More stories

  • in

    Nasal microbiome disruption and recovery after mupirocin treatment in Staphylococcus aureus carriers and noncarriers

    Study population and study designThis is a prospective interventional cohort study of healthy S. aureus carriers and noncarriers in the Netherlands. All experiments were performed in accordance with the Dutch Medical Research Involving Human Subjects Act (WMO). The study protocol was approved by the local Medical Ethical Committee of the Erasmus University Medical Centre Rotterdam, The Netherlands (MEC-2018-091). Written informed consent was obtained for all participants. Participants were recruited through advertisements at Dutch universities and the research teams social networks. Exclusion criteria were age  8 CFU/mL for each culture. Noncarriers were defined as 2 S. aureus-negative cultures. Intermittent S. aureus carriers were excluded from further participation in the study. Eligible volunteers were enrolled on a first-come, first-served basis.Eligible participants were asked to fill out a questionnaire regarding risk factors for S. aureus acquisition. All participants received decolonization treatment. Decolonization consisted of mupirocin nasal ointment (2%, GlaxoSmithKline BV, Zeist, the Netherlands) twice daily and chlorhexidine gluconate cutaneous solution (4%w/v, Regent Medical Overseas Limited, Oldham, UK) once daily, both for 5 days.Nasal samples were taken 1 day before decolonization (D0) and 2 days (D7), 1 month (M1), 3 months (M3) and 6 months (M6) after decolonization. All participants received a personal demonstration for nasal sampling by the executive researcher. Thereafter, all specimens were taken by the participants by inserting a swab (ESwab, 490CE.A, Copan Italia, Brescia, Italy) into one nostril and rotating 5 times, repeating this in the second nostril using the same swab. Swabs were collected in a container filled with 1 ml modified Liquid Amies, a collection and transport solution, and sent through regular mail service (non-temperature controlled) or deposited at the laboratory personally.
    Staphylococcus aureus quantitative cultureQuantitative S. aureus cultures were conducted to examine the dynamics of S. aureus carriage over the 6-month follow-up period after decolonization. Swab containers were vortexed for 20 s before plating. Serial dilutions of Amies medium were plated onto phenol mannitol salt agar (PHMA) and incubated for 2 days at 37 °C. Swabs were placed in phenol mannitol salt broth (PHMB) and incubated for 7 days at 37 °C for enrichment. S. aureus growth was confirmed by a latex agglutination test (Staph Plus Latex Kit, Diamondial, Vienna, Austria). Morphologically different S. aureus colonies were selected for spa typing and methicillin resistance screening using BBL CHROMagar MRSA II agar (BD, Breda, The Netherlands).
    Spa typingMolecular typing of S. aureus isolates was performed to infer whether recolonization with S. aureus in decolonized carriers involved the same spa-type. Typing was limited to the last S. aureus positive culture moment and the last S. aureus positive culture moment after decolonization in recolonised carriers. S. aureus DNA lysates were prepared by boiling in 10 mM Tris–HCl, 1 mM disodium EDTA, pH 8.0 or extraction with the QIAamp DNA Mini Kit (QIAGEN, Venlo, The Netherlands) according to the manufacturer’s instructions. Amplification of the S. aureus protein A (spa) repeat region was performed by PCR with 2 sets of primers. One set consisted of forward primer spa-1113, 5′-TAAAGACGATCCTTCGGTGAGC-3′ and reverse primer spa-1514, 5′-CAGCAGTAGTGCCGTTTGCTT-3′24. The other set consisted of forward primers spa-F1, 5′-AACAACGTAACGGCTTCATCC-3′ and spa-F2 5′-AGACGATCCTTCAGTGAGC-3′ and reverse primer spa-R1 5′-GCTTTTGCAATGTCATTTACTG-3′. Amplicons were purified with ExoSAP-IT (Applied Biosystems) according to the manufacturer’s instructions and sent for sequence analysis (Baseclear, Leiden, The Netherlands). Resulting sequences were analysed using BioNumerics v7.6 (Applied Maths NV, Sint-Martens-Latem, Belgium) and the spa types were assigned by use of the RidomStaphType database (Ridom GmbH, Würzburg, Germany).16S ribosomal RNA sequencing of nasal microbiotaThe impact of decolonization on the nasal microbiome and the recovery of the microbiome structure after decolonization were examined by means of 16S rRNA metabarcoding. Amies medium from each nasal swab container was stored at − 80 °C on the day of receipt at the study laboratory in Rotterdam, NL, then sent at − 80 °C to the microbiome analysis laboratory in Lyon, FR. To properly capture the impact of decolonization on the living microbiota, metabarcoding used RNA-based 16S ribosomal RNA (rRNA, which is preserved in living cells but quickly cleared after cell death or lysis) rather than the DNA coding sequence, as DNA can persist for prolonged time periods after cell death25,26,27,28. RNA was extracted using the Mag Bind® Total RNA 96 Kit (Omega Bio-tek) tissue protocol from 150 µL of samples’ material. Cell lysis was performed using beads (Disruptor plate C plus—Omega Bio-tek) and proteinase K for 15 min at 2600 rpm, followed by 10 min at room temperature without agitation, and finished with a DNase I digestion of 20 min at room temperature. RNA was quantified using QuantiFluor RNA kit on Tecan Safire (TECAN). 10 ng total RNA was used for reverse transcription using FIREScript RT cDNA synthesis kit (Solis Biodyne) with random primers, then cDNA was purified with SPRIselect reagent (Beckman coulter) and quantified.The rRNA V1–V3 region was PCR amplified using the 5× HOT BIOAmp® BlendMaster Mix 12,5 mM MgCl 2 (Biofidal), 10× GC rich Enhancer (Biofidal) and BSA 20 mg/mL. The PCR reaction consisted of 30 cycles at 56 °C using the forward primer 27F, 5′-TCGTCGGCAGCGTCAGATGTGTATAAGAGACAG AGAGTTTGATCCTGGCTCAG-3′ and reverse primer 534R, 5′-GTCTCGTGGGCTCGGAGATGTGTATAAGAGACAGATTACCGCGGCTGCTGG-3′ in 25 µL of solution. PCR products were purified using SPRIselect beads (Beckman Coulter) in 20 µL nuclease-free water and quantified using QuantiFluor dsDNA (Promega). Samples were indexed with lllumina’s barcodes with the same PCR reagents during a 12 cycles PCR, then purified and quantified as previously mentioned. Samples were normalized and pooled, then sequenced using Illumina MiSeq V3 Flow Cell following the constructor’s recommendations for a 2 × 300 bp paired-end application. A mean of 130 k proofread reads per sample was obtained.Experiment buffers were used as negative controls to detect contamination by out-of-sample bacterial RNA. RNA extraction was controlled using an in-house mix of live Staphylococcus aureus ATCC29213 and Escherichia coli ATCC25922 in equal proportions, allowing for assessing extraction bias in Gram-positive and -negative bacteria. PCR amplification bias was controlled using a commercial DNA mix of 8 bacterial species (ZymoBIOMICS™ Microbial Community DNA Standard).Bioinformatics and statistical analysesSequencing reads were quality checked and trimmed. Paired-ended read pairs were merged using BBMap version 38.49 (available at https://sourceforge.net/projects/bbmap/), with default options besides a minimum single size of 150 bp with an average Phred quality score higher than 10, and a total pair size of minimum 400 bp. PCR adapters were removed with cutadapt v.2.1 (Martin 2011) then dereplicated using vsearch v.2.12.029 with the sizeout option. For species assignment, reads were aligned to sequences of NCBI blast 16S_ribosomal_RNA database (version date 03.12.2020) using Blastn v.2.11.0+30,31, keeping a maximum of 20 reference targets. Read counts per bacterial species were normalized to account for taxon-specific variations of the copy number of 16S rRNA genes using NCBI rrnDB-5.5 database based on the mean gene copy number in the taxon.To optimize the resolution of sequencing read taxonomic assignment, we used in-house bioinformatic software publicly available at https://github.com/rasigadelab/taxonresolve. Briefly, when a read matches sequences from several species with identical alignment scores, taxonomic assignment pipelines typically output the higher taxonomic level such as the genus (e.g., Staphylococcus spp. when a read matches S. aureus and S. epidermidis). This loss of information can be problematic when species-level discrimination is important. To prevent losing species-level information, the taxonresolve software assigns reads with uncertain species to groups of species rather than to genera.Bacterial species deemed present from contaminating sources such as kits reagents and found in negative controls, mostly from the Bacillus genera, were removed. A total of 1376 species or group of species were retained. The rarefaction curves corresponding to the sequencing effort to assess the species richness within samples are shown in Supplementary Fig. 3. Most samples reached a plateau after 40,000 sequences.Given the small sample size compared to the number of variables and species considered in this study, no hypothesis testing was performed, and we provide a descriptive assessment of the results. In figures, 95% confidence intervals of the means were computed based on normal approximation, after log transformation for CFU/mL and log odds transformation for quantities restricted to the [0, 1] interval, such as proportions.In microbial diversity analyses, we retained the 9 most prevalent bacterial species and pooled the other species into an ‘Others’ category. To assess the disruption and possible recovery of the microbiota, the divergence of sampled microbiota relative to the initial, pre-treatment microbiota (D0) was assessed using the Bray–Curtis dissimilarity at each sampling time point relative to the first sample of the same patient.Software code of the analyses are available at https://github.com/rasigadelab/macotra-metabarcoding. Data are available at https://zenodo.org/record/6382657. Analyses and figures used R software v3.6.032 with packages dplyr33, ggplot234, vegan35, and MicrobiomAnalyst available at https://www.microbiomeanalyst.ca36,37. More

  • in

    The role of individual variability on the predictive performance of machine learning applied to large bio-logging datasets

    Tsai, C.-W., Lai, C.-F., Chao, H.-C. & Vasilakos, A. V. Big data analytics: a survey. J. Big Data 2, 21 (2015).
    Google Scholar 
    Lemoine, F. et al. Renewing Felsenstein’s phylogenetic bootstrap in the era of big data. Nature 556, 452–456 (2018).ADS 
    CAS 

    Google Scholar 
    Manzoni, C. et al. Genome, transcriptome and proteome: the rise of omics data and their integration in biomedical sciences. Brief. Bioinform. 19, 286–302 (2018).CAS 

    Google Scholar 
    Lichtman, J. W., Pfister, H. & Shavit, N. The big data challenges of connectomics. Nat. Neurosci. 17, 1448–1454 (2014).CAS 

    Google Scholar 
    Altaf-Ul-Amin, M., Afendi, F. M., Kiboi, S. K. & Kanaya, S. Systems biology in the context of big data and networks. Biomed. Res. Int. 2014, 428570 (2014).
    Google Scholar 
    Xia, J., Wang, J. & Niu, S. Research challenges and opportunities for using big data in global change biology. Glob. Change Biol. 26, 6040–6061 (2020).ADS 

    Google Scholar 
    Hindell, M. A. et al. Tracking of marine predators to protect Southern Ocean ecosystems. Nature 580, 87–92 (2020).ADS 
    CAS 

    Google Scholar 
    Hussey, N. E. et al. Ecology. Aquatic animal telemetry: A panoramic window into the underwater world. Science 348, 1255642 (2015).
    Google Scholar 
    Kays, R., Crofoot, M. C., Jetz, W. & Wikelski, M. Terrestrial animal tracking as an eye on life and planet. Science 348, aaa2478 (2015).
    Google Scholar 
    Sherub, S., Fiedler, W., Duriez, O. & Wikelski, M. Bio-logging, new technologies to study conservation physiology on the move: a case study on annual survival of Himalayan vultures. J. Comp. Physiol. A Neuroethol. Sens. Neural Behav. Physiol. 203, 531–542 (2017).
    Google Scholar 
    Nathan, R. et al. Big-data approaches lead to an increased understanding of the ecology of animal movement. Science 375, eabg1780 (2022).CAS 

    Google Scholar 
    Wilson, R. P. et al. Estimates for energy expenditure in free-living animals using acceleration proxies: A reappraisal. J. Anim. Ecol. 89, 161–172 (2020).
    Google Scholar 
    Patterson, A., Gilchrist, H. G., Chivers, L., Hatch, S. & Elliott, K. A comparison of techniques for classifying behavior from accelerometers for two species of seabird. Ecol. Evol. 9, 3030–3045 (2019).
    Google Scholar 
    Masello, J. F. et al. How animals distribute themselves in space: energy landscapes of Antarctic avian predators. Mov. Ecol. 9, 24 (2021).
    Google Scholar 
    Shepard, E. L. C. et al. Energy landscapes shape animal movement ecology. Am. Nat. 182, 298–312 (2013).
    Google Scholar 
    Elliott, K. H., Le Vaillant, M., Kato, A., Speakman, J. R. & Ropert-Coudert, Y. Accelerometry predicts daily energy expenditure in a bird with high activity levels. Biol. Lett. 9, 20120919 (2013).
    Google Scholar 
    Nickel, B. A., Suraci, J. P., Nisi, A. C. & Wilmers, C. C. Energetics and fear of humans constrain the spatial ecology of pumas. Proc. Natl. Acad. Sci. USA 118, e2004592118 (2021).
    Eisaguirre, J. M., Booms, T. L., Barger, C. P., Lewis, S. B. & Breed, G. A. Novel step selection analyses on energy landscapes reveal how linear features alter migrations of soaring birds. J. Anim. Ecol. 89, 2567–2583 (2020).
    Google Scholar 
    Wittemyer, G., Northrup, J. M. & Bastille-Rousseau, G. Behavioural valuation of landscapes using movement data. Philos. Trans. R. Soc. Lond. B Biol. Sci. 374, 20180046 (2019).
    Google Scholar 
    Chimienti, M. et al. The use of an unsupervised learning approach for characterizing latent behaviors in accelerometer data. Ecol. Evol. 6, 727–741 (2016).
    Google Scholar 
    Hounslow, J. L. et al. Assessing the effects of sampling frequency on behavioural classification of accelerometer data. J. Exp. Mar. Bio. Ecol. 512, 22–30 (2019).
    Google Scholar 
    Glass, T. W., Breed, G. A., Robards, M. D., Williams, C. T. & Kielland, K. Accounting for unknown behaviors of free-living animals in accelerometer-based classification models: Demonstration on a wide-ranging mesopredator. Ecol. Inf. 60, 101152 (2020).
    Google Scholar 
    Wang, Y. et al. Movement, resting, and attack behaviors of wild pumas are revealed by tri-axial accelerometer measurements. Mov. Ecol. 3, 2 (2015).
    Google Scholar 
    Chakravarty, P., Cozzi, G., Ozgul, A. & Aminian, K. A novel biomechanical approach for animal behaviour recognition using accelerometers. Methods Ecol. Evol. https://doi.org/10.1111/2041-210X.13172 (2019).Article 

    Google Scholar 
    Clarke, T. M. et al. Using tri-axial accelerometer loggers to identify spawning behaviours of large pelagic fish. Mov. Ecol. 9, 26 (2021).
    Google Scholar 
    Zhang, J., O’Reilly, K. M., Perry, G. L. W., Taylor, G. A. & Dennis, T. E. Extending the functionality of behavioural change-point analysis with k-Means clustering: a case study with the little penguin (Eudyptula minor). PLoS ONE 10, e0122811 (2015).
    Google Scholar 
    Korpela, J. et al. Machine learning enables improved runtime and precision for bio-loggers on seabirds. Commun. Biol. 3, 633 (2020).
    Google Scholar 
    Jeantet, L. et al. Behavioural inference from signal processing using animal-borne multi-sensor loggers: a novel solution to extend the knowledge of sea turtle ecology. R. Soc. Open Sci. 7, 200139 (2020).ADS 

    Google Scholar 
    Wang, G. Machine learning for inferring animal behavior from location and movement data. Ecol. Inf. 49, 69–76 (2019).
    Google Scholar 
    Dunford, C. E. et al. Surviving in steep terrain: a lab-to-field assessment of locomotor costs for wild mountain lions (Puma concolor). Mov. Ecol. 8, 34 (2020).
    Google Scholar 
    Jeanniard-du-Dot, T., Guinet, C., Arnould, J. P. Y., Speakman, J. R. & Trites, A. W. Accelerometers can measure total and activity-specific energy expenditures in free-ranging marine mammals only if linked to time-activity budgets. Funct. Ecol. 31, 377–386 (2017).
    Google Scholar 
    Hicks, O. et al. Acceleration predicts energy expenditure in a fat, flightless, diving bird. Sci. Rep. 10, 21493 (2020).ADS 
    CAS 

    Google Scholar 
    Dentinger, J. E. et al. A probabilistic framework for behavioral identification from animal-borne accelerometers. Ecol. Model. 464, 109818 (2022).
    Google Scholar 
    Chakravarty, P., Maalberg, M., Cozzi, G., Ozgul, A. & Aminian, K. Behavioural compass: animal behaviour recognition using magnetometers. Mov. Ecol. 7, 28 (2019).
    Google Scholar 
    Hammond, T. T., Palme, R. & Lacey, E. A. Ecological specialization, variability in activity patterns and response to environmental change. Biol. Lett. 14, 20180115 (2018).
    Google Scholar 
    Lynch, H. J. & LaRue, M. A. First global census of the Adélie Penguin. Auk 131, 457–466 (2014).
    Google Scholar 
    Riaz, J., Bestley, S., Wotherspoon, S., Freyer, J. & Emmerson, L. From trips to bouts to dives: temporal patterns in the diving behaviour of chick-rearing Adélie penguins East Antarctica. Mar. Ecol. Prog. Ser. 654, 177–194 (2020).ADS 

    Google Scholar 
    Cherel, Y. Isotopic niches of emperor and Adélie penguins in Adélie Land, Antarctica. Mar. Biol. 154, 813–821 (2008).
    Google Scholar 
    Little Penguin (Eudyptula minor) – BirdLife species factsheet. at Carroll, G., Harcourt, R., Pitcher, B. J., Slip, D. & Jonsen, I. Recent prey capture experience and dynamic habitat quality mediate short-term foraging site fidelity in a seabird. Proc. Biol. Sci. 285, 20180788 (2018).
    Google Scholar 
    Meyer, X. et al. Oceanic thermal structure mediates dive sequences in a foraging seabird. Ecol. Evol. 10, 6610–6622 (2020).
    Google Scholar 
    Cavallo, C. et al. Quantifying prey availability using the foraging plasticity of a marine predator, the little penguin. Funct. Ecol. https://doi.org/10.1111/1365-2435.13605 (2020).Article 

    Google Scholar 
    Ropert-Coudert, Y., Chiaradia, A. & Kato, A. An exceptionally deep dive by a Little Penguin Eudyptula minor. Mar. Ornithol 34, 71–74 (2006).
    Google Scholar 
    Ropert-Coudert, Y., Kato, A., Wilson, R. P. & Cannell, B. Foraging strategies and prey encounter rate of free-ranging Little Penguins. Mar. Biol. 149, 139–148 (2006).
    Google Scholar 
    Rodríguez, A., Chiaradia, A., Wasiak, P., Renwick, L. & Dann, P. Waddling on the dark side: ambient light affects attendance behavior of little penguins. J. Biol. Rhythms 31, 194–204 (2016).
    Google Scholar 
    Ropert-Coudert, Y. et al. Happy feet in a hostile world? the future of penguins depends on proactive management of current and expected threats. Front. Mar. Sci. 6, 248 (2019).
    Google Scholar 
    Shuert, C. R., Pomeroy, P. P. & Twiss, S. D. Assessing the utility and limitations of accelerometers and machine learning approaches in classifying behaviour during lactation in a phocid seal. Anim. Biotelemetry 6, 14 (2018).
    Google Scholar 
    Dickinson, E. R. et al. Limitations of using surrogates for behaviour classification of accelerometer data: refining methods using random forest models in Caprids. Mov. Ecol. 9, 28 (2021).
    Google Scholar 
    Conway, A. M., Durbach, I. N., McInnes, A. & Harris, R. N. Frame-by-frame annotation of video recordings using deep neural networks. Ecosphere 12, e03384 (2021).
    Google Scholar 
    Ravindran, S. Five ways deep learning has transformed image analysis. Nature 609, 864–866 (2022).ADS 
    CAS 

    Google Scholar 
    Del Caño, M. et al. Fine-scale body and head movements allow to determine prey capture events in the Magellanic Penguin (Spheniscus magellanicus). Mar. Biol. 168, 84 (2021).
    Google Scholar 
    Johnson, J. M. & Khoshgoftaar, T. M. Survey on deep learning with class imbalance. J. Big Data 6, 27 (2019).
    Google Scholar 
    Hazen, E. L. et al. Marine top predators as climate and ecosystem sentinels. Front. Ecol. Environ. 17, 565–574 (2019).
    Google Scholar 
    Sánchez, S. et al. Within-colony spatial segregation leads to foraging behaviour variation in a seabird. Mar. Ecol. Prog. Ser. 606, 215–230 (2018).ADS 

    Google Scholar 
    Patrick, S. C., Martin, J. G. A., Ummenhofer, C. C., Corbeau, A. & Weimerskirch, H. Albatrosses respond adaptively to climate variability by changing variance in a foraging trait. Glob. Change Biol. https://doi.org/10.1111/gcb.15735 (2021).Article 

    Google Scholar 
    Bonar, M. et al. Geometry of the ideal free distribution: individual behavioural variation and annual reproductive success in aggregations of a social ungulate. Ecol. Lett. 23, 1360–1369 (2020).
    Google Scholar 
    Michelot, C., Kato, A., Raclot, T. & Ropert-Coudert, Y. Adélie penguins foraging consistency and site fidelity are conditioned by breeding status and environmental conditions. PLoS ONE 16, e0244298 (2021).CAS 

    Google Scholar 
    Mahoney, P. J. et al. Navigating snowscapes: scale-dependent responses of mountain sheep to snowpack properties. Ecol. Appl. 28, 1715–1729 (2018).
    Google Scholar 
    Watanabe, Y. Y., Ito, K., Kokubun, N. & Takahashi, A. Foraging behavior links sea ice to breeding success in Antarctic penguins. Sci. Adv. 6, eaba4828 (2020).ADS 

    Google Scholar 
    Lescroël, A. et al. Working less to gain more: when breeding quality relates to foraging efficiency. Ecology 91, 2044–2055 (2010).
    Google Scholar 
    Zimmer, I., Ropert-Coudert, Y., Kato, A., Ancel, A. & Chiaradia, A. Does foraging performance change with age in female little penguins (Eudyptula minor)?. PLoS ONE 6, e16098 (2011).ADS 
    CAS 

    Google Scholar 
    Hertel, A. G., Royauté, R., Zedrosser, A. & Mueller, T. Biologging reveals individual variation in behavioural predictability in the wild. J. Anim. Ecol. 90, 723–737 (2021).
    Google Scholar 
    Dickinson, E. R., Stephens, P. A., Marks, N. J., Wilson, R. P. & Scantlebury, D. M. Best practice for collar deployment of tri-axial accelerometers on a terrestrial quadruped to provide accurate measurement of body acceleration. Anim. Biotelemetry 8, 9 (2020).
    Google Scholar 
    Garde, B. et al. Ecological inference using data from accelerometers needs careful protocols. Methods Ecol. Evol. https://doi.org/10.1111/2041-210X.13804 (2022).Article 

    Google Scholar 
    Watanabe, Y. Y., Ito, M. & Takahashi, A. Testing optimal foraging theory in a penguin-krill system. Proc. Biol. Sci. 281, 20132376 (2014).
    Google Scholar 
    Grémillet, D. et al. Energetic fitness: Field metabolic rates assessed via 3D accelerometry complement conventional fitness metrics. Funct. Ecol. 32, 1203–1213 (2018).
    Google Scholar 
    Chimienti, M. et al. Quantifying behavior and life-history events of an Arctic ungulate from year-long continuous accelerometer data. Ecosphere 12, e03565 (2021).
    Google Scholar 
    Sutton, G. J., Botha, J. A., Speakman, J. R. & Arnould, J. P. Y. Validating accelerometry-derived proxies of energy expenditure using the doubly-labelled water method in the smallest penguin species. Biol. Open 10, bio055475 (2021).
    Google Scholar 
    Pagano, A. M. & Williams, T. M. Estimating the energy expenditure of free-ranging polar bears using tri-axial accelerometers: A validation with doubly labeled water. Ecol. Evol. 9, 4210–4219 (2019).
    Google Scholar 
    Ballance, L. T., Ainley, D. G., Ballard, G. & Barton, K. An energetic correlate between colony size and foraging effort in seabirds, an example of the Adélie penguin Pygoscelis adeliae. J. Avian Biol. 40, 279–288 (2009).
    Google Scholar 
    Wilson, R. P. et al. Long-term attachment of transmitting and recording devices to penguins and other seabirds. Wildl. Soc. Bull. 25, 101–106 (1997).
    Google Scholar 
    Shepard, E. L. C. et al. Identification of animal movement patterns using tri-axial accelerometry. Endanger. Species Res. 10, 47–60 (2008).ADS 

    Google Scholar 
    Kato, A., Ropert-Coudert, Y., Grémillet, D. & Cannell, B. Locomotion and foraging strategy in foot-propelled and wing-propelled shallow-diving seabirds. Mar. Ecol. Prog. Ser. 308, 293–301 (2006).ADS 

    Google Scholar 
    Team, R. C. R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna, Austria. URL https://url.org/www.R-project.org/ (2021).
    Ainley, D. The Adélie Penguin: Bellwether of Climate Change (New York: Columbia University Press) (2006).Langrognet, F. et al. Rmixmod: Classification with Mixture Modelling. (2020).Bishop, C. M. Pattern Recognition and Machine Learning. Springer Science+Business Media, LLC, New
    York, NY. (2006).Amélineau, F. et al. Intra- and inter-individual changes in little penguin diving and isotopic composition over the breeding season. Mar. Biol. 168, 62 (2021).
    Google Scholar 
    Stoffel, M. A., Nakagawa, S. & Schielzeth, H. rptR: repeatability estimation and variance decomposition by generalized linear mixed-effects models. Methods Ecol. Evol. 8, 1639–1644 (2017).
    Google Scholar 
    Kuhn, M. & Wickham, H. Tidymodels: a collection of packages for modeling and machine learning using tidyverse principles. https://www.tidymodels.org (2020).Wright, M. N. & Ziegler, A. Ranger : A fast implementation of random forests for high dimensional data in C++ andR. J. Stat. Softw. 77, 1–17 (2017).
    Google Scholar  More

  • in

    Climate change will redefine taxonomic, functional, and phylogenetic diversity of Odonata in space and time

    Pecl, G. T. et al. Biodiversity redistribution under climate change: impacts on ecosystems and human well-being. Science 355, eaai9214 (2017).Article 
    PubMed 

    Google Scholar 
    Bellard, C., Bertelsmeier, C., Leadley, P., Thuiller, W. & Courchamp, F. Impacts of climate change on the future of biodiversity. Ecol. Lett. 15, 365–377 (2012).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Diamond, S. E. Contemporary climate‐driven range shifts: putting evolution back on the table. Functional Ecol. 32, 1652–1665 (2018).Article 

    Google Scholar 
    Perry, A. L., Low, P. J., Ellis, J. R. & Reynolds, J. D. Climate change and distribution shifts in marine fishes. Science 308, 1912–1915 (2005).Article 
    CAS 
    PubMed 

    Google Scholar 
    Chen, I. C., Hill, J. K., Ohlemüller, R., Roy, D. B. & Thomas, C. D. Rapid range shifts of species associated with high levels of climate warming. Science 333, 1024–1026 (2011).Article 
    CAS 
    PubMed 

    Google Scholar 
    Lenoir, J. et al. Species better track climate warming in the oceans than on land. Nat. Ecol. Evol. 4, 1044–1059 (2020).Article 
    PubMed 

    Google Scholar 
    Nelson, E. J. et al. Climate change’s impact on key ecosystem services and the human well‐being they support in the US. Front. Ecol. Environ. 11, 483–893 (2013).Article 

    Google Scholar 
    Prather, C. M. et al. Invertebrates, ecosystem services and climate change. Biol. Rev. 88, 327–348 (2013).Article 
    PubMed 

    Google Scholar 
    Cardinale, B. J. et al. Biodiversity loss and its impact on humanity. Nature 486, 59–67 (2012).Article 
    CAS 
    PubMed 

    Google Scholar 
    Ripple, W. J. et al. World scientists’ warning of a climate emergency 2021. BioScience 71, 894–898 (2021).Article 

    Google Scholar 
    Gallagher, R. V., Hughes, L. & Leishman, M. R. Species loss and gain in communities under future climate change: consequences for functional diversity. Ecography 36, 531–540 (2013).Article 

    Google Scholar 
    Saladin, B. et al. Rapid climate change results in long-lasting spatial homogenization of phylogenetic diversity. Nat. Commun. 11, 1–8 (2020).Article 

    Google Scholar 
    Stewart, P. S. et al. Global impacts of climate change on avian functional diversity. Ecol. Lett. 25, 673–685 (2022).Article 
    PubMed 

    Google Scholar 
    Mammola, S., Carmona, C. P., Guillerme, T. & Cardoso, P. Concepts and applications in functional diversity. Funct. Ecol. 35, 1869–1885 (2021).Article 
    CAS 

    Google Scholar 
    Tucker, C. M. et al. A guide to phylogenetic metrics for conservation, community ecology and macroecology. Biol. Rev. 92, 698–715 (2017).Article 
    PubMed 

    Google Scholar 
    Pavoine, S. & Bonsall, M. B. Measuring biodiversity to explain community assembly: a unified approach. Biol. Rev. 86, 792–812 (2011).Article 
    CAS 
    PubMed 

    Google Scholar 
    Petchey, O. L. & Gaston, K. J. Functional diversity (FD), species richness and community composition. Ecol. Lett. 5, 402–411 (2002).Article 

    Google Scholar 
    Wang, S. & Loreau, M. Ecosystem stability in space: α, β and γ variability. Ecol. Lett. 17, 891–901 (2014).Article 
    PubMed 

    Google Scholar 
    Cardoso, P. et al. Partitioning taxon, phylogenetic and functional beta diversity into replacement and richness difference components. J. Biogeogr. 41, 749–761 (2014).Article 

    Google Scholar 
    Hassall, C. Odonata as candidate macroecological barometers for global climate change. Freshwater Sci. 34, 1040–1049 (2015).Article 

    Google Scholar 
    Grewe, Y., Hof, C., Dehling, D. M., Brandl, R. & Brändle, M. Recent range shifts of European dragonflies provide support for an inverse relationship between habitat predictability and dispersal. Global Ecol. Biogeogr. 22, 403–409 (2013).Article 

    Google Scholar 
    Moore, M. P. et al. Sex-specific ornament evolution is a consistent feature of climatic adaptation across space and time in dragonflies. Proc. Natl Acad. Sci. 118, https://doi.org/10.1073/pnas.2101458118 (2021).Castillo-Pérez, E. U., Suárez-Tovar, C. M., González-Tokman, D., Schondube, J. E. & Córdoba-Aguilar, A. Insect thermal limits in warm and perturbed habitats: Dragonflies and damselflies as study cases. J. Thermal Biol. 103, 103164 (2022).Article 

    Google Scholar 
    May, M. L. Odonata: Who they are and what they have done for us lately: Classification and ecosystem services of dragonflies. Insects 10, 62 (2019).Article 
    PubMed Central 

    Google Scholar 
    Hickling, R., Roy, D. B., Hill, J. K. & Thomas, C. D. A northward shift of range margins in British Odonata. Global Change biology 11, 502–506 (2005).Article 

    Google Scholar 
    Hickling, R., Roy, D. B., Hill, J. K., Fox, R. & Thomas, C. D. The distributions of a wide range of taxonomic groups are expanding polewards. Global Change Biol. 12, 450–455 (2006).Heino, J., Virkkala, R. & Toivonen, H. Climate change and freshwater biodiversity: detected patterns, future trends and adaptations in northern regions. Biol. Rev. 84, 39–54 (2009).Article 
    PubMed 

    Google Scholar 
    Mustonen, K. R. et al. Thermal and hydrologic responses to climate change predict marked alterations in boreal stream invertebrate assemblages. Global Change Biol. 24, 2434–2446 (2018).Article 

    Google Scholar 
    Cadotte, M. W. & Tucker, C. M. Difficult decisions: strategies for conservation prioritization when taxonomic, phylogenetic and functional diversity are not spatially congruent. Biol. Conserv. 225, 128–133 (2018).Article 

    Google Scholar 
    Wong, J. S. et al. Comparing patterns of taxonomic, functional and phylogenetic diversity in reef coral communities. Coral Reefs 37, 737–750 (2018).Article 

    Google Scholar 
    Arnan, X., Cerdá, X. & Retana, J. Relationships among taxonomic, functional, and phylogenetic ant diversity across the biogeographic regions of Europe. Ecography 40, 448–457 (2017).Article 

    Google Scholar 
    Strecker, A. L., Olden, J. D., Whittier, J. B. & Paukert, C. P. Defining conservation priorities for freshwater fishes according to taxonomic, functional, and phylogenetic diversity. Ecol. Appl. 21, 3002–3013 (2011).Article 

    Google Scholar 
    Eisenhauer, N., Bonn, A. & Guerra, C. A. Recognizing the quiet extinction of invertebrates. Nat. Commun. 10, 1–3 (2019).Article 

    Google Scholar 
    Cardoso, P. et al. Scientists’ warning to humanity on insect extinctions. Biol. Conserv. 242, 108426 (2020).Article 

    Google Scholar 
    Ovaskainen, O., Rybicki, J. & Abrego, N. What can observational data reveal about metacommunity processes? Ecography 42, 1877–1886 (2019).Article 

    Google Scholar 
    Thomas, C. D. The development of Anthropocene biotas. Philos. Trans. R. Soc. B 375, 20190113 (2020).Article 

    Google Scholar 
    Krosby, M. et al. Climate-induced range overlap among closely related species. Nat. Clim. Change 5, 883–886 (2015).Article 

    Google Scholar 
    Sánchez-Guillén, R. A., Wellenreuther, M., Cordero-Rivera, A. & Hansson, B. Introgression and rapid species turnover in sympatric damselflies. BMC Evol. Biol. 11, 1–17 (2011).Article 

    Google Scholar 
    Bybee, S. et al. Odonata (dragonflies and damselflies) as a bridge between ecology and evolutionary genomics. Front. Zool. 13, 1–20 (2016).Article 

    Google Scholar 
    Tobias, N. & Monika, W. Does taxonomic homogenization imply functional homogenization in temperate forest herb layer communities? Plant Ecol. 213, 431–443 (2012).Article 

    Google Scholar 
    Pauls, S. U., Nowak, C., Bálint, M. & Pfenninger, M. The impact of global climate change on genetic diversity within populations and species. Mol. Ecol. 22, 925–946 (2013).Article 
    PubMed 

    Google Scholar 
    Ball-Damerow, J. E., M’Gonigle, L. K. & Resh, V. H. Changes in occurrence, richness, and biological traits of dragonflies and damselflies (Odonata) in California and Nevada over the past century. Biodiversity Conserv. 23, 2107–2126 (2014).Article 

    Google Scholar 
    McGoff, E., Solimini, A. G., Pusch, M. T., Jurca, T. & Sandin, L. Does lake habitat alteration and land-use pressure homogenize European littoral macroinvertebrate communities? J. Appl. Ecol. 50, 1010–1018 (2013).Article 

    Google Scholar 
    Vilenica, M., Kerovec, M., Pozojević, I. & Mihaljević, Z. Odonata assemblages in anthropogenically impacted lotic habitats. J. Limnol. 80, 1968 (2021).Mammola, S. et al. Challenges and opportunities of species distribution modelling of terrestrial arthropod predators. Diversity Distrib. 27, 2596–2614 (2021).Article 

    Google Scholar 
    Fourcade, Y., Besnard, A. G. & Secondi, J. Paintings predict the distribution of species, or the challenge of selecting environmental predictors and evaluation statistics. Global Ecol. Biogeogr. 27, 245–256 (2018).Article 

    Google Scholar 
    Kalkman, V. J. et al. Diversity and conservation of European dragonflies and damselflies (Odonata). Hydrobiologia 811, 269–282 (2018). .Barve, N. et al. The crucial role of the accessible area in ecological niche modeling and species distribution modeling. Ecol. Model. 222, 1810–1819 (2011).Article 

    Google Scholar 
    Miller, J. A. & Holloway, P. Incorporating movement in species distribution models. Progr. Phys. Geogr. 39, 837–849 (2015).Article 

    Google Scholar 
    Freeman, B. G., Scholer, M. N., Ruiz-Gutierrez, V. & Fitzpatrick, J. W. Climate change causes upslope shifts and mountaintop extirpations in a tropical bird community. Proc. Natl Acad. Sci. 115, 11982–11987 (2018).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Parmesan, C. Ecological and evolutionary responses to recent climate change. Annu. Rev. Ecol. Evol. Syst. 37, 637–669 (2006).Article 

    Google Scholar 
    Pinkert, S. et al. Evolutionary processes, dispersal limitation and climatic history shape current diversity patterns of European dragonflies. Ecography 41, 795–804 (2018).Article 

    Google Scholar 
    Comte, L., Murienne, J. & Grenouillet, G. Species traits and phylogenetic conservatism of climate-induced range shifts in stream fishes. Nat. Commun. 5, 1–9 (2014).Article 

    Google Scholar 
    Buckley, L. B. & Kingsolver, J. G. Functional and phylogenetic approaches to forecasting species’ responses to climate change. Ann. Rev. Ecol. Evol. Syst. 43, 205–226 (2012).Article 

    Google Scholar 
    Tikhonov, G. et al. Joint species distribution modelling with the R‐package Hmsc. Methods Ecol. Evol. 11, 442–447 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Corbet, P. S. The life-history of the emperor dragonfly Anax imperator Leach (Odonata: Aeshnidae). J. Animal Ecol. 1–69. https://doi.org/10.2307/1781 (1957).Guisan, A. & Thuiller, W. Predicting species distribution: offering more than simple habitat models. Ecol. Lett. 8(9), 993–1009 (2005).Article 
    PubMed 

    Google Scholar 
    Peterson, A. T. et al. Ecological niches and geographic distributions (MPB-49) (Princeton University Press, 2011).Franklin, J. Mapping species distributions: spatial inference and prediction (Cambridge University Press, 2010).Ryo, M. et al. Explainable artificial intelligence enhances the ecological interpretability of black‐box species distribution models. Ecography 44(2), 199–205 (2021).Article 

    Google Scholar 
    Guisan, A. et al. Predicting species distributions for conservation decisions. Ecol. Lett. 16, 1424–1435 (2013).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Adams, M. P. et al. Prioritizing localized management actions for seagrass conservation and restoration using a species distribution model. Aquat. Conserv. Marine Freshwater Ecosyst. 26, 639–659 (2016).Ficetola, G. F., Thuiller, W. & Padoa‐Schioppa, E. From introduction to the establishment of alien species: bioclimatic differences between presence and reproduction localities in the slider turtle. Diversity Distrib. 15, 108–116 (2009).Article 

    Google Scholar 
    Wang, Y., Xie, B., Wan, F., Xiao, Q. & Dai, L. Application of ROC curve analysis in evaluating the performance of alien species’ potential distribution models. Biodiversity Sci. 15, 365 (2007).Article 

    Google Scholar 
    Santini, L., Benítez‐López, A., Maiorano, L., Čengić, M. & Huijbregts, M. A. Assessing the reliability of species distribution projections in climate change research. Diversity Distrib. 27, 1035–1050 (2021).Article 

    Google Scholar 
    Guyennon, A. et al. Colonization and extinction dynamics and their link to the distribution of European trees at the continental scale. J. Biogeogr. 49, 117–129 (2022).Article 

    Google Scholar 
    Pritchard, G. & Leggott, M. A. Temperature, incubation rates and origins of dragonflies. Adv. Odonatol. 3, 121–126 (1987).
    Google Scholar 
    Clausnitzer, V. et al. Odonata enter the biodiversity crisis debate: the first global assessment of an insect group. Biol. Conserv. 142, 1864–1869 (2009).Article 

    Google Scholar 
    Córdoba-Aguilar, A. (Ed.). Dragonflies and damselflies: model organisms for ecological and evolutionary research (OUP Oxford, 2008).Corbet, P. S. et al. Dragonflies: behaviour and ecology of Odonata (Harley books, 1999).Troast, D., Suhling, F., Jinguji, H., Sahlén, G. & Ware, J. A global population genetic study of Pantala flavescens. PloS One 11, e0148949 (2016).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Harabiš, F. & Dolný, A. The effect of ecological determinants on the dispersal abilities of central European dragonflies (Odonata). Odonatologica 40, 17 (2011).
    Google Scholar 
    Boudot, J. P. & Kalkman, V. J. (eds) Atlas of the European dragonflies and damselflies (KNNV publishing, 2015).Dijkstra, K. D. & Schröter, A. Field guide to the dragonflies of Britain and Europe (Bloomsbury Publishing, 2020).Titley, M. A., Snaddon, J. L. & Turner, E. C. Scientific research on animal biodiversity is systematically biased towards vertebrates and temperate regions. PloS One 12, e0189577 (2017).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Beck, J., Böller, M., Erhardt, A. & Schwanghart, W. Spatial bias in the GBIF database and its effect on modeling species’ geographic distributions. Ecol. Inf. 19, 10–15 (2014).Article 

    Google Scholar 
    Zizka, A. et al. No one-size-fits-all solution to clean GBIF. PeerJ 8, e9916 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Burgman, M. A. & Fox, J. C. Bias in species range estimates from minimum convex polygons: implications for conservation and options for improved planning, Animal Conservation Forum (6, No. 1, pp. 19–28 (Cambridge University Press, 2003). https://doi.org/10.1017/S1367943003003044Calenge, C. The package “adehabitat” for the R software: a tool for the analysis of space and habitat use by animals. Ecol. Modell. 197, 516–519 (2006).Article 

    Google Scholar 
    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 
    Hijmans, R. J. Raster: geographic data analysis and modeling. https://CRAN.R-project.org/package=raster (2020).Hijmans, R. J., Phillips S., Leathwick J. & Elith J. Dismo: species distribution modeling. https://CRAN.R-project.org/package=dismo (2020).Title, P. O. & Bemmels, J. B. ENVIREM: an expanded set of bioclimatic and topographic variables increases flexibility and improves performance of ecological niche modeling. Ecography 41, 291–307 (2018).Article 

    Google Scholar 
    Dormann, C. F. et al. Collinearity: a review of methods to deal with it and a simulation study evaluating their performance. Ecography 36, 27–46 (2013).Article 

    Google Scholar 
    Mukaka, M. M. A guide to appropriate use of correlation coefficient in medical research. Malawi Med. J. 24, 69–71 (2012).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Zuur, A. F., Ieno, E. N. & Elphick, C. S. A protocol for data exploration to avoid common statistical problems. Methods Ecol. Evol. 1, 3–14 (2010).Article 

    Google Scholar 
    Hausfather, Z. & Peters, G. P. Emissions–the ‘business as usual’story is misleading https://doi.org/10.1038/d41586-020-00177-3 (2020)Mammola, S., Milano, F., Vignal, M., Andrieu, J. & Isaia, M. Associations between habitat quality, body size and reproductive fitness in the alpine endemic spider Vesubia jugorum. Global Ecol. Biogeogr. 28, 1325–1335 (2019).Article 

    Google Scholar 
    Mammola, S., Goodacre, S. L. & Isaia, M. Climate change may drive cave spiders to extinction. Ecography 41(1), 233–243 (2018).Article 

    Google Scholar 
    Hastie, T. J. & Tibshirani, R. J. Generalized additive models (Routledge, 2017).Phillips, S. J., Anderson, R. P. & Schapire, R. E. Maximum entropy modeling of species geographic distributions. Ecol. Model. 190(3-4), 231–259 (2006).Article 

    Google Scholar 
    Phillips, S. J., Dudík, M. & Schapire, R. E. (2004). A maximum entropy approach to species distribution modeling. In Proceedings of the twenty-first international conference on Machine learning (p. 83).https://doi.org/10.1145/1015330.1015412 (2004).Elith, J., Leathwick, J. R. & Hastie, T. A working guide to boosted regression trees. J. Animal Ecol. 77, 802–813 (2008).Article 
    CAS 

    Google Scholar 
    Elith, J. et al. Novel methods improve prediction of species’ distributions from occurrence data. Ecography 29, 129–151 (2006).Article 

    Google Scholar 
    Araújo, M. B. & New, M. Ensemble forecasting of species distributions. Trends Ecol. Evol. 22, 42–47 (2007).Article 
    PubMed 

    Google Scholar 
    Grenouillet, G., Buisson, L., Casajus, N. & Lek, S. Ensemble modelling of species distribution: the effects of geographical and environmental ranges. Ecography 34, 9–17 (2011).Article 

    Google Scholar 
    Phillips, S. J. et al. Sample selection bias and presence‐only distribution models: implications for background and pseudo‐absence data. Ecol. Appl. 19, 181–197 (2009).Article 
    PubMed 

    Google Scholar 
    Hirzel, A. H., Le Lay, G., Helfer, V., Randin, C. & Guisan, A. Evaluating the ability of habitat suitability models to predict species presences. Ecol. Model. 199, 142–152 (2006).Article 

    Google Scholar 
    Zhang, Z. et al. Lineage‐level distribution models lead to more realistic climate change predictions for a threatened crayfish. Diversity Distrib. 27, 684–695 (2021).Article 

    Google Scholar 
    Liu, C., Berry, P. M., Dawson, T. P. & Pearson, R. G. Selecting thresholds of occurrence in the prediction of species distributions. Ecography 28, 385–393 (2005).Article 

    Google Scholar 
    Martín‐Vélez, V. & Abellán, P. Effects of climate change on the distribution of threatened invertebrates in a Mediterranean hotspot. Insect Conserv. Divers. https://doi.org/10.1111/icad.12563 (2022).Qiao, H., Soberon, J. & Peterson, A. T. No silver bullets in correlative ecological niche modelling: insights from testing among many potential algorithms for niche estimation. Methods Ecol. Evol. 6, 1126–1136 (2015).Article 

    Google Scholar 
    Zurell, D. et al. A standard protocol for reporting species distribution models. Ecography 43, 1261–1277 (2020).Article 

    Google Scholar 
    Petchey, O. L. & Gaston, K. J. Functional diversity: back to basics and looking forward. Ecol. Lett. 9, 741–758 (2006).Article 
    PubMed 

    Google Scholar 
    Faith, D. P. Conservation evaluation and phylogenetic diversity. Biol. Conserv. 61(1), 1–10 (1992).Article 

    Google Scholar 
    Cadotte, M. W. et al. Phylogenetic diversity metrics for ecological communities: integrating species richness, abundance and evolutionary history. Ecol. Lett. 13, 96–105 (2010).Article 
    PubMed 

    Google Scholar 
    Pollock, L. J. et al. Protecting biodiversity (in all its complexity): new models and methods. Trends Ecol. Evol. 35, 1119–1128 (2020).Article 
    PubMed 

    Google Scholar 
    Corbet, P. S. ‘Biology of Odonata’. Ann. Rev. Entomol. 25, 189–217 (1980).Article 

    Google Scholar 
    Mitchell. Dragonfly locomotion: Ecology, form and function. PhD thesis, (University of Leeds, 2018). https://etheses.whiterose.ac.uk/21211/.The GIMP Development Team. GIMP (version 2.10.12). https://www.gimp.org (2019).Weller, H. Colordistance: distance metrics for image color similarity. https://CRAN.R-project.org/package=colordistance (2020).R Core Team. R: a language and environment for statistical computing (R Foundation for Statistical Computing, Vienna, Austria, 2020). https://www.R-project.org/.de Bello, F., Botta‐Dukát, Z., Lepš, J. & Fibich, P. Towards a more balanced combination of multiple traits when computing functional differences between species. Methods Ecol. Evol. 12, 443–448 (2021).Article 

    Google Scholar 
    Hassall, C. & Thompson, D. J. The effects of environmental warming on Odonata: a review. Int. J. Odonatol. 11, 131–153 (2008).Article 

    Google Scholar 
    Acquah‐Lamptey, D., Brändle, M., Brandl, R. & Pinkert, S. Temperature‐driven color lightness and body size variation scale to local assemblages of European Odonata but are modified by propensity for dispersal. Ecol. Evol. 10, 8936–8948 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Outomuro, D. & Johansson, F. Wing morphology and migration status, but not body size, habitat or Rapoport’s rule predict range size in North‐American dragonflies (Odonata: Libellulidae). Ecography 42, 309–320 (2019).Article 

    Google Scholar 
    Rundle, S. D., Bilton, D. T., Abbott, J. C. & Foggo, A. Range size in North American Enallagma damselflies correlates with wing size. Freshwater Biol. 52, 471–477 (2007).Article 

    Google Scholar 
    Finlayson, C. M. et al. The second warning to humanity–providing a context for wetland management and policy. Wetlands 39, 1–5 (2019).Article 

    Google Scholar 
    Okude, G. & Futahashi, R. Pigmentation and color pattern diversity in Odonata. Curr. Opin. Genet. Dev. 69, 14–20 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Mani, M. S. Ecology and biogeography of high altitude insects, vol. 4. (Springer Science & Business Media, 2013).Suárez‐Tovar, C. M., Guillermo‐Ferreira, R., Cooper, I. A., Cezário, R. R. & Córdoba‐Aguilar, A. Dragon colors: the nature and function of Odonata (dragonfly and damselfly) coloration. J. Zool. https://doi.org/10.1111/jzo.12963 (2022).Katoh, K. & Standley, D. M. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol. Biol. Evol. 30, 772–780 (2013).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Vaidya, G., Lohman, D. J. & Meier, R. SequenceMatrix: concatenation software for the fast assembly of multi‐gene datasets with character set and codon information. Cladistics 27, 171–180 (2011).Article 
    PubMed 

    Google Scholar 
    Lanfear, R., Frandsen, P. B., Wright, A. M., Senfeld, T. & Calcott, B. PartitionFinder 2: new methods for selecting partitioned models of evolution for molecular and morphological phylogenetic analyses. Mol. Biol. Evol. 34, 772–773 (2017).CAS 
    PubMed 

    Google Scholar 
    Bouckaert, R. et al. BEAST 2.5: an advanced software platform for Bayesian evolutionary analysis. PLoS Comput. Biol. 15, e1006650 (2019).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Rambaut, A., Drummond, A. J., Xie, D., Baele, G. & Suchard, M. A. Posterior summarization in Bayesian phylogenetics using Tracer 1.7. Syst. Biol. 67, 901–904 (2018).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Cardoso, P., Stefano, M., Francois, R. & Jose, C. C. BAT: biodiversity assessment tools. https://CRAN.R-project.org/package=BAT (2021).Robert J. H. geosphere: spherical trigonometry. R package version 1.5-14. https://CRAN.R-project.org/package=geosphere (2021).Revell, L. J. phytools: an R package for phylogenetic comparative biology (and other things). Methods Ecol. Evol. 217–223 https://doi.org/10.1111/j.2041-210X.2011.00169.x (2012).Joy, J. B., Liang, R. H., McCloskey, R. M., Nguyen, T. & Poon, A. F. Ancestral reconstruction. PLoS Comput. Biol. 12, e1004763 (2016).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Orme, D. et al. caper: comparative analyses of phylogenetics and evolution in R. R package version 1.0.1 (2018).Silva, L. F. et al. Functional responses of Odonata larvae to human disturbances in neotropical savanna headwater streams. Ecol. Indic. 133, 108367 (2021).Article 

    Google Scholar  More

  • in

    The impact of environmental and climatic variables on genetic diversity and plant functional traits of the endangered tuberous orchid (Orchis mascula L.)

    Read, Q. D., Moorhead, L. C., Swenson, N. G., Bailey, J. K. & Sanders, N. J. Convergent effects of elevation on functional leaf traits within and among species. Funct. Ecol. 28, 37–45. https://doi.org/10.1111/1365-2435.12162 (2014).Article 

    Google Scholar 
    Isbell, F. et al. Biodiversity increases the resistance of ecosystem productivity to climate extremes. Nature 526, 574–577. https://doi.org/10.1038/nature15374 (2015).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Vellend, M. & Geber, M. A. Connections between species diversity and genetic diversity. Ecol. Lett. 8, 767–781. https://doi.org/10.1111/j.1461-0248.2005.00775.x (2005).Article 

    Google Scholar 
    Hart, S. P., Schreiber, S. J. & Levine, J. M. How variation between individuals affects species coexistence. Ecol. Lett. 19, 825–838. https://doi.org/10.1111/ele.12618 (2016).Article 
    PubMed 

    Google Scholar 
    Paschke, M. C. & Schmid, B. Relationship between population size, allozyme variation, and plant performance in the narrow endemic Cochlearia bavarica. Conserv. Genet. 3, 131–144 (2002).Article 
    CAS 

    Google Scholar 
    Soleimani, V., Baum, B. & Johnson, D. A. AFLP and pedigree-based genetic diversity estimates in modern cultivars of durum wheat [Triticum turgidum L. subsp. durum (Desf.) Husn.]. Theor. Appl. Genet. 104, 350–357. https://doi.org/10.1007/s001220100714 (2002).Article 
    CAS 
    PubMed 

    Google Scholar 
    Hughes, A. R., Inouye, B. D., Johnson, M. T., Underwood, N. & Vellend, M. Ecological consequences of genetic diversity. Ecol. Lett. 11, 609–623. https://doi.org/10.1111/j.1461-0248.2008.01179.x (2008).Article 
    PubMed 

    Google Scholar 
    Prati, D., Peintinger, M. & Fischer, M. Genetic composition, genetic diversity and small-scale environmental variation matter for the experimental reintroduction of a rare plant. J. Plant Ecol. https://doi.org/10.1093/jpe/rtv067 (2016).Article 

    Google Scholar 
    Barrett, R. D. & Schluter, D. Adaptation from standing genetic variation. Trends Ecol. Evol. 23, 38–44. https://doi.org/10.1016/j.tree.2007.09.008 (2008).Article 
    PubMed 

    Google Scholar 
    Atwater, D. Z. & Callaway, R. M. Testing the mechanisms of diversity-dependent overyielding in a grass species. Ecology 96, 3332–3342. https://doi.org/10.1890/15-0889.1 (2015).Article 
    PubMed 

    Google Scholar 
    Cook-Patton, S. C., McArt, S. H., Parachnowitsch, A. L., Thaler, J. S. & Agrawal, A. A. A direct comparison of the consequences of plant genotypic and species diversity on communities and ecosystem function. Ecology 92, 915–923. https://doi.org/10.1890/10-0999.1 (2011).Article 
    PubMed 

    Google Scholar 
    Whitney, K. D. et al. Experimental drought reduces genetic diversity in the grassland foundation species Bouteloua eriopoda. Oecologia 189, 1107–1120. https://doi.org/10.1007/s00442-019-04371-7 (2019).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Violle, C. et al. Let the concept of trait be functional!. Oikos 116, 882–892. https://doi.org/10.1111/j.0030-1299.2007.15559.x (2007).Article 

    Google Scholar 
    Kattge, J. et al. TRY plant trait database-enhanced coverage and open access. Glob. Change Biol. 26, 119–188. https://doi.org/10.1111/gcb.14904 (2020).Article 
    ADS 

    Google Scholar 
    König, P. et al. Advances in flowering phenology across the Northern Hemisphere are explained by functional traits. Glob. Ecol. Biogeogr. 27, 310–321 (2018).Article 

    Google Scholar 
    Robinson, K. M., Ingvarsson, P. K., Jansson, S. & Albrectsen, B. R. Genetic variation in functional traits influences arthropod community composition in aspen (Populus tremula L.). PLoS ONE 7, e37679. https://doi.org/10.1371/journal.pone.0037679 (2012).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Karbstein, K., Prinz, K., Hellwig, F. & Römermann, C. Plant intraspecific functional trait variation is related to within-habitat heterogeneity and genetic diversity in Trifolium montanum L. Ecol. Evol. 10, 5015–5033. https://doi.org/10.1002/ece3.6255 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kaya, S. & Tekin, A. R. The effect of salep content on the rheological characteristics of a typical ice-cream mix. J. Food Eng. 47, 59–62. https://doi.org/10.1016/S0260-8774(00)00093-5 (2001).Article 

    Google Scholar 
    Ktistis, G. & Georgakopoulos, P. P. Rheology of salep mucilages. Pharmazie 46, 55–56 (1991).CAS 

    Google Scholar 
    Kayacier, A. & Dogan, M. Rheological properties of some gums-salep mixed solutions. J. Food Eng. 72, 261–265. https://doi.org/10.1016/j.jfoodeng.2004.12.005 (2006).Article 
    CAS 

    Google Scholar 
    Sen, M. A., Palabiyik, I. & Kurultay, S. The effect of saleps obtained from various Orchidacease species on some physical and sensory properties of ice cream. Food Sci. Technol. 39, 82–87. https://doi.org/10.1590/fst.26017 (2019).Article 

    Google Scholar 
    Farhoosh, R. & Riazi, A. A compositional study on two current types of salep in Iran and their rheological properties as a function of concentration and temperature. Food Hydrocoll. 21, 660–666. https://doi.org/10.1016/j.foodhyd.2006.07.021 (2007).Article 
    CAS 

    Google Scholar 
    Ghorbani, A., Zarre, S., Gravendeel, B. & de Boer, H. J. Illegal wild collection and international trade of CITES-listed terrestrial orchid tubers in Iran. Traffic Bullet. 26, 52–58 (2014).
    Google Scholar 
    Lenoir, J., Gégout, J.-C., Marquet, P., De Ruffray, P. & Brisse, H. A significant upward shift in plant species optimum elevation during the 20th century. Science 320, 1768–1771. https://doi.org/10.1126/science.1156831 (2008).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Chen, Z.-Q., Algeo, T. J. & Fraiser, M. L. Organism-environment interactions during the Permian-Triassic mass extinction and its aftermath. Palaios 28, 661–663. https://doi.org/10.2110/palo.2012.p12-102r (2013).Article 
    ADS 

    Google Scholar 
    Ebrahimi, A. et al. Evaluation of phenotypic diversity of the endangered orchid (Orchis mascula): Emphasizing on breeding, conservation and development. S. Afr. J. Bot. 132, 304–315. https://doi.org/10.1016/j.sajb.2020.05.013 (2020).Article 

    Google Scholar 
    Ghorbani, A., Gravendeel, B., Naghibi, F. & de Boer, H. Wild orchid tuber collection in Iran: A wake-up call for conservation. Biodivers. Conserv. 23, 2749–2760. https://doi.org/10.1007/s10531-014-0746-y (2014).Article 

    Google Scholar 
    Barrett, S. C. & Kohn, J. R. The application of minimum viable population theory to plants. Genetics and conservation of rare plants 3–1 (Oxford University Press, 1991).
    Google Scholar 
    Yun, S. A., Son, H.-D., Im, H.-T. & Kim, S.-C. Genetic diversity and population structure of the endangered orchid Pelatantheria scolopendrifolia (Orchidaceae) in Korea. PLoS ONE 15, e0237546. https://doi.org/10.1371/journal.pone.0237546 (2020).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Gholami, S., Vafaee, Y., Nazari, F. & Ghorbani, A. Exploring genetic variations in threatened medicinal orchids using start codon targeted (SCoT) polymorphism and marker-association with seed morphometric traits. Physiol. Mol. Biol. Plants 27, 769–785. https://doi.org/10.1007/s12298-021-00978-4 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Gholami, S., Vafaee, Y., Nazari, F. & Ghorbani, A. Molecular characterization of endangered Iranian terrestrial orchids using ISSR markers and association with floral and tuber-related phenotypic traits. Physiol. Mol. Biol. Plants 27, 53–68. https://doi.org/10.1007/s12298-020-00920-0 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kaki, A., Vafaee, Y. & Khadivi, A. Genetic variation of Anacamptis coriophora, Dactylorhiza umbrosa, Himantoglossum affine, Orchis mascula, and Ophrys schulzei in the western parts of Iran. Ind. Crops Prod. 156, 112854. https://doi.org/10.1016/j.indcrop.2020.112854 (2020).Article 
    CAS 

    Google Scholar 
    Falk, D., & Holsinger, K. E. Genetic sampling guidelines for conservation collections of endangered plants (1991).Dempewolf, H. et al. Past and future use of wild relatives in crop breeding. Crop Sci. 57, 1070–1082. https://doi.org/10.2135/cropsci2016.10.0885 (2017).Article 

    Google Scholar 
    Renz, J. Flora Iranica. Part 126: Orchidaceae (1978).Shahsavari, A. Flora of Iran. Part 57: Orchidaceae (2008).Boulila, A., Béjaoui, A., Messaoud, C. & Boussaid, M. Genetic diversity and population structure of Teucrium polium (Lamiaceae) in Tunisia. Biochem. Gen. 48, 57–70. https://doi.org/10.1007/s10528-009-9295-6 (2010).Article 
    CAS 

    Google Scholar 
    Zannou, A., Struik, P., Richards, P., Zoundjih, E. & Yam, J. (Dioscorea spp.) responses to the environmental variability in the Guinea Sudan zone of Benin. Afr. J. Agric. Res. 10, 4913–4925. https://doi.org/10.5897/AJAR2013.8099 (2015).Article 

    Google Scholar 
    Sujii, P. et al. Morphological and molecular characteristics do not confirm popular classification of the Brazil nut tree in Acre, Brazil. Genet. Mol. Res. https://doi.org/10.4238/2013.september.27.3 (2013).Article 
    PubMed 

    Google Scholar 
    Nadeem, M. A. et al. DNA molecular markers in plant breeding: current status and recent advancements in genomic selection and genome editing. Biotechnol. Biotechnol. Equip. 32, 261–285. https://doi.org/10.1080/13102818.2017.1400401 (2018).Article 
    CAS 

    Google Scholar 
    Jacquemyn, H., Brys, R., Adriaens, D., Honnay, O. & Roldán-Ruiz, I. Effects of population size and forest management on genetic diversity and structure of the tuberous orchid Orchis mascula. Conserv. Genet. 10, 161–168. https://doi.org/10.1007/s10592-008-9543-z (2009).Article 

    Google Scholar 
    Mitchell, P. & Woodward, F. Responses of three woodland herbs to reduced photosynthetically active radiation and low red to far-red ratio in shade. J. Ecol. https://doi.org/10.2307/2260575 (1988).Article 

    Google Scholar 
    Likens, G. E., Bormann, F. H., Johnson, N. M., Fisher, D. & Pierce, R. S. Effects of forest cutting and herbicide treatment on nutrient budgets in the Hubbard Brook watershed-ecosystem. Ecol. Monog. 40, 23–47. https://doi.org/10.2307/1942440 (1970).Article 

    Google Scholar 
    Jacquemyn, H. & Brys, R. Lack of strong selection pressures maintains wide variation in floral traits in a food-deceptive orchid. Ann. Bot. 126, 445–453. https://doi.org/10.1093/aob/mcaa080 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Olaya-Arenas, P., Meléndez-Ackerman, E. J., Pérez, M. E. & Tremblay, R. Demographic response by a small epiphytic orchid. Am. J. Bot. 98, 2040–2048. https://doi.org/10.3732/ajb.1100223 (2011).Article 
    PubMed 

    Google Scholar 
    Primack, R. B., Miao, S. & Becker, K. R. Costs of reproduction in the pink lady’s slipper orchid (Cypripedium acaule): Defoliation, increased fruit production, and fire. Am. J. Bot. 81, 1083–1090. https://doi.org/10.2307/2446500 (1994).Article 

    Google Scholar 
    Whigham, D. F. & O’Neill, J. P. Dynamics of flowering and fruit production in two eastern North American terrestrial orchids. In Tipularia Discolor and Liparis Lilifolia in Population Ecology of Terrestrial Orchids (eds Wells, T. C. E. & Willems, J. H.) 89–101 (SPB Academic Publishers, 1991).
    Google Scholar 
    Tekinşen, K. K. & Güner, A. Chemical composition and physicochemical properties of tubera salep produced from some Orchidaceae species. Food Chem. 121, 468–471. https://doi.org/10.1016/j.foodchem.2009.12.066 (2010).Article 
    CAS 

    Google Scholar 
    Whigham, D. F. Biomass and nutrient allocation of Tipularia discolor (Orchidaceae). Oikos https://doi.org/10.2307/3544398 (1984).Article 

    Google Scholar 
    Mattila, E. & Kuitunen, M. T. Nutrient versus pollination limitation in Platanthera bifolia and Dactylorhiza incarnata (Orchidaceae). Oikos 89, 360–366. https://doi.org/10.1034/j.1600-0706.2000.890217.x (2000).Article 

    Google Scholar 
    Xu, W. et al. Drought stress condition increases root to shoot ratio via alteration of carbohydrate partitioning and enzymatic activity in rice seedlings. Acta Physiol. Plant. 37, 9. https://doi.org/10.1007/s11738-014-1760-0 (2015).Article 
    CAS 

    Google Scholar 
    March-Salas, M., Fandos, G. & Fitze, P. S. Effects of intrinsic environmental predictability on intra-individual and intra-population variability of plant reproductive traits and eco-evolutionary consequences. Ann. Bot. 127, 413–423. https://doi.org/10.1093/aob/mcaa096 (2021).Article 
    CAS 
    PubMed 

    Google Scholar 
    Jump, A. S., Marchant, R. & Peñuelas, J. Environmental change and the option value of genetic diversity. Trends Plant Sci. 14, 51–58. https://doi.org/10.1016/j.tplants.2008.10.002 (2009).Article 
    CAS 
    PubMed 

    Google Scholar 
    Huang, J. et al. Global semi-arid climate change over last 60 years. Clim. Dyn. 46, 1131–1150. https://doi.org/10.1007/s00382-015-2636-8 (2016).Article 

    Google Scholar 
    Crémieux, L., Bischoff, A., Müller-Schärer, H. & Steinger, T. Gene flow from foreign provenances into local plant populations: Fitness consequences and implications for biodiversity restoration. Am. J. Bot. 97, 94–100. https://doi.org/10.3732/ajb.0900103 (2010).Article 
    PubMed 

    Google Scholar 
    Garant, D., Forde, S. E. & Hendry, A. P. The multifarious effects of dispersal and gene flow on contemporary adaptation. Funct. Ecol. 21, 434–443. https://doi.org/10.1111/j.1365-2435.2006.01228.x (2007).Article 

    Google Scholar 
    Jacquemyn, H. et al. Multigenerational analysis of spatial structure in the terrestrial, food-deceptive orchid Orchis mascula. J. Ecol. 97, 206–216. https://doi.org/10.1111/j.1365-2745.2008.01464.x (2009).Article 

    Google Scholar 
    Siepielski, A. M. et al. Precipitation drives global variation in natural selection. Science 355, 959–962. https://doi.org/10.1126/science.aag2773 (2017).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Ene, C. O., Ogbonna, P. E., Agbo, C. U. & Chukwudi, U. P. Studies of phenotypic and genotypic variation in sixteen cucumber genotypes. Chilean J. Agric. Res. 76, 307–313. https://doi.org/10.4067/S0718-58392016000300007 (2016).Article 

    Google Scholar 
    Pradhan, S. K. et al. Population structure, genetic diversity and molecular marker-trait association analysis for high temperature stress tolerance in rice. PLoS ONE 11, e0160027. https://doi.org/10.1371/journal.pone.0160027 (2016).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Swarup, S. et al. Genetic diversity is indispensable for plant breeding to improve crops. Crop Sci. 61, 839–852. https://doi.org/10.1002/csc2.20377 (2021).Article 

    Google Scholar 
    Patzak, A. Plantaginaceae in KH Rechinger Flora Iranica 15: 1–21 (Academische Druck und Verlagsantalt, 1965).
    Google Scholar 
    Mehrvarz Saeidi, S. Plantaginaceae Family Vol. 14 (Research Institute of Forests and Rangelands, 1995).
    Google Scholar 
    Limited, M. I. I. Glucomannan assay procedure KGLUM 10/04. Ireland (2004).Limited, M. I. I. Total starch assay procedure (amyloglucosidase/a-Amylase Method) AA/AMG 11/01. AOAC Method 996.11.Ireland (2004).Bradshaw, H., Otto, K. G., Frewen, B. E., McKay, J. K. & Schemske, D. W. Quantitative trait loci affecting differences in floral morphology between two species of monkeyflower (Mimulus). Genetics 149, 367–382. https://doi.org/10.1093/genetics/149.1.367 (1998).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Del Sal, G., Manfioletti, G. & Schneider, C. The CTAB-DNA precipitation method: A common mini-scale preparation of template DNA from phagemids, phages or plasmids suitable for sequencing. Biotechniques 7, 514–520 (1989).PubMed 

    Google Scholar 
    Vos, P. et al. AFLP: A new technique for DNA fingerprinting. Nucl. Acid Res. 23, 4407–4414. https://doi.org/10.1093/nar/23.21.4407 (1995).Article 
    CAS 

    Google Scholar 
    Bassam, B. J., Caetano-Anollés, G. & Gresshoff, P. M. Fast and sensitive silver staining of DNA in polyacrylamide gels. Anal. Biochem. 196, 80–83. https://doi.org/10.1016/0003-2697(91)90120-I (1991).Article 
    CAS 
    PubMed 

    Google Scholar 
    Husson, F., Josse, J., Le, S., Mazet, J. & Husson, M. F. Package ‘FactoMineR’. An R package 96, 698 (2016).
    Google Scholar 
    Lê, S., Josse, J. & Husson, F. FactoMineR: An R package for multivariate analysis. J. Stat. Softw. 25, 1–18 (2008).Article 

    Google Scholar 
    Galili, T. in The R User Conference, useR! 2017 July 4–7 2017 Brussels, Belgium. 219.Wei, T. et al. Package ‘corrplot’. 56, e24 (2017).Yeh, F. POPGENE (version 1.3. 1). Microsoft Window-Bases Freeware for Population Genetic Analysis. http://www.ualbertaca/~fyeh/ (1999).Wickham, H. & Chang, W. URL: http://CRAN.R-project.org/package=ggplot2.ggplot2: An implementation of the Grammar of Graphics. 3 (2008).Kolde, R. & Kolde, M. R. Package’ pheatmap’. R package 1, 790 (2015).
    Google Scholar 
    Landgraf, A. J. & Lee, Y. Dimensionality reduction for binary data through the projection of natural parameters. J. Mult. Anal. 180, 104668. https://doi.org/10.1016/j.jmva.2020.104668 (2020).Article 
    MathSciNet 
    MATH 

    Google Scholar 
    Pritchard, J. K., Stephens, M. & Donnelly, P. Inference of population structure using multilocus genotype data. Genetics 155, 945–959. https://doi.org/10.1093/genetics/155.2.945 (2000).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Earl, D. A. & VonHoldt, B. M. STRUCTURE HARVESTER: A website and program for visualizing STRUCTURE output and implementing the Evanno method. Biol. J. Linn. Soc. 4, 359–361. https://doi.org/10.1007/s12686-011-9548-7 (2012).Article 

    Google Scholar 
    Oksanen, J. et al. Vegan: Community Ecology Package. R package version 2.4-3 (2016).Kamvar, Z. N., Tabima, J. F. & Grünwald, N. J. Poppr: An R package for genetic analysis of populations with clonal, partially clonal, and/or sexual reproduction. Peer. J. 2, e281 (2014).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Vavrek, M. J. Fossil: Palaeoecological and palaeogeographical analysis tools. Palaeontol. Electron. 14, 16. https://doi.org/10.7717/peerj.281 (2011).Article 

    Google Scholar 
    Bradbury, P. J. et al. TASSEL: Software for association mapping of complex traits in diverse samples. Bioinformatics 23, 2633–2635. https://doi.org/10.1093/bioinformatics/btm308 (2007).Article 
    CAS 
    PubMed 

    Google Scholar  More

  • in

    Population status, distribution and trophic implications of Pinna nobilis along the South-eastern Italian coast

    According to the target of the present study, the mortality incidence on P. nobilis in local populations along the Apulia peninsula (the Southeast coast of Italy) following the MME was assessed. In addition, an investigation on the species distribution and densities in the Adriatic and the Ionian Sea was carried out, which allowed us to build a picture of species populations before the MME.Concerning the P. nobilis distribution in the Apulia region before the MME, unfortunately, there is a lack of information at the wide scale, and literature reports only concern semi-enclosed systems such as the Taranto basins17,18,19 and the Aquatina lagoon20. No large-scale monitoring program on P. nobilis, in fact, has been carried out previously along the Apulian coast, although this kind of surveys is indispensable for the management of a protected species and must become mandatory for a critically endangered species such has become P. nobilis. The present data-gathering, that is aimed to partially address this information gap, based on the monitoring of recently dead specimens, allowed to realize a plausible map of P. nobilis distribution and densities before the MME in 30 areas distributed along the entire Apulian region coast.Along the Ionian coast, recently dead P. nobilis were detected in all the areas studied, highlighting a continuous distribution of the species prior to the MME, differently from the not continuous distribution along the Adriatic coast. The occurrence of P. nobilis was recorded in the areas surveyed in the south, from A7 to A17, but no traces were found along the northernmost areas except for the Tremiti archipelago, suggesting that the northernmost Adriatic coast of the region does not meet the environmental conditions suitable for hosting this species. Nevertheless, in the Gulf of Manfredonia multiple reports from fisherman indicating the presence of the species in a local Cymodocea nodosa meadow before the 1980s, suggest that this area may have been an exception in the past. Therefore, we can assume that excessive fishing and anthropogenic activities in this area are likely to have caused the species to disappear many decades ago.Data regarding the mortality incidence after the MME in Apulian populations is scarce. Panarese et al.11 reported the advent of the disease in Mar Piccolo di Taranto but without describing the disease incidence. In this study, a mortality incidence of 100% in all basins, bathymetric (down to 15 m) and habitat types, was recorded, demonstrating the severity of the situation along the entire Apulian coast, both inshore and offshore, and in lagoon and marine-protected areas.Although the availability of nutrients and the trophic conditions are assumed to be very different between offshore, inshore, and transitional systems, the archipelago of Tremiti islands, located 13 miles away from the coast, showed no differences in mortality incidence from sites along the coast, evidencing the same critical conditions in all environments.Many Mediterranean lagoon systems, including the Ebro Delta, Mar Menor Lagoon in Spain21, the Rhone delta, Leucate and Thau in France22,23,24, Venice, Grado-Marano and Faro in Italy25,26,27, Bizerte in Tunisia24 are considered the last healthy shelters for P. nobilis populations in the Mediterranean Sea22. These systems seem to offer a degree of resistance against the disease and are all characterized by high seasonal fluctuations of environmental parameters, such as temperature and salinity. It has been supposed that the effect of these fluctuations could make these environments less suitable for the spread of the disease and reduce the rate of transmission21,22. In the present study, two lagoon systems were also investigated, but no live specimens were found. These systems are strongly affected by the saltwater intrusion and the freshwater inputs became very low during the dry season. Hence, we can assume that during the summer season, when P. nobilis become susceptible to the disease, no salinity barrier against the pathogen spread persists in these lagoons systems.Considering that the lagoon refuges currently represent the main source of larval production for P. nobilis recruitment22,28, the collapse of these populations confirms the severity of the situation for species conservation. For the Italian coast, the last live populations are those in the lagoons located in the northen Adriatic Sea (Venice and Grado-Marano lagoon). These environments can act as larval exporters for the Adriatic Sea taking advantage of the mobility of the larvae that can spread over hundreds of kms28.Regarding the timeframe of the spread of the MME along the Apulian coast, the first report of the infection dates back to 201818, in the Mar Piccolo di Taranto. Compared to the first MME event observed in the Spanish coast in 20165,7, the disease has spread from the western to the eastern basin of the Mediterranean Sea over a period of 2 years. Our surveys, carried out in 2020, showed that 91% of the shells were still undamaged and with joined valves. Based on the state of conservation of the shells29 it is possible to hypothesize that the death of the specimens was a recent phenomenon that had occurred in Apulia in the two years preceding our surveys, and most probably it should be dated back to 2019.Kersting and Ballesteros30 have suggested that other species, such as P. rudis, could benefit from the collapse of the P. nobilis population. During our surveys, only 5 specimens of P. rudis were found, located in 2 sites, but it must be considered that the survey was carried out only a short time after the MME of P. nobilis. Further studies aimed at assessing an increase in P. rudis in the investigated areas would be of great interest to corroborate this hypothesis.In these surveys, P. nobilis showed transverse distribution among habitat types occurring both in marine and lagoon systems, inside and outside seagrass meadows, on sandy, rocky, and maerl beds substrate. Nevertheless, on a spatial macro (from a few kilometers to tens of kilometers) and mesoscale (from hundreds to thousands of meters), an overlap with the distributional range of seagrass meadows emerges. A clear cross-boundary subsidy trend was evidenced by the data collected on P. nobilis distribution in association with seagrasses. The specimens inside seagrass meadows were almost double than those detected nearby and a gradual decrease was observed with the increase of the distance from the seagrass patches (Fig. 2). This is particularly evident along the northern Adriatic coast of the region, where extended seagrass meadows are absent and, no trace of P. nobilis was encountered, except in the Tremiti archipelago where both P. oceanica meadows and pen shells were found. By contrast, present data reporting P. nobilis as associated with various seagrass species, such as P. oceanica, C. nodosa, and Zostera sp., are consistent with the macroscale and mesoscale association between P. nobilis and seagrass meadows sensu lato and most literature reporting ubiquitous distribution of P. nobilis both in lagoon-estuarine21,22,24,25,26,31 and in marine ecosystems4,7,9,14,16,24.However, regarding their microscale distribution, the pen shells in our surveys were recorded also outside the seagrass meadows boundaries, at times up to 1 km away. Hence, seagrass sheltering can potentially be ruled out as the sole explanatory factor for the distribution pattern of the species. The pattern emerging from this study led us to hypothesize that a trophic link with the seagrass detritus food-chain may explain both the macroscale–mesoscale association with seagrass species and the microscale cross-boundary distribution. In fact, seagrass detritus is highly refractory, since it is largely exported to the nearby areas where it can represent the major food source for other invertebrates32,33,34. This hypothesis is consistent with the stomach contents observations reported by Davenport et al.3 indicating detritus as the bulk component, accounting for 95% of the total ingested material.One of the main factors underlying the distribution pattern in benthic invertebrates is indeed food availability35,36. According to the Ideal Free Distribution (IFD) theory, the individuals in a population disperse to different resource patches within their environment, minimizing competition and maximizing fitness37. When the IFD assumptions are met, the number of individuals who aggregate in patches is proportional to the amount of food resource available in each one. Accordingly, the distribution of large, long life, and sessile organisms such as P. nobilis would be expected to depict the species trophic supply, by analyzing the resources available in those patches.Studies on the seagrass system energy flow have shown that seagrass debris must be fractionated before entering the food chain33. In this way, plant material becomes fine particulates moving in the boundary layer over the sediment–water interface38,39. These processes take time, and while the matter is transported, heterotrophic bacteria grow exponentially, turning it into a high quality and protein-enriched food for consumers. Hence, bacteria adhering to seagrass detritus may play a key role in this benthic food chain and sediment–water interface consumers may incorporate more energy from associated microbes than from the detritus itself32,38. On the basis of these considerations, it is reasonable to hypothesize that the quantity, composition and origin of the suspended particles are regulated by a drift mechanism and that this mechanism may explain local densities of P. nobilis as a response to sinking rates and resuspension effects. This hypothesis explain also the species distribution in systems, characterized by strong dominant current and shallow seabeds where the seagrass detritus can be spread/drift several kilometers away from the meadows. An example of this condition is encountered in the north Adriatic Sea (e.g., Gulf of Trieste) where extensive population of P. nobilis develops on several sink areas even kilometers downstream from the meadows. The assumption of the species’ ability to feed on seagrass detritus, together with the high biomasses reached (large size specimens and high density), lead us to suppose that P. nobilis may play a key role in the processing of matter and in the energy pathway deriving from seagrass detritus in Mediterranean coastal areas. This makes the repercussions of the MME not only a problem of conservation, but also and above all, an ecological-functional issue.We can, therefore, conclude that Mediterranean seagrass meadows not only constitute a habitat for P. nobilis, but probably also a food source through refractory detritus generation which is transferred and transformed outside the meadows. Unfortunately, literature is lacking on this topic and further investigations are needed to define the trophic role and function of these filter feeders in the different seagrass meadows.The density values that emerged were significantly different among basins. In the Adriatic Sea, where all the coastal values were recorded, the densities were consistently lower than those reported in the Ionian Sea, except for the two southernmost areas. In the Adriatic basin, it was also possible to recognize a north-south trend when considering the densities of pen shells in the coastal areas. Although the values recorded along the southern coast of the region were much greater than those recorded in the central coast, they were far lower than those reported by Čižmek et al.40 in the Croatian coast (North Adriatic Sea). Similar values to ours within the same basin were reported by Celebicic et al.41 in Bosnian waters (0.12 individuals/100 m2).On the other hand, in the Ionian areas, the values recorded were consistently >0.1 individuals/100 m2. The values recorded in the Mar Grande di Taranto were higher than those reported by Centoducati et al.17 (0.1–0.7 ind/ha2). From interviews with fishermen, it emerged that illegal trawling in this area has strongly impacted the natural populations of the Mar Grande di Taranto, and a partial reduction of this activity, in recent years could explain the slight increase in density compared to the 2004 survey data17.In interpreting our data, it should be considered that the surveys were carried out employing an extensive sampling protocol conceived to assess wide surface densities on coastal areas investigating across several habitat types. Therefore, literature density values focused only on local areas or habitat patchiness that were not randomly selected must be contextualized when compared with these data. In addition, given the scale of the presented surveys, emphasis must be given to P. nobilis absence data of which the literature appears poor. Indeed, contrary to the data on presence, reliable absence data are difficult to obtain requiring much greater effort to rule out a rare occurrence42. The absence data obtained in this study derive from the merger of two different data types. The first come from the local ecological knowledge obtained from interviews with the local fishermen, which allowed us to confirm our data, excluding spot occurrences in the same areas. Furthermore the interviews allowed us to collect information on a historical series of species presence/absence in the areas, which was helpful to confirm local absence when no P. nobilis specimens were recorded in our surveys. The second derives from the complete vision of divers during the field surveys. Indeed the scuba diver’s view was at least 10 times wider than 50 cm from the side around the rope and hence, the perception of absence can be extended over a much larger surface area investigated. By merging these two sources of information, we can assume that the absence data collected in exhaustive and complete.In conclusion, this study investigated different basins, habitat types, and bathymetries along the Apulian coast. The shells spatial distribution that arise from this study allowed to obtain important information on the species trophic ecology. Indeed, the species distributional pattern showed a strong overlap with seagrass meadows on meso and macro geographical scale, however this was not the case on a micro scale. This result indicates that although there is a strong relationship between P. nobilis and seagrass meadows, it is not limited to the habitat patch but crosses the boundaries of seagrass. This result led us to hypothesize that the distribution of P. nobilis displays a trophic link through the cross-boundary subsidy occurring from seagrass meadows to the nearby habitat, by means of the refractory detrital pathway. However, further investigations taking into account other factors such as hydrodynamics, are needed to investigate this topic.No live specimens of P. nobilis were found in >800 km of coastal line, leading us to the conclusion that the coastal and lagoon population had totally collapsed in the region after the MME. The seriousness of the situation on the Apulian coasts, just as in the other Mediterranean ecoregions, indicates that the MME that began in 2016 is still in progress, and no local population can be considered safe. Given the gravity of the current situation, it is vital for species preservation to extend the survey across the entire Italian coast to gain a overall picture of the status of the P. nobilis population on a national scale. Indeed, other regions may reveal the existence of natural shelters, where live populations of P. nobilis may still persist. If this is the case, it is essential to identify and protect them in time. As already suggested by Kersting et al.9, this initiative should be conducted in parallel by all the nations of the Mediterranean basin to implement standard guidelines for the monitoring, protection, and recovery of this critically endangered species. More

  • in

    Host biology, ecology and the environment influence microbial biomass and diversity in 101 marine fish species

    Whitman, W. B., Coleman, D. C. & Wiebe, W. J. Prokaryotes: the unseen majority. Proc. Natl Acad. Sci. USA 95, 6578–6583 (1998).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Locey, K. J. & Lennon, J. T. Scaling laws predict global microbial diversity. Proc. Natl Acad. Sci. USA 113, 5970–5975 (2016).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Louca, S. et al. Function and functional redundancy in microbial systems. Nat. Ecol. Evol. 2, 936–943 (2018).PubMed 

    Google Scholar 
    Dinsdale, E. A. et al. Functional metagenomic profiling of nine biomes. Nature 452, 629–632 (2008).ADS 
    CAS 
    PubMed 

    Google Scholar 
    Nayfach, S. et al. A genomic catalog of Earth’s microbiomes. Nat. Biotechnol. 39, 499–509 (2021).CAS 
    PubMed 

    Google Scholar 
    Thompson, L. R. et al. A communal catalogue reveals Earth’s multiscale microbial diversity. Nature 551, 457–463 (2017).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Song, S. J. et al. Comparative analyses of vertebrate gut microbiomes reveal convergence between birds and bats. MBio 11, e02901-19 (2020).Ley, R. E. et al. Evolution of mammals and their gut microbes. Science 320, 1647–1651 (2008).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    McDonald, D. et al. American Gut: an open platform for citizen science microbiome research. mSystems 3, e00031-18 (2018).Minich, J. J. et al. Temporal, environmental, and biological drivers of the mucosal microbiome in a wild marine fish, Scomber japonicus. mSphere 5, e00401-20 (2020).Sullam, K. E. et al. Environmental and ecological factors that shape the gut bacterial communities of fish: a meta-analysis. Mol. Ecol. 21, 3363–3378 (2012).PubMed 

    Google Scholar 
    Youngblut, N. D. et al. Host diet and evolutionary history explain different aspects of gut microbiome diversity among vertebrate clades. Nat. Commun. 10, 2200 (2019).ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bosco, N. & Noti, M. The aging gut microbiome and its impact on host immunity. Genes Immun. 22, 289–303 (2021).PubMed 
    PubMed Central 

    Google Scholar 
    McKenzie, V. J. et al. The effects of captivity on the mammalian gut microbiome. Integr. Comp. Biol. 57, 690–704 (2017).PubMed 
    PubMed Central 

    Google Scholar 
    Groussin, M. et al. Unraveling the processes shaping mammalian gut microbiomes over evolutionary time. Nat. Commun. 8, 14319 (2017).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Ross, A. A., Müller, K. M., Weese, J. S. & Neufeld, J. D. Comprehensive skin microbiome analysis reveals the uniqueness of human skin and evidence for phylosymbiosis within the class Mammalia. Proc. Natl Acad. Sci. USA 115, E5786–E5795 (2018).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hobbie, J. E., Daley, R. J. & Jasper, S. Use of nuclepore filters for counting bacteria by fluorescence microscopy. Appl. Environ. Microbiol. 33, 1225–1228 (1977).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Prussin, A. J. 2nd, Garcia, E. B. & Marr, L. C. Total virus and bacteria concentrations in indoor and outdoor air. Environ. Sci. Technol. Lett. 2, 84–88 (2015).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Gomez, D., Sunyer, J. O. & Salinas, I. The mucosal immune system of fish: the evolution of tolerating commensals while fighting pathogens. Fish. Shellfish Immunol. 35, 1729–1739 (2013).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Lowrey, L., Woodhams, D. C., Tacchi, L. & Salinas, I. Topographical mapping of the Rainbow Trout (Oncorhynchus mykiss) microbiome reveals a diverse bacterial community with antifungal properties in the skin. Appl. Environ. Microbiol. 81, 6915–6925 (2015).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Minich, J. J. et al. Microbial ecology of Atlantic Salmon (Salmo salar) hatcheries: impacts of the built environment on fish mucosal microbiota. Appl. Environ. Microbiol. 86, 20 (2020).Minich, J. J. et al. Impacts of the marine hatchery built environment, water and feed on mucosal microbiome colonization across ontogeny in Yellowtail Kingfish, Seriola lalandi. Front. Mar. Sci. 0, 676731 (2021).Minich, J. J. et al. The Southern Bluefin Tuna mucosal microbiome is influenced by husbandry method, net pen location, and anti-parasite treatment. Front. Microbiol. 11, 2015 (2020).PubMed 
    PubMed Central 

    Google Scholar 
    Ruiz-Rodríguez, M. et al. Host species and body site explain the variation in the microbiota associated to wild sympatric Mediterranean teleost fishes. Microb. Ecol. 80, 212–222 (2020).PubMed 

    Google Scholar 
    Tarnecki, A. M., Burgos, F. A., Ray, C. L. & Arias, C. R. Fish intestinal microbiome: diversity and symbiosis unravelled by metagenomics. J. Appl. Microbiol. 123, 2–17 (2017).CAS 
    PubMed 

    Google Scholar 
    Liu, H. et al. The gut microbiome and degradation enzyme activity of wild freshwater fishes influenced by their trophic levels. Sci. Rep. 6, 24340 (2016).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Egerton, S., Culloty, S., Whooley, J., Stanton, C. & Ross, R. P. The gut microbiota of marine fish. Front. Microbiol. 9, 873 (2018).PubMed 
    PubMed Central 

    Google Scholar 
    Woodhams, D. C. et al. Host-associated microbiomes are predicted by immune system complexity and climate. Genome Biol. 21, 23 (2020).PubMed 
    PubMed Central 

    Google Scholar 
    Karachle, P. K. & Stergiou, K. I. Gut length for several marine fish: relationships with body length and trophic implications. Mar. Biodivers. Rec. 3, 1–10 (2010).Ghilardi, M. et al. Phylogeny, body morphology, and trophic level shape intestinal traits in coral reef fishes. Ecol. Evol. 11, 13218–13231 (2021).PubMed 
    PubMed Central 

    Google Scholar 
    Clements, K. D., Angert, E. R., Linn Montgomery, W. & Howard Choat, J. Intestinal microbiota in fishes: what’s known and what’s not. Mol. Ecol. 23, 1891–1898 (2014).PubMed 

    Google Scholar 
    Zhu, D., Delgado-Baquerizo, M., Ding, J., Gillings, M. R. & Zhu, Y.-G. Trophic level drives the host microbiome of soil invertebrates at a continental scale. Microbiome 9, 189 (2021).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Minich, J. J. et al. KatharoSeq enables high-throughput microbiome analysis from low-biomass samples. mSystems 3, e00218-17 (2018).Davis, C. Enumeration of probiotic strains: review of culture-dependent and alternative techniques to quantify viable bacteria. J. Microbiol. Methods 103, 9–17 (2014).CAS 
    PubMed 

    Google Scholar 
    Rastogi, G., Tech, J. J., Coaker, G. L. & Leveau, J. H. J. A PCR-based toolbox for the culture-independent quantification of total bacterial abundances in plant environments. J. Microbiol. Methods 83, 127–132 (2010).CAS 
    PubMed 

    Google Scholar 
    Stämmler, F. et al. Adjusting microbiome profiles for differences in microbial load by spike-in bacteria. Microbiome 4, 28 (2016).PubMed 
    PubMed Central 

    Google Scholar 
    Tourlousse, D. M. et al. Synthetic spike-in standards for high-throughput 16S rRNA gene amplicon sequencing. Nucleic Acids Res. 45, e23 (2017).PubMed 

    Google Scholar 
    Gloor, G. B., Macklaim, J. M., Pawlowsky-Glahn, V. & Egozcue, J. J. Microbiome datasets are compositional: and this is not optional. Front. Microbiol. 8, 2224 (2017).PubMed 
    PubMed Central 

    Google Scholar 
    Morton, J. T. et al. Establishing microbial composition measurement standards with reference frames. Nat. Commun. 10, 2719 (2019).ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Louca, S., Doebeli, M. & Parfrey, L. W. Correcting for 16S rRNA gene copy numbers in microbiome surveys remains an unsolved problem. Microbiome 6, 41 (2018).PubMed 
    PubMed Central 

    Google Scholar 
    Smith, N. C., Rise, M. L. & Christian, S. L. A comparison of the innate and adaptive immune systems in cartilaginous fish, ray-finned fish, and lobe-finned fish. Front. Immunol. 10, 2292 (2019).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Azam, F. et al. The ecological role of water-column microbes in the sea. Mar. Ecol. Prog. Ser. 10, 257–263 (1983).ADS 

    Google Scholar 
    Fuhrman, J. A., Cram, J. A. & Needham, D. M. Marine microbial community dynamics and their ecological interpretation. Nat. Rev. Microbiol. 13, 133–146 (2015).CAS 
    PubMed 

    Google Scholar 
    Chong-Seng, K. M., Mannering, T. D., Pratchett, M. S., Bellwood, D. R. & Graham, N. A. J. The influence of coral reef benthic condition on associated fish assemblages. PLoS ONE 7, e42167 (2012).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Yahel, G. et al. Fish activity: a major mechanism for sediment resuspension and organic matter remineralization in coastal marine sediments. Mar. Ecol. Prog. Ser. 372, 195–209 (2008).ADS 
    CAS 

    Google Scholar 
    Glover, C. N., Bucking, C. & Wood, C. M. The skin of fish as a transport epithelium: a review. J. Comp. Physiol. B 183, 877–891 (2013).CAS 
    PubMed 

    Google Scholar 
    León-Zayas, R., McCargar, M., Drew, J. A. & Biddle, J. F. Microbiomes of fish, sediment and seagrass suggest connectivity of coral reef microbial populations. PeerJ 8, e10026 (2020).PubMed 
    PubMed Central 

    Google Scholar 
    Hess, S., Wenger, A. S., Ainsworth, T. D. & Rummer, J. L. Exposure of clownfish larvae to suspended sediment levels found on the Great Barrier Reef: impacts on gill structure and microbiome. Sci. Rep. 5, 10561 (2015).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Sparagon, W. J. et al. Fine scale transitions of the microbiota and metabolome along the gastrointestinal tract of herbivorous fishes. Anim. Microbiome 4, 33 (2022).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Edward Stevens, C. & Hume, I. D. Comparative Physiology of the Vertebrate Digestive System (Cambridge University Press, 2004).Wilson, J. M. & Castro, L. F. C. Morphological diversity of the gastrointestinal tract in fishes. Fish Physiol. 1–55 https://doi.org/10.1016/s1546-5098(10)03001-3 (2010).Shirakashi, S. et al. Morphology and distribution of blood fluke eggs and associated pathology in the gills of cultured Pacific bluefin tuna, Thunnus orientalis. Parasitol. Int. 61, 242–249 (2012).PubMed 

    Google Scholar 
    Ogawa, K. & Fukudome, M. Mass mortality caused by Blood Fluke(Paradeontacylix) among Amberjack(Seriola dumeili) imported to Japan. Fish. Pathol. 29, 265–269 (1994).
    Google Scholar 
    Wilson, J. M. & Laurent, P. Fish gill morphology: inside out. J. Exp. Zool. 293, 192–213 (2002).PubMed 

    Google Scholar 
    Huang, Q. et al. Diversity of gut microbiomes in marine fishes is shaped by host-related factors. Mol. Ecol. 29, 5019–5034 (2020).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kohl, K. D., Amaya, J., Passement, C. A., Dearing, M. D. & McCue, M. D. Unique and shared responses of the gut microbiota to prolonged fasting: a comparative study across five classes of vertebrate hosts. FEMS Microbiol. Ecol. 90, 883–894 (2014).CAS 
    PubMed 

    Google Scholar 
    Lall, S. P. & Tibbetts, S. M. Nutrition, feeding, and behavior of fish. Vet. Clin. North Am. Exot. Anim. Pract. 12, 361–372 (2009). xi.PubMed 

    Google Scholar 
    Hacquard, S. et al. Microbiota and host nutrition across plant and animal kingdoms. Cell Host Microbe 17, 603–616 (2015).CAS 
    PubMed 

    Google Scholar 
    Day, R. D., German, D. P. & Tibbetts, I. R. Why can’t young fish eat plants? Neither digestive enzymes nor gut development preclude herbivory in the young of a stomachless marine herbivorous fish. Comp. Biochem. Physiol. Part B: Biochem. Mol. Biol. 158, 23–29 (2011).
    Google Scholar 
    Lim, S. J. & Bordenstein, S. R. An introduction to phylosymbiosis. Proc. Biol. Sci. 287, 20192900 (2020).PubMed 
    PubMed Central 

    Google Scholar 
    Brucker, R. M. & Bordenstein, S. R. The hologenomic basis of speciation: gut bacteria cause hybrid lethality in the genus Nasonia. Science 341, 667–669 (2013).ADS 
    CAS 
    PubMed 

    Google Scholar 
    Mazel, F. et al. Is host filtering the main driver of phylosymbiosis across the tree of life? mSystems 3, e00097-18 (2018).Ross, A. A., Rodrigues Hoffmann, A. & Neufeld, J. D. The skin microbiome of vertebrates. Microbiome 7, 79 (2019).PubMed 
    PubMed Central 

    Google Scholar 
    Javůrková, V. G. et al. Unveiled feather microcosm: feather microbiota of passerine birds is closely associated with host species identity and bacteriocin-producing bacteria. ISME J. 13, 2363–2376 (2019).PubMed 
    PubMed Central 

    Google Scholar 
    Doane, M. P. et al. The skin microbiome of elasmobranchs follows phylosymbiosis, but in teleost fishes, the microbiomes converge. Microbiome 8, 93 (2020).PubMed 
    PubMed Central 

    Google Scholar 
    Chiarello, M. et al. Skin microbiome of coral reef fish is highly variable and driven by host phylogeny and diet. Microbiome 6, 147 (2018).PubMed 
    PubMed Central 

    Google Scholar 
    Sylvain, F.-É. et al. Fish skin and gut microbiomes show contrasting signatures of host species and habitat. Appl. Environ. Microbiol. 86, e00789-20 (2020).Smith, C. C. R., Snowberg, L. K., Gregory Caporaso, J., Knight, R. & Bolnick, D. I. Dietary input of microbes and host genetic variation shape among-population differences in stickleback gut microbiota. ISME J. 9, 2515–2526 (2015).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Human Microbiome Project Consortium. Structure, function and diversity of the healthy human microbiome. Nature 486, 207–214 (2012).ADS 

    Google Scholar 
    Choat, J. H. & Clements, K. D. Vertebrate herbivores in marine and terrestrial environments: a nutritional ecology perspective. Annu. Rev. Ecol. Syst. 29, 375–403 (1998).
    Google Scholar 
    Sale, P. F. Reef fish communities: open nonequilibrial systems. In The Ecology of Fishes on Coral Reefs. 564–598. https://doi.org/10.1016/b978-0-08-092551-6.50024-6 (Academic Press Inc., San Diego, 1991).Reese, A. T. & Dunn, R. R. Drivers of microbiome biodiversity: a review of general rules, feces, and ignorance. MBio 9, e01294-18 (2018).Press, C. McL & Evensen, Ø. The morphology of the immune system in teleost fishes. Fish Shellfish Immunol. 9, 309–318 (1999).Koppang, E. O., Kvellestad, A. & Fischer, U. Fish mucosal immunity: gill. In Mucosal Health in Aquaculture (eds Beck, B. & Peatman, E.) 93–133. https://doi.org/10.1016/b978-0-12-417186-2.00005-4 (Elsevier Inc., 2015).Esteban, M. Á. & Cerezuela, R. Fish mucosal immunity: skin. In Mucosal Health in Aquaculture (eds Beck, B. & Peatman, E.) 67–92. https://doi.org/10.1016/b978-0-12-417186-2.00004-2 (Elsevier Inc., 2015).Song, S. J. et al. Preservation methods differ in fecal microbiome stability, affecting suitability for field studies. mSystems 1, e00021-16 (2016).Rabosky, D. L. et al. An inverse latitudinal gradient in speciation rate for marine fishes. Nature 559, 392–395 (2018).ADS 
    CAS 
    PubMed 

    Google Scholar 
    Love, M. S., Bizzarro, J. J., Maria Cornthwaite, A., Frable, B. W. & Maslenikov, K. P. Checklist of marine and estuarine fishes from the Alaska–Yukon Border, Beaufort Sea, to Cabo San Lucas, Mexico. Zootaxa 5053, 1–285 (2021).PubMed 

    Google Scholar 
    Allen, L. G. & Horn, M. H. The Ecology of Marine Fishes: California and Adjacent Waters (University of California Press, 2006).Al-Hussaini, A. H. On the functional morphology of the alimentary tract of some fish in relation to differences in their feeding habits; anatomy and histology. Q. J. Microsc. Sci. 90(Pt. 2), 109–139 (1949).PubMed 

    Google Scholar 
    Maddock, L., Bone, Q. & Rayner, J. M. V. (eds). In Mechanics and Physiology of Animal Swimming (Press Syndicate-of the University of Cambridge, 1994).Gonzalez, A. et al. Qiita: rapid, web-enabled microbiome meta-analysis. Nat. Methods 15, 796–798 (2018).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Cruz, G. N. F., Christoff, A. P. & de Oliveira, L. F. V. Equivolumetric protocol generates library sizes proportional to total microbial load in 16S amplicon sequencing. Front. Microbiol. 12, 638231 (2021).PubMed 
    PubMed Central 

    Google Scholar 
    Minich, J. J. et al. Quantifying and understanding well-to-well contamination in microbiome research. mSystems 4, e00186-19 (2019).Minich, J. J. et al. High-throughput miniaturized 16S rRNA amplicon library preparation reduces costs while preserving microbiome integrity. mSystems 3, e00166-18 (2018).Walters, W. et al. Improved bacterial 16S rRNA gene (V4 and V4–5) and fungal internal transcribed spacer marker gene primers for microbial community surveys. mSystems 1, e00009-15 (2016).Amir, A. et al. Deblur rapidly resolves single-nucleotide community sequence patterns. mSystems 2, e00191-16 (2017).Whittaker, R. J., Willis, K. J. & Field, R. Scale and species richness: towards a general, hierarchical theory of species diversity. J. Biogeogr. 28, 453–470 (2001).
    Google Scholar 
    Bolyen, E. et al. Reproducible, interactive, scalable and extensible microbiome data science using QIIME 2. Nat. Biotechnol. 37, 852–857 (2019).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Lozupone, C., Lladser, M. E., Knights, D., Stombaugh, J. & Knight, R. UniFrac: an effective distance metric for microbial community comparison. ISME J. 5, 169–172 (2011).PubMed 

    Google Scholar 
    McDonald, D. et al. Striped UniFrac: enabling microbiome analysis at unprecedented scale. Nat. Methods 15, 847–848 (2018).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Lozupone, C. & Knight, R. UniFrac: a new phylogenetic method for comparing microbial communities. Appl. Environ. Microbiol. 71, 8228–8235 (2005).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Janssen, S. et al. Phylogenetic placement of exact amplicon sequences improves associations with clinical information. mSystems 3, e00021-18 (2018).Anderson, M. J. A new method for non-parametric multivariate analysis of variance. Austral Ecol. 26, 32–46 (2001).
    Google Scholar 
    Minich, J. J. et al. Microbial effects of livestock manure fertilization on freshwater aquaculture ponds rearing tilapia (Oreochromis shiranus) and North African catfish (Clarias gariepinus). Microbiologyopen 7, e00716 (2018).PubMed 
    PubMed Central 

    Google Scholar 
    Van Doan, H. et al. Host-associated probiotics: a key factor in sustainable aquaculture. Rev. Fish. Sci. Aquac. 28, 16–42 (2020).
    Google Scholar 
    Kumar, S., Stecher, G., Suleski, M. & Hedges, S. B. TimeTree: a resource for timelines, timetrees, and divergence times. Mol. Biol. Evol. 34, 1812–1819 (2017).CAS 
    PubMed 

    Google Scholar  More

  • in

    Senescence of the immune defences and reproductive trade-offs in females of the mealworm beetle, Tenebrio molitor

    Williams, G. C. Natural selection, the costs of reproduction, and a refinement of lack’s principle. Am. Nat. 100, 687–690 (1966).
    Google Scholar 
    Stearns, S. C. The evolution of life histories. (Oxford University Press, 1992).Kirkwood, T. B. L. Evolution of ageing. Nature 270, 301–304 (1977).ADS 
    CAS 
    PubMed 

    Google Scholar 
    Partridge, L., Prowse, N. & Pignatelli, P. Another set of responses and correlated responses to selection on age at reproduction in Drosophila melanogaster. Proc. R. Soc. B. 266, 255–261 (1999).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Metcalfe, N. Growth versus lifespan: Perspectives from evolutionary ecology. Exp. Gerontol. 38, 935–940 (2003).PubMed 

    Google Scholar 
    Lee, W.-S., Monaghan, P. & Metcalfe, N. B. Experimental demonstration of the growth rate–lifespan trade-off. Proc. R. Soc. B. 280, 20122370 (2013).PubMed 
    PubMed Central 

    Google Scholar 
    Lemaître, J.-F. et al. Early-late life trade-offs and the evolution of ageing in the wild. Proc. R. Soc. B. 282, 20150209 (2015).PubMed 
    PubMed Central 

    Google Scholar 
    Jehan, C., Sabarly, C., Rigaud, T. & Moret, Y. Late-life reproduction in an insect: Terminal investment, reproductive restraint or senescence. J. Anim. Ecol. 90, 282–297 (2021).PubMed 

    Google Scholar 
    Pawelec, G. Age and immunity: What is “immunosenescence”?. Exp. Gerontol. 105, 4–9 (2018).CAS 
    PubMed 

    Google Scholar 
    Schwenke, R. A., Lazzaro, B. P. & Wolfner, M. F. Reproduction–immunity trade-offs in insects. Annu. Rev. Entomol. 61, 239–256 (2016).CAS 
    PubMed 

    Google Scholar 
    Maklakov, A. A. & Chapman, T. Evolution of ageing as a tangle of trade-offs: Energy versus function. Proc. R. Soc. B. 286, 20191604 (2019).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Hamel, S. et al. Fitness costs of reproduction depend on life speed: empirical evidence from mammalian populations: Fitness costs of reproduction in mammals. Ecol. Lett. 13, 915–935 (2010).PubMed 

    Google Scholar 
    Graham, A. L., Allen, J. E. & Read, A. F. Evolutionary causes and consequences of immunopathology. Annu. Rev. Ecol. Evol. Syst. 36, 373–397 (2005).
    Google Scholar 
    Sorci, G. & Faivre, B. Inflammation and oxidative stress in vertebrate host–parasite systems. Phil. Trans. R. Soc. B. 364, 71–83 (2009).PubMed 

    Google Scholar 
    Ashley, N. T., Weil, Z. M. & Nelson, R. J. Inflammation: Mechanisms, costs, and natural variation. Annu. Rev. Ecol. Evol. Syst. 43, 385–406 (2012).
    Google Scholar 
    Babin, A., Moreau, J. & Moret, Y. Storage of carotenoids in crustaceans as an adaptation to modulate immunopathology and optimize immunological and life history strategies. BioEssays 41, 1800254 (2019).
    Google Scholar 
    Vasto, S. et al. Inflammatory networks in ageing, age-related diseases and longevity. Mech. Ageing Dev. 128, 83–91 (2007).CAS 
    PubMed 

    Google Scholar 
    Finch, C. E. & Crimmins, E. M. Inflammatory exposure and historical changes in human life-spans. Science 305, 1736–1739 (2004).ADS 
    CAS 
    PubMed 

    Google Scholar 
    Licastro, F. et al. Innate immunity and inflammation in ageing: A key for understanding age-related diseases. Immun. Ageing 2, 8 (2005).PubMed 
    PubMed Central 

    Google Scholar 
    Pawelec, G., Goldeck, D. & Derhovanessian, E. Inflammation, ageing and chronic disease. Curr. Opin. Immunol. 29, 23–28 (2014).CAS 
    PubMed 

    Google Scholar 
    Pursall, E. R. & Rolff, J. Immune responses accelerate ageing: Proof-of-principle in an insect model. PLoS ONE 6, e19972 (2011).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Khan, I., Agashe, D. & Rolff, J. Early-life inflammation, immune response and ageing. Proc. R. Soc. B. 284, 20170125 (2017).PubMed 
    PubMed Central 

    Google Scholar 
    Vigneron, A., Jehan, C., Rigaud, T. & Moret, Y. Immune defenses of a beneficial pest: The mealworm beetle, Tenebrio molitor. Front. Physiol. 10, 138 (2019).PubMed 
    PubMed Central 

    Google Scholar 
    Jehan, C., Chogne, M., Rigaud, T. & Moret, Y. Sex-specific patterns of senescence in artificial insect populations varying in sex-ratio to manipulate reproductive effort. BMC Evol. Biol. 20, 18 (2020).PubMed 
    PubMed Central 

    Google Scholar 
    Jehan, C., Sabarly, C., Rigaud, T. & Moret, Y. Age-specific fecundity under pathogenic threat in an insect: Terminal investment versus reproductive restraint. J. Anim. Ecol. 91, 101–111 (2022).PubMed 

    Google Scholar 
    Chung, K.-H. & Moon, M.-J. Fine structure of the hemopoietic tissues in the mealworm beetle, Tenebrio molitor. Entomol. Res. 34, 131–138 (2004).
    Google Scholar 
    Urbański, A., Adamski, Z. & Rosiński, G. Developmental changes in haemocyte morphology in response to Staphylococcus aureus and latex beads in the beetle Tenebrio molitor L.. Micron 104, 8–20 (2018).PubMed 

    Google Scholar 
    Vommaro, M. L., Kurtz, J. & Giglio, A. Morphological characterisation of haemocytes in the mealworm beetle Tenebrio molitor (Coleoptera, Tenebrionidae). Insects 12, 423 (2021).PubMed 
    PubMed Central 

    Google Scholar 
    Söderhäll, K. & Cerenius, L. Role of the prophenoloxidase-activating system in invertebrate immunity. Curr. Opin. Immunol. 10, 23–28 (1998).PubMed 

    Google Scholar 
    Siva-Jothy, M. T., Moret, Y. & Rolff, J. Insect immunity: an evolutionary ecology perspective. in Advances in Insect Physiology vol. 32 1–48 (Elsevier, 2005).Nappi, A. J. & Ottaviani, E. Cytotoxicity and cytotoxic molecules in invertebrates. BioEssays 22, 469–480 (2000).CAS 
    PubMed 

    Google Scholar 
    Sadd, B. M. & Siva-Jothy, M. T. Self-harm caused by an insect’s innate immunity. Proc. R. Soc. B. 273, 2571–2574 (2006).PubMed 
    PubMed Central 

    Google Scholar 
    Daukšte, J., Kivleniece, I., Krama, T., Rantala, M. J. & Krams, I. Senescence in immune priming and attractiveness in a beetle: Immunosenescence in a beetle. J. Evol. Biol. 25, 1298–1304 (2012).PubMed 

    Google Scholar 
    Krams, I. et al. Trade-off between cellular immunity and life span in mealworm beetles Tenebrio molitor. Curr. Zool. 59, 340–346 (2013).
    Google Scholar 
    Moon, H. J., Lee, S. Y., Kurata, S., Natori, S. & Lee, B. L. Purification and molecular cloning of cDNA for an inducible antibacterial protein from larvae of the coleopteran, Tenebrio molitor. J. Biochem. 116, 53–58 (1994).CAS 
    PubMed 

    Google Scholar 
    Lee, Y. J. et al. Structure and expression of the tenecin 3 gene in Tenebrio molitor. Biochem. Biophys. Res. Comm. 218, 6–11 (1996).CAS 
    PubMed 

    Google Scholar 
    Kim, D. H. et al. Bacterial expression of tenecin 3, an insect antifungal protein isolated from Tenebrio molitor, and its efficient purification. Mol. Cells 8, 786–789 (1998).CAS 
    PubMed 

    Google Scholar 
    Roh, K.-B. et al. Proteolytic cascade for the activation of the insect toll pathway induced by the fungal cell wall component. J. Biol. Chem. 284, 19474–19481 (2009).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Park, J.-W. et al. Beetle Immunity. in Invertebrate Immunity (ed. Söderhäll, K.) vol. 708 163–180 (Springer US, 2010).Chae, J.-H. et al. Purification and characterization of tenecin 4, a new anti-Gram-negative bacterial peptide, from the beetle Tenebrio molitor. Dev. Comp. Immunol. 36, 540–546 (2012).CAS 
    PubMed 

    Google Scholar 
    Haine, E. R., Pollitt, L. C., Moret, Y., Siva-Jothy, M. T. & Rolff, J. Temporal patterns in immune responses to a range of microbial insults (Tenebrio molitor). J. Insect Physiol. 54, 1090–1097 (2008).CAS 
    PubMed 

    Google Scholar 
    Dhinaut, J., Chogne, M. & Moret, Y. Immune priming specificity within and across generations reveals the range of pathogens affecting evolution of immunity in an insect. J. Anim. Ecol. 87, 448–463 (2018).PubMed 

    Google Scholar 
    Hoffmann, J. A., Reichhart, J.-M. & Hetru, C. Innate immunity in higher insects. Curr. Opin. Immunol. 8, 8–13 (1996).CAS 
    PubMed 

    Google Scholar 
    Moret, Y. Explaining variable costs of the immune response: selection for specific versus non-specific immunity and facultative life history change. Oikos 102, 213–216 (2003).
    Google Scholar 
    Khan, I., Prakash, A. & Agashe, D. Immunosenescence and the ability to survive bacterial infection in the red flour beetle Tribolium castaneum. J. Anim. Ecol. 85, 291–301 (2016).PubMed 

    Google Scholar 
    Rolff, J. Effects of age and gender on immune function of dragonflies (Odonata, Lestidae) from a wild population. Can. J. Zool. 79, 2176–2180 (2001).
    Google Scholar 
    Doums, C., Moret, Y., Benelli, E. & Schmid-Hempel, P. Senescence of immune defence in Bombus workers. Ecol. Entomol. 27, 138–144 (2002).
    Google Scholar 
    Schmid, M. R., Brockmann, A., Pirk, C. W. W., Stanley, D. W. & Tautz, J. Adult honeybees (Apis mellifera L.) abandon hemocytic, but not phenoloxidase-based immunity. J. Insect Physiol. 54, 439–444 (2008).CAS 
    PubMed 

    Google Scholar 
    Moret, Y. & Schmid-Hempel, P. Immune responses of bumblebee workers as a function of individual and colony age: senescence versus plastic adjustment of the immune function. Oikos 118, 371–378 (2009).
    Google Scholar 
    Armitage, S. A. O. & Boomsma, J. J. The effects of age and social interactions on innate immunity in a leaf-cutting ant. J. Insect Physiol. 56, 780–787 (2010).CAS 
    PubMed 

    Google Scholar 
    Korner, P. & Schmid-Hempel, P. In vivo dynamics of an immune response in the bumble bee Bombus terrestris. J. Invert. Pathol. 87, 59–66 (2004).CAS 

    Google Scholar 
    Li, T., Yan, D., Wang, X., Zhang, L. & Chen, P. Hemocyte changes during immune melanization in Bombyx Mori infected with Escherichia coli. Insects 10, 301 (2019).PubMed Central 

    Google Scholar 
    Chase, M. R., Raina, K., Bruno, J. & Sugumaran, M. Purification, characterization and molecular cloning of prophenoloxidases from Sarcophaga bullata. Insect Biochem. Mol. Biol. 30, 953–967 (2000).CAS 
    PubMed 

    Google Scholar 
    Kanost, M. R. & Gorman, M. J. Phenoloxidases in insect immunity. in Insect Immunology 69–96 (Elsevier, 2008).Sadd, B. M. et al. Modulation of sexual signalling by immune challenged male mealworm beetles (Tenebrio molitor L.): Evidence for terminal investment and dishonesty. J. Evol. Biol. 19, 321–325 (2006).CAS 
    PubMed 

    Google Scholar 
    Gálvez, D. & Chapuisat, M. Immune priming and pathogen resistance in ant queens. Ecol. Evol. 4, 1761–1767 (2014).PubMed 
    PubMed Central 

    Google Scholar 
    Armitage, S. A. O. & Siva-Jothy, M. T. Immune function responds to selection for cuticular colour in Tenebrio molitor. Heredity 94, 650–656 (2005).CAS 
    PubMed 

    Google Scholar 
    Armitage, S. A. O., Thompson, J. J. W., Rolff, J. & Siva-Jothy, M. T. Examining costs of induced and constitutive immune investment in Tenebrio molitor. J. Evol. Biol. 16, 1038–1044 (2003).CAS 
    PubMed 

    Google Scholar 
    Kokoza, V. A. et al. Transcriptional regulation of the mosquito vitellogenin gene via a blood meal-triggered cascade. Gene 274, 47–65 (2001).CAS 
    PubMed 

    Google Scholar 
    Isaac, P. G. & Bownes, M. Ovarian and fat-body vitellogenin synthesis in Drosophila melanogaster. Europ. J. Biochem. 123, 527–534 (2005).
    Google Scholar 
    Hoffmann, J. A. The immune response of Drosophila. Nature 426, 33–38 (2003).ADS 
    CAS 
    PubMed 

    Google Scholar 
    Tzou, P. et al. Tissue-specific inducible expression of antimicrobial peptide genes in Drosophila surface epithelia. Immunity 13, 737–748 (2000).CAS 
    PubMed 

    Google Scholar 
    Haine, E. R., Moret, Y., Siva-Jothy, M. T. & Rolff, J. Antimicrobial defense and persistent infection in insects. Science 322, 1257–1259 (2008).ADS 
    CAS 
    PubMed 

    Google Scholar 
    Moret, Y. & Siva-Jothy, M. T. Adaptive innate immunity? Responsive-mode prophylaxis in the mealworm beetle, Tenebrio molitor. Proc. R. Soc. B. 270, 2475–2480 (2003).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Du Rand, N. & Laing, M. D. Determination of insecticidal toxicity of three species of entomopathogenic spore-forming bacterial isolates against Tenebrio molitor L. (Coleoptera: Tenebrionidae). Afr. J. Microbiol. Res. 5, 2222–2228 (2011).
    Google Scholar 
    Jurat-Fuentes, J. L. & Jackson, T. Bacterial entomopathogens. In Insect Pathology 2nd edn (eds Kaya, H. & Vera, F.) 265–349 (Elsevier Academic Press, Cambridge, Mass, 2012).
    Google Scholar 
    Dhinaut, J., Balourdet, A., Teixeira, M., Chogne, M. & Moret, Y. A dietary carotenoid reduces immunopathology and enhances longevity through an immune depressive effect in an insect model. Sci. Rep. 7, 12429 (2017).ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Moreau, J., Martinaud, G., Troussard, J.-P., Zanchi, C. & Moret, Y. Trans-generational immune priming is constrained by the maternal immune response in an insect. Oikos 121, 1828–1832 (2012).
    Google Scholar 
    Lee, H. S. et al. The pro-phenoloxidase of coleopteran insect, Tenebrio molitor, larvae was activated during cell clump/cell adhesion of insect cellular defense reactions. FEBS Lett. 444, 255–259 (1999).CAS 
    PubMed 

    Google Scholar 
    Zanchi, C., Troussard, J.-P., Martinaud, G., Moreau, J. & Moret, Y. Differential expression and costs between maternally and paternally derived immune priming for offspring in an insect. J. Anim. Ecol. 80, 1174–1183 (2011).PubMed 

    Google Scholar 
    Moret, Y. ‘Trans-generational immune priming’: Specific enhancement of the antimicrobial immune response in the mealworm beetle, Tenebrio molitor. Proc. R. Soc. B. 273, 1399–1405 (2006).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Dubuffet, A. et al. Trans-generational immune priming protects the eggs only against gram-positive bacteria in the mealworm beetle. PLoS Pathog. 11, e1005178 (2015).PubMed 
    PubMed Central 

    Google Scholar  More

  • in

    Health risk assessment and source apportionment of potentially toxic metal(loid)s in windowsill dust of a rapidly growing urban settlement, Iran

    PTM concentrationsIn Table 1, the descriptive statistics of PTMs in 50 dust samples from Qom city are described. The background values were used based on the concentrations of metals in the Upper Continental Crust. The mean concentration of As, Cd, Cu, Mo, Pb, Sb, and, Zn exceeded the background value. Also, Cd, Cu, Mo, Pb, Sb, and, Zn had a coefficient of variation (C.V.) greater than 50%, indicating a severe variability in PTMs concentrations in the atmospheric dust of the studied area2. Metals with C.V.  Pb  > Zn  > As  > Cd  > Cu  > Mo  > Cr  > Mn  > Ni = Co. Antimony (38.55) and Pb (35.13) had the highest average EF values, which means they were enriched very high in the windowsill dust. Also, they had a wide range of EF values in the 50 stations: from 4.0 to 227.0 for Sb, and from 8.3 to 140.8 for Pb which might reflect the existence of discrete multiple sources in the studied area. The degree of enrichment for Pb and Sb in the industrial sector was extreme and in the commercial sector was very high; also, the other sectors were significantly enriched. Zinc and As had a more homogenous enrichment in the area. In all the functional sectors, 95% and 84% of stations were significantly enriched by As and Zn, respectively. Copper, Cd, and Mo were moderately enriched in all functional sectors, but the greenspace sector had minimal enrichment by these elements. Some areas in the industrial sector had significant to very high enrichment of Cd. The EF value indicated Co, Cr, Mn, and Ni were minimally enriched in all the stations.Figure 2Box plot of the (a) enrichment factor (EF), and (b) geo accumulation index (Igeo) for the dust samples in the studied area.Full size imageThe highest average values of Igeo were obtained in the order of Pb  > Sb  > As  > Zn. PTMs included Co, Cr, Ni, Fe, and Mn were categorized as unpolluted and Cd, Cu, and Mo were in the category of unpolluted to moderately polluted. In the industrial zone, the windowsill dust was extremely polluted with Sb and Pb. The sequence of contamination intensity with Pb, Zn and Sb according to land use was: industrial  > commercial  > residential  > greenspace. The highest concentration of arsenic in the study area belongs to the industrial area.To evaluate the pollution level based on land use, PLI and mCd indices were utilized (Fig. 3). These cumulative indices showed that the dust in Qom city is considerably contaminated with PTMs. According to the PLI index, all the stations were categorized as polluted sites. The PLIzone values were in the order of industrial (3.77)  > commercial (2.05)  > residential (1.67)  > green space (1.38). This pattern was also repeated with the mCd index. The mCd for the industrial sector ranged from 6.98 (high contamination level) to 39.60 (ultra-high contamination level). In the commercial sector, fifty percent of dust samples were classified as having a high degree of contamination. All the greenspace stations were in the moderate pollution category. This shows the possible effect of tree density in diminishing the risk of dust pollution to the receptors.Figure 3Pollution level indexes (a) mCd and (b) PLI, based on four functional areas.Full size imageSpatial distribution of PTMsThe As, Cd, Cu, Sb, Pb, and Zn content in 100% of the dust samples exceeded the background value. Spatial distribution maps were generated for the hotspot PTMs (As, Sb, Pb, Cd, Cu, Mo and Zn) by applying the inverse distance weighted (IDW) interpolation method (ArcGIS 10.3). Figure 4 demonstrates that PTMs dispersions were slightly influenced by the prevailing wind direction (from the west), suggesting they came from the point- or area- sources. On the other hand, the K–S test showed that the overall distribution of PTMs was not normal in the studied region. This might signify the influence of industrial activities and the presence of multiple sources of dust.Figure 4Spatial distribution maps of seven PTMs in windowsill dusts of Qom, Iran. This map was constructed using ArcGIS version 10.3. (https://www.esri.com/en-us/arcgis/products/arcgisdesktop/overview).Full size imageThe highest pollution load of PTMs belonged to the industrial section. The level of pollution gradually decreased from Shokouhieh to Mahmoudabad industrial zones. The reason is related to more active industries, a closed environment, and more construction existing in Shokouhieh industrial town than in Mahmoudabad industrial town.There is a clear decreasing trend from the central part to southern (downtown area) and southwestern (suburb area) parts of the city. In fact, these parts are diffusely populated and the southwestern part is almost new with lots of barren lands. Copper, Mo, and Cd show high concentrations toward the central part of the city. Educational, cultural and commercial activities are mainly located in the central part of the city. Also, historical and religious districts in the city center are accompanied by a huge influx of tourists throughout the year. For this reason, the central part of the city has various public transportations such as bus stands and taxi stations, and is dominated by a high load of motorcycles.In the eastern part of the city, some hotspots can be observed (Fig. 4). This part includes an important transportation system (like highways and a complex interchange) where exhaust traffic emissions might be a probable source of As, Sb, Pb, and Zn. Unlike Pb and Zn, several peaks of As are scattered in the western part, suggesting an area source might exist in the region. It is noteworthy that the western area is densely populated with lots of residential buildings. Bisht et al. (2022)35 also observed hotspots of As in the residential area of Dehradun, India.PTM potential sourcesTo evaluate the relationship between PTMs in dust samples, the Spearman correlation and PCA were developed (Fig. 5) and more details are given in Table S7. Statistical analysis can help to identify the potential source of contamination in urban dust. The Spearman correlation was significant at p  residential ≈ greenspace. The five PTMs with the highest overall HI are ranked as follows: Pb  > As  > Cr  > Mn  > Sb (Fig. 7). The HI values in all the sections were lower than the permissible level (1.00), except for Pb. In the industrial section, Pb recorded the highest HI value for children (HI = 1.73) which exceeded the acceptable value. The HI values were 10 times higher for children than adults indicating they are more susceptible to PTMs in the dust.The dominant pathway for noncancerous risk was ingestion followed by dermal contact and inhalation. The trend is in line with previous research25,51,52. However, for Co and Mn, the descending order was different as follows ingestion  > inhalation  > dermal contact. The highest contribution of HQinh and HQderm to HI was measured for Co (34.0%) and Cd (29.0%), respectively.In this study, the carcinogenic risk from windowsill dust was estimated for the carcinogens including Cd, Co, Cr and Ni, Pb, and As through the possible routes (Fig. 8, Table S9). The contribution of PTMs to CR decreased in the order of Cr (3.24E−05)  > As (2.05E−05)  > Pb (2.52E−06)  > Co (6.91E−09)  > Ni (1.72E−09)  > Cd (2.58E−10). The average CR values for target PTMs through inhalation ranged from 7.9E−10 to 1.7E−07, which remained in the safety zone (CR  inhalation  > dermal (Fig. 8). While the contribution of Cr to carcinogenic risk was higher through inhalation than ingestion. The reports concluded that the primary exposure route of Cr is inhalation54. Considering the predominant forms of Cr in the environment, CrVI is more toxic than CrIII. Exposure to CrVI can cause immunological diseases, dental effects and carcinogenic effects (lung cancer, nose and nasal sinus cancer, suspected laryngeal and stomach cancers)54,55.The result of health risk from target PTMs in windowsills of Qom indicates significant chronic exposure to Pb can take place for children in the industrial zone. The ingestion route is the most probable pathway for children due to their hand–to–mouth behavior56. Lead can bio-accumulate in the body without any obvious symptoms of toxicity56. The total CR values for Pb, Cr and As in different land-use types were in the range of tolerable carcinogenic risk (1 × 10−4  More