–96 (Academic Press, Cambridge, 2016).
    Google Scholar 

    31.
    Uteau, D. et al. Oxygen and redox potential gradients in the rhizosphere of alfalfa grown on a loamy soil. J. Plant Nutr. Soil Sci. 178, 278–287 (2015).
    CAS  Article  Google Scholar 

    32.
    Tian, C., Wang, C., Tian, Y., Wu, X. & Xiao, B. Root radial oxygen loss and the effects on rhizosphere microarea of two submerged plants. Polish J. Environ. Studies 24, 1795–1802 (2015).
    Article  Google Scholar 

    33.
    Shimamura, S., Mochizuki, T., Nada, Y. & Fukuyama, M. Secondary aerenchyma formation and its relation to nitrogen fixation in root nodules of soybean plants (Glycine max) grown under flooded conditions. Plant Product. Sci. 5, 294–300 (2002).
    CAS  Article  Google Scholar 

    34.
    Shiba, H. & Daimon, H. Histological observation of secondary aerenchyma formed immediately after flooding in Sesbania cannabina and S. rostrata. Plant Soil 255, 209–215 (2003).
    CAS  Article  Google Scholar 

    35.
    Somavilla, N. S. & Graciano-Ribeiro, D. Ontogeny and characterization of aerenchymatous tissues of Melastomataceae in the flooded and well-drained soils of a Neotropical savanna. Flora 207, 212–222 (2012).
    Article  Google Scholar 

    36.
    Thomas, A. L., Guerreiro, S. M. C. & Sodek, L. Aerenchyma formation and recovery from hypoxia of the flooded root system of nodulated soybean. Ann. Bot. 96, 1191–1198 (2005).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    37.
    Wiengweera, A., Greenway, H. & Thomson, C. J. The use of agar nutrient solution to simulate lack of convection in waterlogged soils. Ann. Bot. 80, 115–123 (1997).
    Article  Google Scholar 

    38.
    Dacey, J. W. Internal winds in water lilies: An adaptation for life in anaerobic sediments. Science 210, 1017–1019 (1980).
    ADS  CAS  PubMed  Article  Google Scholar 

    39.
    Drew, M. C., Saglio, P. H. & Pradet, A. J. P. Larger adenylate energy charge and ATP/ADP ratios in aerenchymatous roots of Zea mays in anaerobic media as a consequence of improved internal oxygen transport. Planta 165, 51–58 (1985).
    CAS  PubMed  Article  Google Scholar 

    40.
    Drew, M. C. Oxygen deficiency and root metabolism: Injury and acclimation under hypoxia and anoxia. Annu. Rev. Plant Physiol. Plant Mol. Biol. 48, 223–250 (1997).
    CAS  PubMed  Article  Google Scholar 

    41.
    Shimamura, S., Yoshida, S. & Mochizuki, T. Cortical aerenchyma formation in hypocotyl and adventitious roots of Luffa cylindrica subjected to soil flooding. Ann. Bot. 100, 1431–1439 (2007).
    PubMed  PubMed Central  Article  Google Scholar 

    42.
    Armstrong, W., Cousins, D., Armstrong, J., Turner, D. W. & Beckett, P. M. Oxygen distribution in wetland plant roots and permeability barriers to gas-exchange with the rhizosphere: A microelectrode and modelling study with Phragmites australis. Ann. Bot. 86, 687–703 (2000).
    Article  Google Scholar 

    43.
    Herzog, M. & Pedersen, O. Partial versus complete submergence: Snorkelling aids root aeration in Rumex palustris but not in R. acetosa. Plant Cell Environ. 37, 2381–2390 (2014).
    CAS  PubMed  Google Scholar 

    44.
    Tanaka, K., Masumori, M., Yamanoshita, T. & Tange, T. Morphological and anatomical changes of Melaleuca cajuputi under submergence. Trees 25, 695–704 (2011).
    Article  Google Scholar 

    45.
    Armstrong, W. Polarographic oxygen electrodes and their use in plant aeration studies. Proc. R. Soc. Edinburgh Sect. B. Biol. Sci. 102, 511–527 (1994).
    Article  Google Scholar 

    46.
    Hitchman, M. L. Measurement of Dissolved Oxygen (Wiley, New York, 1978).
    Google Scholar 

    47.
    Ober, E. S. & Sharp, R. E. A microsensor for direct measurement of O2 partial pressure within plant tissues. J. Exp. Bot. 47, 447–454 (1996).
    CAS  Article  Google Scholar  More

  • in

    Chimpanzees balance resources and risk in an anthropogenic landscape of fear

    1.
    Dirzo, R. et al. Defaunation in the anthropocene. Science 345, 401–406 (2014).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 
    2.
    Boivin, N. L. et al. Ecological consequences of human niche construction: examining long-term anthropogenic shaping of global species distributions. Proc. Natl. Acad. Sci. 113, 6388–6396 (2016).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    3.
    Newbold, T. et al. Global effects of land use on local terrestrial biodiversity. Nature 520, 45–50 (2015).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    4.
    Hagen, M. et al. Biodiversity, species interactions and ecological networks in a fragmented world. In Advances in Ecological Research Vol. 46 (eds Jacob, U. & Woodward, G.) 89–210 (Academic Press, Cambridge, 2012).
    Google Scholar 

    5.
    Gallego-Zamorano, J. et al. Combined effects of land use and hunting on distributions of tropical mammals. Conserv. Biol. 34, 1271–1280 (2020).
    PubMed  PubMed Central  Article  Google Scholar 

    6.
    Ellis, E. C. & Ramankutty, N. Putting people in the map: anthropogenic biomes of the world. Front. Ecol. Environ. 6, 439–447 (2008).
    Article  Google Scholar 

    7.
    Estrada, A., Raboy, B. E. & Oliveira, L. C. Agroecosystems and primate conservation in the tropics: a review. Am. J. Primatol. 74, 696–711 (2012).
    PubMed  Article  Google Scholar 

    8.
    Bhagwat, S. A., Willis, K. J., Birks, H. J. B. & Whittaker, R. J. Agroforestry: a refuge for tropical biodiversity?. Trends Ecol. Evol. 23, 261–267 (2008).
    PubMed  Article  Google Scholar 

    9.
    Galán-Acedo, C. et al. The conservation value of human-modified landscapes for the world’s primates. Nat. Commun. 10, 1–8 (2019).
    Article  CAS  Google Scholar 

    10.
    Arroyo-Rodríguez, V. et al. Designing optimal human-modified landscapes for forest biodiversity conservation. Ecol. Lett. 23, 1404–1420 (2020).
    PubMed  Article  Google Scholar 

    11.
    Kshettry, A., Vaidyanathan, S., Sukumar, R. & Athreya, V. Looking beyond protected areas: identifying conservation compatible landscapes in agro-forest mosaics in north-eastern India. Glob. Ecol. Conserv. 22, e00905 (2020).
    Article  Google Scholar 

    12.
    Osborn, F. V. & Hill, C. M. Techiques to reduce crop loss: human and technical dimensions in Africa. In People and Wildlife, Conflict or Co-existence? 72–85 (Cambridge University Press, Cambridge, 2005).

    13.
    McLennan, M. R. & Asiimwe, C. Cars kill chimpanzees: case report of a wild chimpanzee killed on a road at Bulindi, Uganda. Primates J. Primatol. 57, 377–388 (2016).
    Article  Google Scholar 

    14.
    Chapman, C. A. et al. Do food availability, parasitism, and stress have synergistic effects on red colobus populations living in forest fragments?. Am. J. Phys. Anthropol. 131, 525–534 (2006).
    PubMed  Article  Google Scholar 

    15.
    Goldberg, T. L., Gillespie, T. R., Rwego, I. B., Estoff, E. L. & Chapman, C. A. Forest fragmentation as cause of bacterial transmission among nonhuman primates, humans, and livestock, Uganda. Emerg. Infect. Dis. 14, 1375–1382 (2008).
    PubMed  PubMed Central  Article  Google Scholar 

    16.
    McLennan, M. R., Hyeroba, D., Asiimwe, C., Reynolds, V. & Wallis, J. Chimpanzees in mantraps: lethal crop protection and conservation in Uganda. Oryx 46, 598–603 (2012).
    Article  Google Scholar 

    17.
    Kalema-Zikusoka, G., Rubanga, S., Mutahunga, B. & Sadler, R. Prevention of Cryptosporidium and GIARDIA at the human/gorilla/livestock interface. Front. Public Health 6, (2018).

    18.
    Kenney, J., Allendorf, F. W., McDougal, C. & Smith, J. L. D. How much gene flow is needed to avoid inbreeding depression in wild tiger populations?. Proc. R. Soc. B Biol. Sci. 281, 20133337 (2014).
    Article  Google Scholar 

    19.
    Willems, E. P. & Hill, R. A. Predator-specific landscapes of fear and resource distribution: effects on spatial range use. Ecology 90, 546–555 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    20.
    Coleman, B. T. & Hill, R. A. Living in a landscape of fear: the impact of predation, resource availability and habitat structure on primate range use. Anim. Behav. 88, 165–173 (2014).
    Article  Google Scholar 

    21.
    Palmer, M. S., Fieberg, J., Swanson, A., Kosmala, M. & Packer, C. A ‘dynamic’ landscape of fear: prey responses to spatiotemporal variations in predation risk across the lunar cycle. Ecol. Lett. 20, 1364–1373 (2017).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    22.
    Laundré, J. W., Hernandez, L. & Ripple, W. J. The landscape of fear: ecological implications of being afraid. Open Ecol. J. 3, (2010).

    23.
    Theuerkauf, J. & Rouys, S. Habitat selection by ungulates in relation to predation risk by wolves and humans in the Białowieża Forest, Poland. For. Ecol. Manag. 256, 1325–1332 (2008).
    Article  Google Scholar 

    24.
    Ciuti, S. et al. Effects of humans on behaviour of wildlife exceed those of natural predators in a landscape of fear. PLoS ONE 7, e50611 (2012).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    25.
    Nowak, K., Wimberger, K., Richards, S. A., Hill, R. A. & le Roux, A. Samango monkeys (Cercopithecus albogularis labiatus) manage risk in a highly seasonal, human-modified landscape in Amathole Mountains, South Africa. Int. J. Primatol. 38, 194–206 (2017).
    PubMed  Article  Google Scholar 

    26.
    Suraci, J. P., Clinchy, M., Zanette, L. Y. & Wilmers, C. C. Fear of humans as apex predators has landscape-scale impacts from mountain lions to mice. Ecol. Lett. 22, 1578–1586 (2019).
    PubMed  Article  Google Scholar 

    27.
    Carter, N. H., Shrestha, B. K., Karki, J. B., Pradhan, N. M. B. & Liu, J. Coexistence between wildlife and humans at fine spatial scales. Proc. Natl. Acad. Sci. 109, 15360–15365 (2012).
    ADS  CAS  PubMed  Article  Google Scholar 

    28.
    Carter, N. H., Jasny, M., Gurung, B. & Liu, J. Impacts of people and tigers on leopard spatiotemporal activity patterns in a global biodiversity hotspot. Glob. Ecol. Conserv. 3, 149–162 (2015).
    Article  Google Scholar 

    29.
    Lamb, C. T. et al. The ecology of human–carnivore coexistence. Proc. Natl. Acad. Sci. 117, 17876–17883. https://doi.org/10.1073/pnas.1922097117 (2020).
    CAS  Article  PubMed  Google Scholar 

    30.
    Bryson-Morrison, N., Tzanopoulos, J., Matsuzawa, T. & Humle, T. Activity and habitat use of chimpanzees (Pan troglodytes verus) in the anthropogenic landscape of Bossou, Guinea, West Africa. Int. J. Primatol. 38, 282–302 (2017).
    PubMed  PubMed Central  Article  Google Scholar 

    31.
    de Almeida-Rocha, J. M., Peres, C. A. & Oliveira, L. C. Primate responses to anthropogenic habitat disturbance: a pantropical meta-analysis. Biol. Conserv. 215, 30–38 (2017).
    Article  Google Scholar 

    32.
    Galán‐Acedo, C., Arroyo‐Rodríguez, V., Cudney‐Valenzuela, S. J. & Fahrig, L. A global assessment of primate responses to landscape structure. Biol. Rev. 94, 1605–1618 (2019).
    PubMed  Article  PubMed Central  Google Scholar 

    33.
    Garriga, R. M. et al. Factors influencing wild chimpanzee (Pan troglodytes verus) relative abundance in an agriculture-swamp matrix outside protected areas. PLoS ONE 14, e0215545 (2019).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    34.
    Hockings, K. J., Anderson, J. R. & Matsuzawa, T. Road crossing in chimpanzees: a risky business. Curr. Biol. 16, R668–R670 (2006).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    35.
    Estrada, A. et al. Impending extinction crisis of the world’s primates: why primates matter. Sci. Adv. 3, e1600946 (2017).
    ADS  PubMed  PubMed Central  Article  Google Scholar 

    36.
    IUCN SSC Primate Specialist Group. Regional action plan for the conservation of western chimpanzees (Pan troglodytes verus) 2020–2030. (2020).

    37.
    Kalan, A. K. et al. Environmental variability supports chimpanzee behavioural diversity. Nat. Commun. 11, 4451 (2020).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    38.
    Hockings, K. J., Anderson, J. R. & Matsuzawa, T. Socioecological adaptations by chimpanzees, Pan troglodytes verus, inhabiting an anthropogenically impacted habitat. Anim. Behav. 83, 801–810 (2012).
    Article  Google Scholar 

    39.
    McLennan, M. R. & Hockings, K. J. Wild chimpanzees show group differences in selection of agricultural crops. Sci. Rep. 4, 5956 (2014).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    40.
    Kalan, A. K. et al. Novelty response of wild African apes to camera traps. Curr. Biol.  29, 1211–1217.e3 (2019).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    41.
    Hockings, K. J. & McLennan, M. R. From forest to farm: systematic review of cultivar feeding by chimpanzees—management implications for wildlife in anthropogenic landscapes. PLoS ONE 7, e33391 (2012).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    42.
    Hockings, K. J., Anderson, J. R. & Matsuzawa, T. Use of wild and cultivated foods by chimpanzees at Bossou, Republic of Guinea: feeding dynamics in a human-influenced environment. Am. J. Primatol. 71, 636–646 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    43.
    McLennan, M. R. Diet and feeding ecology of chimpanzees (Pan troglodytes) in Bulindi, Uganda: foraging strategies at the forest–farm interface. Int. J. Primatol. 34, 585–614 (2013).
    Article  Google Scholar 

    44.
    McLennan, M. R. & Ganzhorn, J. U. Nutritional characteristics of wild and cultivated foods for chimpanzees (Pan troglodytes) in agricultural landscapes. Int. J. Primatol. 38, 122–150 (2017).
    Article  Google Scholar 

    45.
    Matthews, A. & Matthews, A. Survey of gorillas (Gorilla gorilla gorilla) and chimpanzees (Pan troglodytes troglodytes) in Southwestern Cameroon. Primates 45, 15–24 (2004).
    PubMed  Article  PubMed Central  Google Scholar 

    46.
    Morgan, D. et al. African apes coexisting with logging: comparing chimpanzee (Pan troglodytes troglodytes) and gorilla (Gorilla gorilla gorilla) resource needs and responses to forestry activities. Biol. Conserv. 218, 277–286 (2018).
    Article  Google Scholar 

    47.
    Krief, S. et al. Wild chimpanzees on the edge: nocturnal activities in croplands. PLoS ONE 9, e109925 (2014).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    48.
    Riley, E. P. & Priston, N. E. C. Macaques in farms and folklore: exploring the human–nonhuman primate interface in Sulawesi, Indonesia. Am. J. Primatol. 72, 848–854 (2010).
    PubMed  Article  PubMed Central  Google Scholar 

    49.
    Parathian, H. E., McLennan, M. R., Hill, C. M., Frazão-Moreira, A. & Hockings, K. J. Breaking through disciplinary barriers: human–wildlife interactions and multispecies ethnography. Int. J. Primatol. 39, 749–775. https://doi.org/10.1007/s10764-018-0027-9 (2018).
    Article  PubMed  PubMed Central  Google Scholar 

    50.
    Fuentes, A. & Gamerl, S. Disproportionate participation by age/sex classes in aggressive interactions between long-tailed macaques (Macaca fascicularis) and human tourists at Padangtegal monkey forest, Bali, Indonesia. Am. J. Primatol. 66, 197–204 (2005).
    PubMed  Article  Google Scholar 

    51.
    McLennan, M. R. & Hockings, K. J. The aggressive apes? Causes and contexts of great ape attacks on local persons. In Problematic Wildlife (ed. Angelici, F. M.) 373–394 (Springer, Cham, 2016). https://doi.org/10.1007/978-3-319-22246-2_18.

    52.
    Hill, C. M. & Webber, A. D. Perceptions of nonhuman primates in human–wildlife conflict scenarios. Am. J. Primatol. 72, 919–924 (2010).
    PubMed  Article  Google Scholar 

    53.
    McLennan, M. R. & Hill, C. M. Troublesome neighbours: changing attitudes towards chimpanzees (Pan troglodytes) in a human-dominated landscape in Uganda. J. Nat. Conserv. 20, 219–227 (2012).
    Article  Google Scholar 

    54.
    Mito, Y. & Sprague, D. S. The Japanese and Japanese monkeys: dissonant neighbors seeking accommodation in a shared habitat. In The Macaque Connection: Cooperation and Conflict Between Humans and Macaques (eds Radhakrishna, S. et al.) 33–51 (Springer, Berlin, 2013).
    Google Scholar 

    55.
    Morzillo, A., de Beurs, K. & Martin-Mikle, C. A conceptual framework to evaluate human-wildlife interactions within coupled human and natural systems. Ecol. Soc. 19, (2014).

    56.
    Martin, J. et al. Coping with human disturbance: spatial and temporal tactics of the brown bear (Ursus arctos). Can. J. Zool. 88, 875–883 (2010).
    Article  Google Scholar 

    57.
    Hockings, K. J. et al. Chimpanzees share forbidden fruit. PLoS ONE 2, e886 (2007).
    ADS  PubMed  PubMed Central  Article  Google Scholar 

    58.
    Duvall, C. S. Human settlement ecology and chimpanzee habitat selection in Mali. Landsc. Ecol. 23, 699 (2008).
    Article  Google Scholar 

    59.
    Hockings, K. J., Parathian, H., Bessa, J. & Frazão-Moreira, A. Extensive overlap in the selection of wild fruits by chimpanzees and humans: implications for the management of complex social-ecological systems. Front. Ecol. Evol. 8, (2020).

    60.
    Nowak, K., Hill, R. A., Wimberger, K. & le Roux, A. Risk-taking in samango monkeys in relation to humans at two sites in South Africa. In Ethnoprimatology: Primate Conservation in the 21st Century (ed. Waller, M. T.) 301–314 (Springer, Berlin, 2016).
    Google Scholar 

    61.
    INE. Recenseamento Geral da População e Habitação: População por Região, Sector e Localidades por Sexo Censo 2009. 160 (2009).

    62.
    Heinicke, S. et al. Characteristics of positive deviants in western chimpanzee populations. Front. Ecol. Evol. 7, (2019).

    63.
    Bersacola, E. Zooming in on Human-Wildlife Coexistence: Primate Community Responses in a Shared Agroforest Landscape in Guinea-Bissau (Oxford Brookes University, Oxford, 2020).
    Google Scholar 

    64.
    Bessa, J., Sousa, C. & Hockings, K. J. Feeding ecology of chimpanzees (Pan troglodytes verus) inhabiting a forest-mangrove-savanna-agricultural matrix at Caiquene-Cadique, Cantanhez National Park, Guinea-Bissau. Am. J. Primatol. 77, 651–665 (2015).
    PubMed  Article  Google Scholar 

    65.
    Hockings, K. J. et al. Leprosy in wild chimpanzees. bioRxiv 2020.11.10.374371 (2020) https://doi.org/10.1101/2020.11.10.374371.

    66.
    Hockings, K. J. & Sousa, C. Differential utilization of cashew—a low-conflict crop—by sympatric humans and chimpanzees. Oryx 46, 375–381 (2012).
    Article  Google Scholar 

    67.
    Calenge, C. The package adehabitat for the R software: a tool for the analysis of space and habitat use by animals. Ecol. Model. 197, 516–519 (2006).
    Article  Google Scholar 

    68.
    Schmid, F. & Schmidt, A. Nonparametric estimation of the coefficient of overlapping—theory and empirical application. Comput. Stat. Data Anal. 50, 1583–1596 (2006).
    MathSciNet  MATH  Article  Google Scholar 

    69.
    Ridout, M. S. & Linkie, M. Estimating overlap of daily activity patterns from camera trap data. J. Agric. Biol. Environ. Stat. 14, 322–337 (2009).
    MathSciNet  MATH  Article  Google Scholar 

    70.
    Hijmans, R. J. Raster: geographic data analysis and modeling. (2020).

    71.
    Khorozyan, I., Stanton, D., Mohammed, M., Al-Rail, W. & Pittet, M. Patterns of co-existence between humans and mammals in Yemen: some species thrive while others are nearly extinct. Biodivers. Conserv. 23, 1995–2013 (2014).
    Article  Google Scholar 

    72.
    Sousa, J., Barata, A. V., Sousa, C., Casanova, C. C. N. & Vicente, L. Chimpanzee oil-palm use in southern Cantanhez National Park, Guinea-Bissau. Am. J. Primatol. 73, 485–497 (2011).
    PubMed  Article  PubMed Central  Google Scholar 

    73.
    Tutin, C. E. G. et al. Foraging profiles of sympatric lowland gorillas and chimpanzees in the Lope Reserve, Gabon [and discussion]. Philos. Trans. Biol. Sci. 334, 179–186 (1991).
    ADS  CAS  Article  Google Scholar 

    74.
    Yamakoshi, G. Dietary responses to fruit scarcity of wild chimpanzees at Bossou, Guinea: possible implications for ecological importance of tool use. Am. J. Phys. Anthropol. 106, 283–295 (1998).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    75.
    Wilson, M. L., Hauser, M. D. & Wrangham, R. W. Chimpanzees (Pan troglodytes) modify grouping and vocal behaviour in response to location-specific risk. Behaviour 144, 1621–1653 (2007).
    Article  Google Scholar 

    76.
    Lindshield, S., Danielson, B. J., Rothman, J. M. & Pruetz, J. D. Feeding in fear? How adult male western chimpanzees (Pan troglodytes verus) adjust to predation and savanna habitat pressures. Am. J. Phys. Anthropol. 163, 480–496 (2017).
    PubMed  Article  PubMed Central  Google Scholar 

    77.
    Olson, D. M. et al. Terrestrial ecoregions of the world: a new map of life on Earth. Bioscience 51, 933–938 (2001).
    Article  Google Scholar 

    78.
    Sousa, J., Vicente, L., Gippoliti, S., Casanova, C. & Sousa, C. Local knowledge and perceptions of chimpanzees in Cantanhez National Park, Guinea-Bissau. Am. J. Primatol. 76, 122–134 (2014).
    PubMed  Article  PubMed Central  Google Scholar 

    79.
    Sharma, K. et al. Conservation and people: towards an ethical code of conduct for the use of camera traps in wildlife research. Ecol. Solut. Evid. 1, e12033 (2020).
    Article  Google Scholar 

    80.
    Sun, C. et al. Tree phenology in a tropical montane forest in Rwanda. Biotropica 28, 668–681 (1996).
    Article  Google Scholar 

    81.
    McLennan, M. R. Chimpanzee Ecology and Interactions with People in an Unprotected Human-Dominated Landscape at Bulindi, Western Uganda (Oxford Brookes University, Oxford, 2010).
    Google Scholar 

    82.
    Jenks, K. E. et al. Using relative abundance indices from camera-trapping to test wildlife conservation hypotheses—an example from Khao Yai National Park, Thailand. Trop. Conserv. Sci. 4, 113–131 (2011).
    ADS  Article  Google Scholar 

    83.
    O’Brien, T. G., Kinnaird, M. F. & Wibisono, H. T. Crouching tigers, hidden prey: sumatran tiger and prey populations in a tropical forest landscape. Anim. Conserv. Forum 6, 131–139 (2003).
    Article  Google Scholar 

    84.
    Rue, H., Martino, S. & Chopin, N. Approximate Bayesian inference for latent Gaussian models by using integrated nested Laplace approximations. J. R. Stat. Soc. Ser. B Stat. Methodol. 71, 319–392 (2009).
    MathSciNet  MATH  Article  Google Scholar 

    85.
    Blangiardo, M., Cameletti, M., Baio, G. & Rue, H. Spatial and spatio-temporal models with R-INLA. Spat. Spatio-Temporal Epidemiol. 4, 33–49 (2013).
    Article  Google Scholar 

    86.
    Cameletti, M., Lindgren, F., Simpson, D. & Rue, H. Spatio-temporal modeling of particulate matter concentration through the SPDE approach. AStA Adv. Stat. Anal. 97, 109–131 (2013).
    MathSciNet  MATH  Article  Google Scholar 

    87.
    Lindgren, F., Rue, H. & Lindström, J. An explicit link between Gaussian fields and Gaussian Markov random fields: the stochastic partial differential equation approach. J. R Stat. Soc. Ser. B Stat. Methodol. 73, 423–498 (2011).
    MathSciNet  MATH  Article  Google Scholar 

    88.
    Bakka, H. et al. Spatial modeling with R-INLA: a review. Wiley Interdiscip. Rev. Comput. Stat. 10, e1443 (2018).
    MathSciNet  Article  Google Scholar 

    89.
    Noor, A. M. et al. The changing risk of Plasmodium falciparum malaria infection in Africa: 2000–10: a spatial and temporal analysis of transmission intensity. Lancet 383, 1739–1747 (2014).
    PubMed  PubMed Central  Article  Google Scholar 

    90.
    Rue, H. et al. Bayesian computing with INLA: a review. Annu. Rev. Stat. Its Appl. 4, 395–421 (2017).
    ADS  Article  Google Scholar 

    91.
    Cressie, N. & Wikle, C. K. Statistics for Spatio-Temporal Data (Wiley, Hoboken, 2015).
    Google Scholar 

    92.
    Lindgren, F. & Rue, H. Bayesian spatial modelling with R-INLA. J. Stat. Softw. 63, 1–25 (2015).
    Article  Google Scholar 

    93.
    Spiegelhalter, D. J., Best, N. G., Carlin, B. P. & Linde, A. V. D. Bayesian measures of model complexity and fit. J. R. Stat. Soc. Ser. B Stat. Methodol. 64, 583–639 (2002).
    MathSciNet  MATH  Article  Google Scholar 

    94.
    R Core Team. R: A language and environment for statistical computing. (R Foundation for Statistical Computing, 2020). More

  • in

    Time-for-space substitution in N-mixture models for estimating population trends: a simulation-based evaluation

    1.
    Williams, B. K., Nichols, J. D. & Conroy, M. J. Analysis and Management of Animal Populations (Academic Press, New York, 2002).
    Google Scholar 
    2.
    Lindenmayer, D. B. & Likens, G. E. Adaptive monitoring: a new paradigm for long-term research and monitoring. Trends Ecol. Evol. 24, 482–486 (2009).
    Article  Google Scholar 

    3.
    MacKenzie, D. I. et al. Estimating site occupancy rates when detection probabilities are less than one. Ecology 83, 2248–2255 (2002).
    Article  Google Scholar 

    4.
    Buckland, S. T., Rexstad, E. A., Marques, T. A. & Oedekoven, C. S. Distance Sampling: Methods and Applications (Springer, Berlin, 2015).
    Google Scholar 

    5.
    Royle, J. A. N-mixture models for estimating population size from spatially replicated counts. Biometrics 60, 108–115 (2004).
    MathSciNet  Article  Google Scholar 

    6.
    Kéry, M. & Royle, J. A. Applied Hierarchical Modelling in Ecology (Academic Press, New York, 2016).
    Google Scholar 

    7.
    Ariefiandy, A. et al. Evaluation of three field monitoring-density estimation protocols and their relevance to Komodo dragon conservation. Biodivers. Conserv. 23, 2473–2490 (2014).
    Article  Google Scholar 

    8.
    Romano, A. et al. Conservation of salamanders in managed forests: methods and costs of monitoring abundance and habitat selection. For. Ecol. Manag. 400, 12–18 (2017).
    Article  Google Scholar 

    9.
    Chandler, R. B., Royle, J. A. & King, D. I. Inference about density and temporary emigration in unmarked populations. Ecology 92, 1429–1435 (2011).
    Article  Google Scholar 

    10.
    Dail, D. & Madsen, L. Models for estimating abundance from repeated counts of an open metapopulation. Biometrics 67, 577–587 (2011).
    MathSciNet  CAS  Article  Google Scholar 

    11.
    Augustynczik, L. D. et al. Diversification of forest management regimes secures tree microhabitats and bird abundance under climate change. Sci. Total Environ. 650, 2717–2730 (2019).
    ADS  CAS  Article  Google Scholar 

    12.
    Peterman, W. E. & Semlitsch, R. D. Fine-scale habitat associations of a terrestrial salamander: the role of environmental gradients and implications for population dynamics. PLoS ONE 8, e62184. https://doi.org/10.1371/journal.pone.0062184 (2013).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    13.
    Balestrieri, R. et al. A guild-based approach to assessing the influence of beech forest structure on bird communities. For. Ecol. Manage. 356, 216–223 (2015).
    Article  Google Scholar 

    14.
    Barker, R. J., Schofield, M. R., Link, W. A. & Sauer, J. R. On the reliability of N-mixture models for count data. Biometrics 74, 369–377 (2017).
    MathSciNet  Article  Google Scholar 

    15.
    Link, W. A., Schofield, M. R., Barker, R. J. & Sauer, J. R. On the robustness of N-mixture models. Ecology 99, 1547–1551 (2018).
    Article  Google Scholar 

    16.
    Kéry, M. Identifiability in N-mixture models: a large-scale screening test with bird data. Ecology 99, 281–288 (2018).
    Article  Google Scholar 

    17.
    Priol, P. M. Using dynamic N-mixture models to test cavity limitation on northern flying squirrel demographic parameters using experimental nest box supplementation. Ecol. Evol. 4, 2165–2177 (2014).
    Article  Google Scholar 

    18.
    Basile, M. et al. Measuring bird abundance—a comparison of methodologies between capture/recapture and audio-visual surveys. Avocetta 40, 55–61 (2016).
    Google Scholar 

    19.
    Ficetola, G. F. et al. N-mixture models reliably estimate the abundance of small vertebrates. Sci. Rep. 8, 10357. https://doi.org/10.1038/s41598-018-28432-8 (2018).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    20.
    Costa, A., Oneto, F. & Salvidio, S. Time-for-space substitution in N-mixture modeling and population monitoring. J. Wildl. Manag. 83, 737–741 (2019).
    Article  Google Scholar 

    21.
    Yamaura, Y. et al. Modelling community dynamics based on species-level abundance models from detection/nondetection data. J. Appl. Ecol. 48, 67–75 (2011).
    Article  Google Scholar 

    22.
    Dennis, E. B., Morgan, B. J. & Ridout, M. S. Computational aspects of N-mixture models. Biometrics 71, 237–246 (2015).
    MathSciNet  Article  Google Scholar 

    23.
    Duarte, A., Adams, M. J. & Peterson, J. T. Fitting N-mixture models to count data with unmodeled heterogeneity: bias, diagnostics, and alternative approaches. Ecol. Model. 374, 51–59 (2018).
    Article  Google Scholar 

    24.
    Costa, A., Romano, A. & Salvidio, S. Reliability of multinomial N-mixture models for estimating abundance of small terrestrial vertebrates. Biodiv. Conserv. 29, 2951–2965 (2020).
    Article  Google Scholar 

    25.
    Ficetola, G. F., Romano, A., Salvidio, S. & Sindaco, R. Optimizing monitoring schemes to detect trends in abundance over broad scales. Anim. Conserv. 21, 221–231 (2018).
    Article  Google Scholar 

    26.
    Fiske, I. & Chandler, R. unmarked: an R package for fitting hierarchical models of wildlife occurrence and abundance. J. Stat. Soft. 43, 1–23 (2011).
    Article  Google Scholar 

    27.
    Kéry, M., Royle, J.A., & Meredith, M. Package AHMbook version 0.1.4 (2016).

    28.
    MacKenzie, D. I. & Bailey, L. L. Assessing the fit of site-occupancy models. J. Agric. Biol. Environ. Stat. 9, 300–318 (2004).
    Article  Google Scholar 

    29.
    Knape, J. et al. Sensitivity of binomial N-mixture models to overdispersion: the importance of assessing model fit. Met. Ecol. Evol. 9, 2102–2114 (2018).
    Article  Google Scholar 

    30.
    Mazerolle, M. J. AICcmodavg: model selection and multimodel inference based on (Q)AIC(c). R package version 2.1-1 (2017).

    31.
    McIntyre, A. P. Empirical and simulation evaluations of an abundance estimator using unmarked individuals of cryptic forest-dwelling taxa. For. Ecol. Manage. 286, 129–136 (2012).
    Article  Google Scholar 

    32.
    Veech, J. A., Ott, J. R. & Troy, J. R. Intrinsic heterogeneity in detection probability and its effect on N-mixture models. Met. Ecol. Evol. 7, 1019–1028 (2016).
    Article  Google Scholar 

    33.
    Gervasi, V. A preliminary estimate of the apennine brown bear population size based on hair-snag sampling and multiple data source mark–recapture huggins models. Ursus 19, 105–121 (2008).
    Article  Google Scholar 

    34.
    Welsh, H. N. & Conroy, M. J. A Case for using plethodontid salamanders for monitoring biodiversity and ecosystem integrity of North American forests. Conserv. Biol. 15, 558–569 (2001).
    Article  Google Scholar 

    35.
    Warton, D. I., Stoklosa, J., Guillera-Arroita, G., MacKenzie, D. I. & Welsh, A. H. Graphical diagnostics for occupancy models with imperfect detection. Met. Ecol. Evol. 8, 408–419 (2017).
    Article  Google Scholar 

    36.
    Lunghi, E. Environmental suitability models predict population density, performance and body condition for microendemic salamanders. Sci. Rep. 8, 7527. https://doi.org/10.1038/s41598-018-25704-1 (2018).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    37.
    Kopler, I. & Malkinson, D. Differential response of mammals to agricultural fences: the effects of species vagility and body size. Basic Appl. Ecol. 33, 79–88 (2018).
    Article  Google Scholar  More

  • in

    Superior predatory ability and abundance predicts potential ecological impact towards early-stage anurans by invasive ‘Killer Shrimp’ (Dikerogammarus villosus)

    1.
    Hoffmann, B. D. & Broadhurst, L. M. The economic cost of managing invasive species in Australia. NeoBiota 31, 1–18 (2016).
    Article  Google Scholar 
    2.
    Dueñas, M. A. et al. The role played by invasive species in interactions with endangered and threatened species in the United States: a systematic review. Biodivers. Conserv. 27, 3171–3183 (2018).
    Article  Google Scholar 

    3.
    Dudgeon, D. et al. Freshwater biodiversity: Importance, threats, status and conservation challenges. Biol. Rev. Camb. Philos. Soc. 81, 163–182 (2006).
    PubMed  Article  PubMed Central  Google Scholar 

    4.
    Ricciardi, A. & MacIsaac, H. J. Impacts of biological invasions on freshwater ecosystems. Fifty Years Invas. Ecol. Legacy Charles Elton https://doi.org/10.1002/9781444329988.ch16 (2010).
    Article  Google Scholar 

    5.
    Moorhouse, T. P. & Macdonald, D. W. Are invasives worse in freshwater than terrestrial ecosystems?. Wiley Interdiscip. Rev. Water 2, 1–8 (2015).
    Article  Google Scholar 

    6.
    Rosewarne, P. J. et al. Feeding behaviour, predatory functional responses and trophic interactions of the invasive Chinese mitten crab (Eriocheir sinensis) and signal crayfish (Pacifastacus leniusculus). Freshw. Biol. 61, 426–443 (2016).
    Article  Google Scholar 

    7.
    Dick, J. T. A. et al. Advancing impact prediction and hypothesis testing in invasion ecology using a comparative functional response approach. Biol. Invasions 16, 735–753 (2014).
    Article  Google Scholar 

    8.
    Dick, J. T. A. et al. Invader Relative Impact Potential: a new metric to understand and predict the ecological impacts of existing, emerging and future invasive alien species. J. Appl. Ecol. 54, 1259–1267 (2017).
    Article  Google Scholar 

    9.
    Cuthbert, R. N., Dickey, J. W. E., Coughlan, N. E., Joyce, P. W. S. & Dick, J. T. A. The Functional Response Ratio (FRR): advancing comparative metrics for predicting the ecological impacts of invasive alien species. Biol. Invasions 21, 2543–2547 (2019).
    Article  Google Scholar 

    10.
    Devin, S., Piscart, C., Beisel, J. N. & Moreteau, J. C. Life History Traits of the Invader Dikerogammarus villosus (Crustacea: Amphipoda) in the Moselle River. France. Int. Rev. Hydrobiol. 89, 21–34 (2004).
    ADS  Article  Google Scholar 

    11.
    Nentwig, W., Bacher, S., Kumschick, S., Pyšek, P. & Vilà, M. More than “100 worst” alien species in Europe. Biol. Invasions 20, 1611–1621 (2018).
    Article  Google Scholar 

    12.
    Gallardo, B. & Aldridge, D. C. Is Great Britain heading for a Ponto-Caspian invasional meltdown?. J. Appl. Ecol. 52, 41–49 (2015).
    Article  Google Scholar 

    13.
    Kramer, A. M. et al. Suitability of Laurentian Great Lakes for invasive species based on global species distribution models and local habitat. Ecosphere 8, e01883 (2017).
    Article  Google Scholar 

    14.
    Van Riel, M. C. et al. Trophic relationships in the Rhine food web during invasion and after establishment of the Ponto-Caspian invader Dikerogammarus villosus. Hydrobiologia 565, 39–58 (2006).
    Article  Google Scholar 

    15.
    MacNeil, C., Boets, P., Lock, K. & Goethals, P. L. M. Potential effects of the invasive ‘killer shrimp’ (Dikerogammarus villosus) on macroinvertebrate assemblages and biomonitoring indices. Freshw. Biol. 58, 171–182 (2013).
    Article  Google Scholar 

    16.
    Dodd, J. A. et al. Predicting the ecological impacts of a new freshwater invader: Functional responses and prey selectivity of the ‘killer shrimp’, Dikerogammarus villosus, compared to the native Gammarus pulex. Freshw. Biol. 59, 337–352 (2014).
    Article  Google Scholar 

    17.
    Bruijs, M. C. M., Kelleher, B., Van Der Velde, G. & De Vaate, A. B. Oxygen consumption, temperature and salinity tolerance of the invasive amphipod Dikerogammarus villosus: Indicators of further dispersal via ballast water transport. Arch. fur Hydrobiol. 152, 633–646 (2001).
    Article  Google Scholar 

    18.
    Pöckl, M. Strategies of a successful new invader in European fresh waters: Fecundity and reproductive potential of the Ponto-Caspian amphipod Dikerogammarus villosus in the Austrian Danube, compared with the indigenous Gammarus fossarum and G. roeseli. Freshw. Biol. 52, 50–63 (2007).

    19.
    Rolla, M., Consuegra, S. & de Leaniz, C. G. Predator recognition and anti-predatory behaviour in a recent aquatic invader, the killer shrimp (Dikerogammarus villosus). Aquat. Invasions 15, 482–496 (2020).
    Article  Google Scholar 

    20.
    Kobak, J., Rachalewski, M. & Bącela-Spychalska, K. Conquerors or exiles? Impact of interference competition among invasive Ponto-Caspian gammarideans on their dispersal rates. Biol. Invasions 18, 1953–1965 (2016).
    Article  Google Scholar 

    21.
    Rewicz, T., Grabowski, M., MacNeil, C. & Bącela-Spychalska, K. The profile of a ‘perfect’ invader – the case of killer shrimp. Dikerogammarus villosus. Aquat. Invasions 9, 267–288 (2014).
    Article  Google Scholar 

    22.
    Hellmann, C. et al. The trophic function of Dikerogammarus villosus (Sowinsky, 1894) in invaded rivers: a case study in the Elbe and Rhine. Aquat. Invasions 10, 385–397 (2015).
    Article  Google Scholar 

    23.
    Platvoet, D., Van Der Velde, G., Dick, J. T. A. & Li, S. Flexible omnivory in Dikerogammarus villosus (Sowinsky, 1894) (Amphipoda) – Amphipod Pilot Species Project (AMPIS) Report 5. Crustaceana 82, 703–720 (2009).
    Article  Google Scholar 

    24.
    Taylor, N. G. & Dunn, A. M. Size matters: predation of fish eggs and larvae by native and invasive amphipods. Biol. Invasions 19, 89–107 (2017).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    25.
    Alford, R. A. Ecology: Bleak future for amphibians. Nature 480, 461–462 (2011).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    26.
    Alroy, J. Current extinction rates of reptiles and amphibians. Proc. Natl. Acad. Sci. https://doi.org/10.1073/pnas.1508681112 (2015).
    Article  PubMed  PubMed Central  Google Scholar 

    27.
    González-del-Pliego, P. et al. Phylogenetic and Trait-Based Prediction of Extinction Risk for Data-Deficient Amphibians. Curr. Biol. 29, 1557–1563.e3 (2019)

    28.
    Fisher, M. C. & Garner, T. W. J. Chytrid fungi and global amphibian declines. Nat. Rev. Microbiol. 18, 332–343 (2020).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    29.
    Hayes, T. B., Falso, P., Gallipeau, S. & Stice, M. The cause of global amphibian declines: a developmental endocrinologist’s perspective. J. Exp. Biol. 213, 921–933 (2010).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    30.
    Bellard, C., Genovesi, P. & Jeschke, J. M. Global patterns in threats to vertebrates by biological invasions. Proc. R. Soc. B Biol. Sci. https://doi.org/10.1098/rspb.2015.2454 (2016).
    Article  Google Scholar 

    31.
    IUCN. The IUCN Red List of Threatened Species. (2020).

    32.
    Nunes, A. L. et al. A global meta-analysis of the ecological impacts of alien species on native amphibians. Proc. R. Soc. B Biol. Sci. https://doi.org/10.1098/rspb.2018.2528 (2019).
    Article  Google Scholar 

    33.
    Ilhéu, M., Bernardo, J. & Fernandes, S. Biological invaders in inland waters: Profiles, distribution, and threats. Biol. invaders Inl. waters profiles, Distrib. Threat. 2, 543–558 (2007).

    34.
    Kats, L. B. & Ferrer, R. P. Alien predators and amphibian declines: Review of two decades of science and the transition to conservation. Divers. Distrib. 9, 99–110 (2003).
    Article  Google Scholar 

    35.
    Beebee, T. J. C. & Griffiths, R. A. The amphibian decline crisis: A watershed for conservation biology?. Biol. Conserv. 125, 271–285 (2005).
    Article  Google Scholar 

    36.
    National Biodiversity Network. NBN Atlas. Nbn (2017).

    37.
    Uehlinger, U., Wantzen, K. M., Leuven, R. S. E. W. & Arndt, H. The Rhine River Basin. in Rivers of Europe 199–245 (2009). https://doi.org/10.1016/B978-0-12-369449-2.00006-0

    38.
    Koester, M., Bayer, B. & Gergs, R. Is Dikerogammarus villosus (Crustacea, Gammaridae) a ‘killer shrimp’ in the River Rhine system?. Hydrobiologia 768, 299–313 (2016).
    Article  Google Scholar 

    39.
    Gergs, R. & Rothhaupt, K. O. Invasive species as driving factors for the structure of benthic communities in Lake Constance. Germany. Hydrobiologia 746, 245–254 (2014).
    Article  CAS  Google Scholar 

    40.
    Haubrock, P. J. et al. Shared histories of co-evolution may affect trophic interactions in a freshwater community dominated by alien species. Frontiers in Ecology and Evolution 7, 355 (2019).
    Article  Google Scholar 

    41.
    Marguillier, S. Stable isotope ratios and food web structure of aquatic ecosystems. (1998).

    42.
    Dick, J. T. A. & Platvoet, D. Invading predatory crustacean Dikerogammarus villosus eliminates both native and exotic species. Proc. R. Soc. B Biol. Sci. 267, 977–983 (2000).
    CAS  Article  Google Scholar 

    43.
    Bollache, L., Dick, J. T., Farnsworth, K. D. & Montgomery, W. I. Comparison of the functional responses of invasive and native amphipods. Biol Lett 4, 166–169 (2008).
    PubMed  Article  PubMed Central  Google Scholar 

    44.
    MacNeil, C. et al. The Ponto-Caspian ‘killer shrimp’, Dikerogammarus villosus (Sowinsky, 1894), invades the British Isles. Aquat. Invasions 5, 441–445 (2010).
    Article  Google Scholar 

    45.
    Worischka, S. et al. Food consumption of the invasive amphipod Dikerogammarus villosus in field mesocosms and its effects on leaf decomposition and periphyton. Aquat. Invasions 13, 261–275 (2018).
    Article  Google Scholar 

    46.
    Jourdan, J. et al. Pronounced species turnover, but no functional equivalence in leaf consumption of invasive amphipods in the river Rhine. Biol. Invasions 18, 763–774 (2016).
    Article  Google Scholar 

    47.
    Fries, G. & Der Tesch, F. W. Einfluss der Massenvorkommens von Gammarus tigrinus Sexton auf Fische und niedere Tierwelt in der Weser. Arch. für Fischer Wiss. 16, 133–150 (1965).
    Google Scholar 

    48.
    Hudgens, B. & Harbert, M. Amphipod Predation on Northern Red-Legged Frog (Rana Aurora) Embryos. Northwest. Nat. 100, 126 (2019).
    Article  Google Scholar 

    49.
    Räsänen, K., Pahkala, M., Laurila, A. & Merilä, J. Does Jelly Envelope Protect the Common Frog Rana Temporaria Embryos From Uv-B Radiation?. Herpetologica 59, 293–300 (2003).
    Article  Google Scholar 

    50.
    Ward, D. & Sexton, O. J. Anti-Predator Role of Salamander Egg Membranes. Copeia 1981, 724 (1981).
    Article  Google Scholar 

    51.
    Henrikson, B.-I. Predation on amphibian eggs and tadpoles by common predators in acidified lakes. Ecography (Cop.) 13, 201–206 (1990).
    Article  Google Scholar 

    52.
    Duellman, W. E. (William E. & Trueb, L. Biology of amphibians. (Johns Hopkins University Press, 1994).

    53.
    Latham, D., Jones, E. & Fasham, M. Amphibians. in Handbook of Biodiversity Methods: Survey, Evaluation and Monitoring (eds. Hill, D., Fasham, M., Tucker, G., Shewry, M. & Shaw, P.) (Cambridge University Press, 2005).

    54.
    Tinsley, R. C., Stott, L. C., Viney, M. E., Mable, B. K. & Tinsley, M. C. Extinction of an introduced warm-climate alien species, Xenopus laevis, by extreme weather events. Biol. Invasions 17, 3183–3195 (2015).
    PubMed  PubMed Central  Article  Google Scholar 

    55.
    Rall, B. C. et al. Universal temperature and body-mass scaling of feeding rates. Philos. Trans. R. Soc. B Biol. Sci. 367, 2923–2934 (2012).

    56.
    Glazier, D. S. A unifying explanation for diverse metabolic scaling in animals and plants. Biol. Rev. 85, 111–138 (2010).
    PubMed  Article  PubMed Central  Google Scholar 

    57.
    Brown, J. H., Gillooly, J. F., Allen, A. P., Savage, V. M. & West, G. B. Toward a metabolic theory of ecology. Ecology 85, 1771–1789 (2004).
    Article  Google Scholar 

    58.
    Mayer, G., Waloszek, D., Maier, G. & Maas, A. Mouthparts of the Ponto-Caspian Invader Dikerogammarus Villosus (Amphipoda: Pontogammaridae). J. Crustac. Biol. 28, 1–15 (2008).
    Article  Google Scholar 

    59.
    Vucic-Pestic, O., Rall, B. C., Kalinkat, G. & Brose, U. Allometric functional response model: Body masses constrain interaction strengths. J. Anim. Ecol. 79, 249–256 (2010).
    PubMed  Article  PubMed Central  Google Scholar 

    60.
    Maazouzi, C., Piscart, C., Legier, F. & Hervant, F. Ecophysiological responses to temperature of the ‘killer shrimp’ Dikerogammarus villosus: Is the invader really stronger than the native Gammarus pulex? Comp. Biochem. Physiol. – A Mol. Integr. Physiol. 159, 268–274 (2011).

    61.
    Álvarez, D. & Nicieza, A. G. Differential success of prey escaping predators: tadpole vulnerability or predator selection??. Copeia 2009, 453–457 (2009).
    Article  Google Scholar 

    62.
    Ward, A. & Webster, M. Sociality. in Sociality: The Behaviour of Group-Living Animals 1–8 (Springer International Publishing, 2016).https://doi.org/10.1007/978-3-319-28585-6_1

    63.
    Price, P. W., Denno, R. F., Eubanks, M. D., Finke, D. L. & Kaplan, I. Insect Ecology: Behaviour, Populations and Communities. (Cambridge University Press, 2011).

    64.
    Juliano, S. A. Nonlinear Curve Fitting: Predation and Functional Response Curves. in Design and Analysis of Ecological Experiments (eds. Cheiner, S. M. & Gurven, J.) 178–196 (Chapman and Hall, 2001).

    65.
    Barrios-O’Neill, D. et al. Fortune favours the bold: A higher predator reduces the impact of a native but not an invasive intermediate predator. J. Anim. Ecol. 83, 693–701 (2014).

    66.
    Sentis, A. & Boukal, D. S. On the use of functional responses to quantify emergent multiple predator effects. Sci. Rep. 8, (2018).

    67.
    Médoc, V., Albert, H. & Spataro, T. Functional response comparisons among freshwater amphipods: ratio-dependence and higher predation for Gammarus pulex compared to the non-natives Dikerogammarus villosus and Echinogammarus berilloni. Biol. Invasions 17, 3625–3637 (2015).
    Article  Google Scholar 

    68.
    Laverty, C., Nentwig, W., Dick, J. & Lucy, F. Alien aquatics in Europe: assessing the relative environmental and socio-economic impacts of invasive aquatic macroinvertebrates and other taxa. Manag. Biol. Invasions 6, 341–350 (2015).
    Article  Google Scholar 

    69.
    Dickey, J. W. E. et al. On the RIP: using Relative Impact Potential to assess the ecological impacts of invasive alien species. NeoBiota 55, 27–60 (2020).
    Article  Google Scholar 

    70.
    Gallardo, B., Errea, M. P. & Aldridge, D. C. Application of bioclimatic models coupled with network analysis for risk assessment of the killer shrimp, Dikerogammarus villosus. Great Britain. Biol. Invasions 14, 1265–1278 (2012).
    Article  Google Scholar 

    71.
    Gallardo, B. & Aldridge, D. C. Priority setting for invasive species management by the water industry. Water Res. 178, 115771 (2020).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    72.
    Gosner, K. L. A simplified table for staging anuran embryos larvae. Herpetodologists’ Leag. 16, 183–190 (1960).
    Google Scholar 

    73.
    Currie, S. P., Combes, D., Scott, N. W., Simmers, J. & Sillar, K. T. A behaviorally related developmental switch in nitrergic modulation of locomotor rhythmogenesis in larval Xenopus tadpoles. J. Neurophysiol. 115, 1446–1457 (2016).
    PubMed  PubMed Central  Article  Google Scholar 

    74.
    Müller, J. C., Schramm, S. & Seitz, A. Genetic and morphological differentiation of Dikerogammarus invaders and their invasion history in Central Europe. Freshw. Biol. 47, 2039–2048 (2002).
    Article  Google Scholar 

    75.
    Blackman, R. C. et al. Detection of a new non-native freshwater species by DNA metabarcoding of environmental samples – first record of gammarus fossarum in the UK. Aquat. Invasions 12, 177–189 (2017).
    Article  Google Scholar 

    76.
    van der Velde, G. et al. Environmental and morphological factors influencing predatory behaviour by invasive non-indigenous gammaridean species. Biol. Invasions 11, 2043–2054 (2009).
    Article  Google Scholar 

    77.
    Dick, J. T. A. et al. Parasitism may enhance rather than reduce the predatory impact of an invader. Biol. Lett. 6, 636–638 (2010).
    PubMed  PubMed Central  Article  Google Scholar 

    78.
    Iltis, C., Spataro, T., Wattier, R. & Médoc, V. Parasitism may alter functional response comparisons: a case study on the killer shrimp Dikerogammarus villosus and two non-invasive gammarids. Biol. Invasions 20, (2018).

    79.
    Welton, J. S. Life-history and production of the amphipod Gammarus pulex in a Dorset chalk stream. Freshw. Biol. 9, 263–275 (1979).
    Article  Google Scholar 

    80.
    Oertli, B. Leaf litter processing and energy flow through macroinvertebrates in a woodland pond (Switzerland). Oecologia 96, 466–477 (1993).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    81.
    Lods-Crozet, B. & Reymond, O. Bathymetric expansion of an invasive gammarid (Dikerogammarus villosus, Crustacea, Amphipoda) in Lake Léman. J. Limnol. 65, 141–144 (2006).
    Article  Google Scholar 

    82.
    Harkness, J. B. The relationships between stressors, macroinvertebrate community structure and leaf processing in stream ecosystems. (University of Sheffield, 2008).

    83.
    Leberfinger, K. & Herrmann, J. Secondary production of invertebrate shredders in open-canopy, intermittent streams on the island of land, southeastern Sweden. J. North Am. Benthol. Soc. 29, 934–944 (2010).
    Article  Google Scholar 

    84.
    Lods-Crozet, B. Long-term biomonitoring of invertebrate neozoans in Lake Geneva. Arch. des Sci. 67, 101–108 (2014).
    Google Scholar 

    85.
    Johns, T., Smith, D. C., Homann, S. & England, J. A. Time-series analysis of a native and a non-native amphipod shrimp in two English rivers. BioInvasions Rec. 7, 101–110 (2018).
    Article  Google Scholar 

    86.
    Clinton, K. E., Mathers, K. L., Constable, D., Gerrard, C. & Wood, P. J. Substrate preferences of coexisting invasive amphipods, Dikerogammarus villosus and Dikerogammarus haemobaphes, under field and laboratory conditions. Biol. Invasions 20, 2187–2196 (2018).
    Article  Google Scholar 

    87.
    Haas, G., Brunke, M. & Streit, B. Fast Turnover in Dominance of Exotic Species in the Rhine River Determines Biodiversity and Ecosystem Function: An Affair Between Amphipods and Mussels. in Invasive Aquatic Species of Europe. Distribution, Impacts and Management 426–432 (2002). doi:https://doi.org/10.1007/978-94-015-9956-6_42

    88.
    Krisp, H. & Maier, G. Consumption of macroinvertebrates by invasive and native gammarids: A comparison. J. Limnol. 64, 55–59 (2005).
    Article  Google Scholar 

    89.
    Mulattieri, P. Etude de l’impact des aménagements riverains sur les macroinvertébrés benthiques des rives genevoises du Léman. (Université de Genève, 2006).

    90.
    Platvoet, D., Dick, J. T. A., MacNeil, C., van Riel, M. C. & van der Velde, G. Invader-invader interactions in relation to environmental heterogeneity leads to zonation of two invasive amphipods, dikerogammarus villosus (sowinsky) and gammarus tigrinus sexton: Amphipod pilot species project (ampis) report 6. Biol. Invasions 11, 2085–2093 (2009).
    Article  Google Scholar 

    91.
    Tricarico, E. et al. The killer shrimp, Dikerogammarus villosus (Sowinsky, 1894), is spreading in Italy. Aquat. Invasions 5, 211–214 (2010).
    Article  Google Scholar 

    92.
    Muskó, I. B., Balogh, C., Tóth, Á. P., Varga, É. & Lakatos, G. Differential response of invasive malacostracan species to lake level fluctuations. Hydrobiologia 590, 65–74 (2007).
    Article  Google Scholar 

    93.
    Hellmann, C., Schöll, F., Worischka, S., Becker, J. & Winkelmann, C. River-specific effects of the invasive amphipod Dikerogammarus villosus (Crustacea: Amphipoda) on benthic communities. Biol. Invasions 19, 381–398 (2017).
    Article  Google Scholar 

    94.
    GBIF.org. Global Biodiversity Information Facility. Choice Reviews Online 41, 41–5289–41–5289 (2004).

    95.
    INaturalist.org. iNaturalist. (2020). Available at: https://www.inaturalist.org/. (Accessed: 16th October 2020)

    96.
    R Core Team. R: A Language and Environment for Statistical Computing. (2018).

    97.
    Pritchard, D. W., Paterson, R. A., Bovy, H. C. & Barrios-O’Neill, D. frair: an R package for fitting and comparing consumer functional responses. Methods Ecol. Evol. 8, 1528–1534 (2017).

    98.
    Rogers, D. Random Search and Insect Population Models. J. Anim. Ecol. 41, 369 (1972).
    Article  Google Scholar 

    99.
    Bolker, B. & R Core Team. bbmle: Tools for General Maximum Likelihood Estimation. R package version 1.0.20. (2017).

    100.
    Laverty, C. et al. Assessing the ecological impacts of invasive species based on their functional responses and abundances. Biol. Invasions 19, 1653–1665 (2017).
    Article  Google Scholar 

    101.
    Cuthbert, R. N., Dick, J. T. A., Callaghan, A. & Dickey, J. W. E. Biological control agent selection under environmental change using functional responses, abundances and fecundities; the Relative Control Potential (RCP) metric. Biol. Control 121, 50–57 (2018).
    Article  Google Scholar 

    102.
    Hothorn, T., Bretz, F. & Westfall, P. Simultaneous inference in general parametric models. Biometrical Journal 50, 346–363 (2008).
    MathSciNet  PubMed  MATH  Article  PubMed Central  Google Scholar  More