More stories

  • in

    Population structure of blackfin tuna (Thunnus atlanticus) in the western Atlantic Ocean inferred from microsatellite loci

    Carvalho, G. R. & Hauser, L. Molecular genetics and the stock concept in fisheries. In Molecular Genetics in Fisheries (eds Carvalho, G. R. & Pitcher, T. J.) 55–79 (Springer Netherlands, 1995). https://doi.org/10.1007/978-94-011-1218-5_3.Chapter 

    Google Scholar 
    Avise, J. C. Conservation genetics in the marine realm. J. Hered. 89, 377–382 (1998).Article 

    Google Scholar 
    Waples, R. S. Separating the wheat from the chaff: Patterns of genetic differentiation in high gene flow species. J. Hered. 89, 438–450 (1998).Article 

    Google Scholar 
    Pecoraro, C. et al. The population genomics of yellowfin tuna (Thunnus albacares) at global geographic scale challenges current stock delineation. Sci. Rep. 8, 13890 (2018).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Nikolic, N. et al. Connectivity and population structure of albacore tuna across southeast Atlantic and southwest Indian Oceans inferred from multidisciplinary methodology. Sci. Rep. 10, 15657 (2020).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Anderson, G., Lal, M., Hampton, J., Smith, N. & Rico, C. Close kin proximity in yellowfin tuna (Thunnus albacares) as a driver of population genetic structure in the tropical western and central Pacific Ocean. Front. Mar. Sci. 6, 341 (2019).Article 

    Google Scholar 
    Collette, B. B. & Nauen, C. E. Scombrids of the World: An Annotated and Illustrated Catalogue of Tunas, Mackerels, Bonitos, and Related Species Known to date v.2 (FAO, 1983).
    Google Scholar 
    Majkowski, J., Arrizabalaga, H. & Carocci, F. C1. Tuna and Tuna-like Species. Review of the state of World Fisheries Resources (FAO, 2005).Mahon, R. Fisheries and research for tunas and tuna-like species in the Western Central Atlantic: implications of the agreement for the implementation of the provisions of the United Nations Convention on the Law of the Sea of the 10 December 1982 relating to the conservation and management of straddling fish stocks and highly migratory fish stocks. (FAO Fisheries Technical Paper, 1996).Doray, M., Stéquert, B. & Taquet, M. Age and growth of blackfin tuna (Thunnus atlanticus ) caught under moored fish aggregating devices, around Martinique Island. Aquat. Living Resour. 17, 13–18 (2004).Article 

    Google Scholar 
    Arocha, F., Barrios, A. & Marcano, J. Blackfin tuna (Thunnus atlanticus) in the Venezuelan fisheries. Collect. Vol. Sci. Pap ICCAT 68(3), 1253–1260 (2012).
    Google Scholar 
    Mathieu, H., Pau, C. & Reynal, L. Chapter 2.1.10.7 THON A NAGEOIRES NOIRES. ICCAT ICCAT Manual. International Commission for the Conservation of Atlantic Tuna. 15 (2013).Maghan, W. B. & Rivas, L. R. The blackfin tuna (Thunnus atlanticus) as an underutilized fishery resource in the tropical western Atlantic Ocean. FAO Fish. Rep. 71(2), 163–172 (1971).
    Google Scholar 
    De Sylva, D. P., Rathjen, W. F. & Higman, J. B. Fisheries development for underutilized Atlantic tunas: Blackfin and little tunny. NOAA Technical Memorandum NMFS-SEFC-191 (1987).Richardson, D. E., Llopiz, J. K., Guigand, C. M. & Cowen, R. K. Larval assemblages of large and medium-sized pelagic species in the Straits of Florida. Prog. Oceanogr. 86, 8–20 (2010).ADS 
    Article 

    Google Scholar 
    Freire, K. M. F., Lessa, R. & Lins-Oliveira, J. E. Fishery and biology of blackfin tuna Thunnus atlanticus off northeastern Brazil. Gulf Caribb. Res. 17, 15–24 (2005).Article 

    Google Scholar 
    Vieira, K. R., Oliveira, J. E. L. & Barbalho, M. C. Aspects of the dynamic population of blackfin tuna (Thunnus atlanticus-Lesson, 1831) caught in the Northeast Brazil. Collect. Vol. Sci. Pap ICCAT 58(5), 1623–1628 (2005).
    Google Scholar 
    FJ Mather, I. I. I. Tunas (genus Thunnus) of the western North Atlantic. Part III. Distribution and behavior of Thunnus species. World Sci. Meeting Biol. Tunas Exper. Pap. Vol. 8, 1–23 (1962)Cornic, M. & Rooker, J. R. Influence of oceanographic conditions on the distribution and abundance of blackfin tuna (Thunnus atlanticus) larvae in the Gulf of Mexico. Fish. Res 201, 1–10 (2018).Article 

    Google Scholar 
    Block, B. A. et al. Electronic tagging and population structure of Atlantic bluefin tuna. Nature 434, 1121–1127 (2005).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Luckhurst, B. E., Trott, T. & Manuel, S. Landings, seasonality, catch per unit effort, and tag-recapture results of yellowfin tuna and blackfin tuna at Bermuda. Am. Fish. Soc. Symp. 25, 225–234 (2001).
    Google Scholar 
    Singh-Renton, S. & Renton, J. CFRAMP’s large pelagic fish tagging program. Gulf Caribb. Res. Vol 19, (2007).Cermeño, P. et al. Electronic tagging of Atlantic bluefin tuna (Thunnus thynnus, L.) reveals habitat use and behaviors in the Mediterranean Sea. PLoS ONE 10, e0116638 (2015).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Begg, G. A., Friedland, K. D. & Pearce, J. B. Stock identification and its role in stock assessment and fisheries management: An overview. Fish. Res 43, 1–8 (1999).Article 

    Google Scholar 
    Saxton, B. Historical demography and genetic population structure of theBlackfin tuna (Thunnus atlanticus) from the Northwest Atlantic Ocean and the Gulf of Mexico. Texas A&M University (2009).Antoni, L., Luque, P. L., Naghshpour, K., Reynal, L. & Saillant, E. A. Development and characterization of microsatellite markers for blackfin tuna (Thunnus atlanticus) with the use of Illumina paired-end sequencing. Fish. Bull. 112, 322–325 (2014).Article 

    Google Scholar 
    Weir, B. S. & Cockerham, C. C. Estimating F-statistics for the analysis of population structure. Evolution 38, 1358 (1984).CAS 
    PubMed 

    Google Scholar 
    Goudet, J. FSTAT (Version 1.2): A computer program to calculate F-statistics. J. Hered 86, 485–486 (1995).Article 

    Google Scholar 
    Rousset, F. Genepop’007: A complete re-implementation of the genepop software for Windows and Linux. Mol. Ecol. Resour. 8, 103–106 (2008).PubMed 
    Article 

    Google Scholar 
    Guo, S. W. & Thompson, E. A. Performing the exact test of Hardy-Weinberg proportion for multiple alleles. Biometrics 48, 361–372 (1992).CAS 
    PubMed 
    MATH 
    Article 

    Google Scholar 
    Van Oosterhout, C., Huthinson, W. F., Wills, D. P. M. & Shipley, P. Micro-checker: Software for identifying and correcting genotyping errors in microsatellite data. Mol. Ecol. Notes 4, 535–538 (2004).Article 
    CAS 

    Google Scholar 
    Excoffier, L., Smouse, P. E. & Quattro, J. M. Analysis of molecular variance inferred from metric distances among DNA haplotypes: Application to human mitochondrial DNA restriction data. Genetics 131, 479–491 (1992).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Excoffier, L. & Lischer, H. E. L. Arlequin suite ver 3.5: A new series of programs to perform population genetics analyses under Linux and Windows. Mol. Ecol. Resour. 10, 564–567 (2010).PubMed 
    Article 

    Google Scholar 
    Benjamini, Y. & Hochberg, Y. Controlling the false discovery rate: A practical and powerful approach to multiple testing. J. R. Stat. Soc. Ser. B (Methodol.) 57, 289–300 (1995).MathSciNet 
    MATH 

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

    Google Scholar 
    Falush, D., Stephens, M. & Pritchard, J. K. Inference of population structure using multilocus genotype data: Linked loci and correlated allele frequencies. Genetics 164, 1567–1587 (2003).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Hubisz, M. J., Falush, D., Stephens, M. & Pritchard, J. K. Inferring weak population structure with the assistance of sample group information. Mol. Ecol. Resour. 9, 1322–1332 (2009).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Jombart, T., Devillard, S. & Balloux, F. Discriminant analysis of principal components: A new method for the analysis of genetically structured populations. BMC Genet. 11, 94 (2010).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Dupanloup, I., Schneider, S. & Excoffier, L. A simulated annealing approach to define the genetic structure of populations. Mol. Ecol. 11, 2571–2581 (2002).CAS 
    PubMed 
    Article 

    Google Scholar 
    Peakall, R. & Smouse, P. E. GenAlEx 6.5: Genetic analysis in Excel. Population genetic software for teaching and research-an update. Bioinformatics 28, 2537–2539 (2012).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Smouse, P. E. & Peakall, R. Spatial autocorrelation analysis of individual multiallele and multilocus genetic structure. Heredity 82(Pt 5), 561–573 (1999).PubMed 
    Article 

    Google Scholar 
    Rousset, F. Genetic differentiation and estimation of gene flow from F-statistics under isolation by distance. Genetics 145, 1219–1228 (1997).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Bezerra, N. P. A. et al. Reproduction of Blackfin tuna Thunnus atlanticus (Perciformes: Scombridae) in Saint Peter and Saint Paul Archipelago, Equatorial Atlantic, Brazil. Rev. Biol. Trop. 61, 1327–1339 (2013).PubMed 
    Article 

    Google Scholar 
    Fitzpatrick, B. M. Power and sample size for nested analysis of molecular variance. Mol. Ecol. 18, 3961–3966 (2009).PubMed 
    Article 

    Google Scholar 
    Ely, B. et al. Consequences of the historical demography on the global population structure of two highly migratory cosmopolitan marine fishes: The yellowfin tuna (Thunnus albacares) and the skipjack tuna (Katsuwonus pelamis). BMC Evol. Biol. 5, 19 (2005).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Alvarado Bremer, J. R., Viñas, J., Mejuto, J., Ely, B. & Pla, C. Comparative phylogeography of Atlantic bluefin tuna and swordfish: The combined effects of vicariance, secondary contact, introgression, and population expansion on the regional phylogenies of two highly migratory pelagic fishes. Mol. Phylogenet. Evol. 36, 169–187 (2005).CAS 
    PubMed 
    Article 

    Google Scholar 
    Hedgecock, D., Barber, P. & Edmands, S. Genetic approaches to measuring connectivity. Oceanography 20, 70–79 (2007).Article 

    Google Scholar 
    Pruett, C. L., Saillant, E. & Gold, J. R. Historical population demography of red snapper (Lutjanus campechanus) from the northern Gulf of Mexico based on analysis of sequences of mitochondrial DNA. Mar. Biol. 147, 593–602 (2005).CAS 
    Article 

    Google Scholar 
    Saillant, E., Bradfield, S. C. & Gold, J. R. Genetic variation and spatial autocorrelation among young-of-the-year red snapper (Lutjanus campechanus) in the northern Gulf of Mexico. ICES J. Mar. Sci 67, 1240–1250 (2010).Article 

    Google Scholar 
    Robledo-Arnuncio, J. J. & Rousset, F. Isolation by distance in a continuous population under stochastic demographic fluctuations. J. Evol. Biol. 23, 53–71 (2010).CAS 
    PubMed 
    Article 

    Google Scholar 
    Rocha, L. A., Craig, M. T. & Bowen, B. W. Phylogeography and the conservation of coral reef fishes. Coral Reefs 26, 501–512 (2007).ADS 
    Article 

    Google Scholar 
    Vasconcellos, A. V., Vianna, P., Paiva, P. C., Schama, R. & Solé-Cava, A. Genetic and morphometric differences between yellowtail snapper (Ocyurus chrysurus, Lutjanidae) populations of the tropical West Atlantic. Genet. Mol. Biol. 31, 308–316 (2008).CAS 
    Article 

    Google Scholar 
    Vieira, K. R., Oliveira, J. E. L. & Barbalho, M. C. Reproductive characteristics of blackfin tuna Thunnus atlanticus (Lesson, 1831), in northeast Brazil. Collect. Vol. Sci. Pap ICCAT 58, 1629–1634 (2005).
    Google Scholar 
    Nielsen, E. E. et al. Genomic signatures of local directional selection in a high gene flow marine organism; the Atlantic cod (Gadus morhua). BMC Evol. Biol. 9, 276 (2009).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Lamichhaney, S. et al. Population-scale sequencing reveals genetic differentiation due to local adaptation in Atlantic herring. Proc. Natl. Acad. Sci. USA 109, 19345–19350 (2012).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Latch, E. K., Dharmarajan, G., Glaubitz, J. C. & Rhodes, O. E. Relative performance of Bayesian clustering software for inferring population substructure and individual assignment at low levels of population differentiation. Conserv. Genet. 7, 295–302 (2006).Article 

    Google Scholar 
    Brophy, D., Rodríguez-Ezpeleta, N., Fraile, I. & Arrizabalaga, H. Combining genetic markers with stable isotopes in otoliths reveals complexity in the stock structure of Atlantic bluefin tuna (Thunnus thynnus). Sci. Rep. 10, 14675 (2020).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar  More

  • in

    Sixth sense in the deep-sea: the electrosensory system in ghost shark Chimaera monstrosa

    Danovaro, et al. Ecological variables for developing a global deep-ocean monitoring and conservation strategy. Nat. Ecol. Evol. 4(2), 181–192. https://doi.org/10.1038/s41559-019-1091-z (2020).Danovaro, R., Snelgrove, P. V. R. & Tyler, P. Challenging the paradigms of deep-sea ecology. Trends Ecol. Evol. 29(8), 465–475. https://doi.org/10.1016/j.tree.2014.06.002 (2014).Article 
    PubMed 

    Google Scholar 
    Collin, S. P. The neuroecology of cartilaginous fishes: sensory strategies for survival. Brain Behav. Evol. 80(2), 80–96. https://doi.org/10.1159/000339870 (2012).Article 
    PubMed 

    Google Scholar 
    Carrier, J. C., Musick, J. A., & Heithaus, M. R. (Eds.). Biology of sharks and their relatives. CRC (2012).Musick, J. A. & Cotton, C. F. Bathymetric limits of chondrichthyans in the deep sea: a re-evaluation. Deep Sea Res. Part II 115, 73–80. https://doi.org/10.1016/j.dsr2.2014.10.010 (2015).Article 

    Google Scholar 
    Treberg, J. R. & Speers-Roesch, B. Does the physiology of chondrichthyan fishes constrain their distribution in the deep sea?. J. Exp. Biol. 219(5), 615–625. https://doi.org/10.1242/jeb.128108 (2016).Article 
    PubMed 

    Google Scholar 
    Didier, D. A., Kemper, J. M. & Ebert, D. A. Phylogeny, biology and classification of extant holocephalans. In Biology of Sharks and Their Relatives, 2nd edn (Carrier, J. C., Musick, J. A. & Heithaus, M. R., eds), pp. 97–124. New York, NY: CRC Pres. (2012).Weigmann, S. Annotated checklist of the living sharks, batoids and chimaeras (Chondrichthyes) of the world, with a focus on biogeographical diversity. J. Fish Biol. 88(3), 837–1037. https://doi.org/10.1111/jfb.12874 (2016).CAS 
    Article 
    PubMed 

    Google Scholar 
    Coates, M. I., Gess, R. W., Finarelli, J. A., Criswell, K. E. & Tietjen, K. A symmoriiform chondrichthyan braincase and the origin of chimaeroid fishes. Nature 541(7636), 208–211. https://doi.org/10.1038/nature20806 (2017).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    Lisney, T. J. A review of the sensory biology of chimaeroid fishes (Chondrichthyes; Holocephali). Rev. Fish Biol. Fisheries 20(4), 571–590. https://doi.org/10.1007/s11160-010-9162-x (2010).Article 

    Google Scholar 
    Finucci, B. et al. Ghosts of the deep–biodiversity, fisheries, and extinction risk of ghost sharks. Fish Fish. 22(2), 391–412. https://doi.org/10.1111/faf.12526 (2021).Article 

    Google Scholar 
    Newton, K. C., Gill, A. B. & Kajiura, S. M. Electroreception in marine fishes: chondrichthyans. J. Fish Biol. 95(1), 135–154. https://doi.org/10.1111/jfb.14068 (2019).Article 
    PubMed 

    Google Scholar 
    Crampton, W. G. Electroreception, electrogenesis and electric signal evolution. J. Fish Biol. 95(1), 92–134. https://doi.org/10.1111/jfb.13922 (2019).Article 
    PubMed 

    Google Scholar 
    Whitehead, D. L. Ampullary organs and electroreception in freshwater Carcharhinus leucas. J. Physiol.-Paris 96(5–6), 391–395. https://doi.org/10.1016/S0928-4257(03)00017-2 (2002).Article 
    PubMed 

    Google Scholar 
    Raschi, W. G., & Gerry, S. Adaptations in the elasmobranch electroreceptive system. Fish Adaptations. Enfield, NH: Scientific Publishers, 233–258 (2003).Atkinson, C. J. L. & Bottaro, M. Ampullary pore distribution of Galeus melastomus and Etmopterus spinax: possible relations with predatory lifestyle and habitat. J. Mar. Biol. Assoc. UK 86(2), 447–448. https://doi.org/10.1017/S0025315406013336 (2006).Article 

    Google Scholar 
    Kempster, R. M. & Collin, S. P. Electrosensory pore distribution and feeding in the basking shark Cetorhinus maximus (Lamniformes: Cetorhinidae). Aquat. Biol. 12(1), 33–36. https://doi.org/10.3354/ab00328 (2011).Article 

    Google Scholar 
    Kempster, R. M., McCarthy, I. D. & Collin, S. P. Phylogenetic and ecological factors influencing the number and distribution of electroreceptors in elasmobranchs. J. Fish Biol. 80(5), 2055–2088. https://doi.org/10.1111/j.1095-8649.2011.03214.x (2012).CAS 
    Article 
    PubMed 

    Google Scholar 
    Whitehead, D. L., Gauthier, A. R., Mu, E. W., Bennett, M. B. & Tibbetts, I. R. Morphology of the Ampullae of Lorenzini in juvenile freshwater Carcharhinus leucas. J. Morphol. 276(5), 481–493. https://doi.org/10.1002/jmor.20355 (2015).Article 
    PubMed 

    Google Scholar 
    Gauthier, A. R. G., Whitehead, D. L., Tibbetts, I. R., Cribb, B. W. & Bennett, M. B. Morphological comparison of the Ampullae of Lorenzini of three sympatric benthic rays. J. Fish Biol. 92(2), 504–514. https://doi.org/10.1111/jfb.13531 (2018).CAS 
    Article 
    PubMed 

    Google Scholar 
    Fields, R. D., Bullock, T. H. & Lange, G. D. Ampullary sense organs, peripheral, central and behavioral electroreception in Chimeras (Hydrolagus, Holocephali, Chondrichthyes). Brain Behav. Evol. 41(6), 269–289. https://doi.org/10.1159/000113849 (1993).CAS 
    Article 
    PubMed 

    Google Scholar 
    Didier, D.A. Phylogenetic systematics of extant chimaeroid fishes (Holocephali, Chimaeroidei). American Museum Novitates; n. 3119 (1995).Serena, F. Field identification guide to the sharks and rays of the Mediterranean and Black Sea (Food and Agriculture Organization, 2005).
    Google Scholar 
    Holt, R. E., Foggo, A., Neat, F. C. & Howell, K. L. Distribution patterns and sexual segregation in chimaeras: implications for conservation and management. ICES J. Mar. Sci. 70(6), 1198–1205. https://doi.org/10.1093/icesjms/fst058 (2013).Article 

    Google Scholar 
    Ragonese, S., Vitale, S., Dimech, M., & Mazzola, S. Abundances of demersal sharks and chimaera from 1994–2009 scientific surveys in the central Mediterranean Sea. PloS one, 8(9). https://doi.org/10.1371/journal.pone.0074865 (2013).Vacchi, M., & Orsi, L. R. Alimentazione di Chimaera monstrosa L. sui fondi batiali liguri. Atti della Società Toscana di Scienze Naturali, Memorie serie B, 86, 388–391 (1979).Macpherson, E. Food and feeding of Chimaera monstrosa, Linnaeus, 1758, in the western Mediterranean. ICES J. Mar. Sci. 39(1), 26–29. https://doi.org/10.1093/icesjms/39.1.26 (1980).Article 

    Google Scholar 
    Mauchline, J. & Gordon, J. D. M. Diets of the sharks and chimaeroids of the Rockall Trough, northeastern Atlantic Ocean. Mar. Biol. 75(2–3), 269–278. https://doi.org/10.1007/BF00406012 (1983).Article 

    Google Scholar 
    Albo-Puigserver, et al. Feeding ecology and trophic position of three sympatric demersal chondrichthyans in the northwestern Mediterranean. Mar. Ecol. Prog. Ser. 524, 255–268. https://doi.org/10.3354/meps11188( (2015).ADS 
    Article 

    Google Scholar 
    Priede, I. G. Deep-sea fishes: biology, diversity, ecology and fisheries. Cambridge University Press (2017).Ferrando, S. et al. First description of a palatal organ in Chimaera monstrosa (Chondrichthyes, Holocephali). Anat. Rec. 299(1), 118–131. https://doi.org/10.1002/ar.23280 (2016).Article 

    Google Scholar 
    Garza-Gisholt, E., Hart, N. S., & Collin, S. P. Retinal morphology and visual specializations in three species of chimaeras, the deep-sea R. pacifica and C. lignaria, and the Vertical Migrator C. milii (Holocephali). Brain, behavior and evolution, 92(1–2), 47–62. https://doi.org/10.1159/000490655 (2018).Pethybridge, H., Daley, R. K. & Nichols, P. D. Diet of demersal sharks and chimaeras inferred by fatty acid profiles and stomach content analysis. J. Exp. Mar. Biol. Ecol. 409(1–2), 290–299. https://doi.org/10.1016/j.jembe.2011.09.009 (2011).Article 

    Google Scholar 
    Rivera-Vicente, A. C., Sewell, J. & Tricas, T. C. Electrosensitive spatial vectors in elasmobranch fishes: implications for source localization. PLoS ONE 6(1), e16008. https://doi.org/10.1371/journal.pone.0016008 (2011).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Kajiura, S. M., Cornett, A. D. & Yopak, K. E. Sensory adaptations to the environment: electroreceptors as a case study. Biol. Sharks Relatives 2, 393–434 (2010).Article 

    Google Scholar 
    Raschi, W. A morphological analysis of the Ampullae of Lorenzini in selected skates (Pisces, Rajoidei). J. Morphol. 189(3), 225–247. https://doi.org/10.1002/jmor.1051890303 (1986).Article 
    PubMed 

    Google Scholar 
    Jordan, L. K. et al. Linking sensory biology and fisheries bycatch reduction in elasmobranch fishes: a review with new directions for research. Conserv. Physiol. 1(1), cot002. https://doi.org/10.1093/conphys/cot002 (2013).Wueringer, B. E., Peverell, S. C., Seymour, J., Squire Jr, L., Kajiura, S. M., & Collin, S. P. Sensory systems in sawfishes. 1. The ampullae of Lorenzini. Brain, behavior and evolution, 78(2), 139–149. https://doi.org/10.1159/000329515 (2011).Bird C.S. The tropho-spatial ecology of deep-sea sharks and chimaeras from a stable isotope perspective. PhD thesis – University of Southampton, UK (2017).Andres, K. H. & Von Düring, M. Comparative anatomy of vertebrate electroreceptors. Prog Brain Res 74, 113–131. https://doi.org/10.1016/S0079-6123(08)63006-X (1998).Article 

    Google Scholar 
    Crooks, N. & Waring, C. P. A study into the sexual dimorphisms of the Ampullae of Lorenzini in the lesser-spotted catshark, Scyliorhinus canicula (Linnaeus, 1758). Environ. Biol. Fishes 96(5), 585–590. https://doi.org/10.1016/S0079-6123(08)63006-X (2013).Article 

    Google Scholar 
    Didier, D. A. Phylogeny and classification of extant Holocephali. Biol. Sharks Relatives 4, 115–138 (2004).Article 

    Google Scholar 
    Wueringer, B. E. & Tibbetts, I. R. Comparison of the lateral line and ampullary systems of two species of shovelnose ray. Rev. Fish Biol. Fisheries 18(1), 47–64. https://doi.org/10.1007/s11160-007-9063-9 (2008).Article 

    Google Scholar 
    Theiss, S. M., Collin, S. P. & Hart, N. S. Morphology and distribution of the ampullary electroreceptors in wobbegong sharks: implications for feeding behaviour. Mar. Biol. 158(4), 723–735. https://doi.org/10.1007/s00227-010-1595-1 (2011).Article 

    Google Scholar 
    Schäfer, B. T. et al. Morphological observations of Ampullae of lorenzini in Squatina guggenheim and S. occulta (Chondrichthyes, Elasmobranchii, Squatinidae). Microscopy Res Tech. 75(9), 1213–1217. https://doi.org/10.1002/jemt.22051 (2012).Brown, B. R. Sensing temperature without ion channels. Nature 421(6922), 495–495. https://doi.org/10.1038/421495a (2003).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    Fields, R. D., Fields, K. D. & Fields, M. C. Semiconductor gel in shark sense organs?. Neurosci. Lett. 426(3), 166–170. https://doi.org/10.1016/j.neulet.2007.08.064 (2007).CAS 
    Article 
    PubMed 
    PubMed Central 
    MATH 

    Google Scholar 
    Brown, B. R. Temperature response in electrosensors and thermal voltages in electrolytes. J. Biol. Phys. 36(2), 121–134. https://doi.org/10.1007/s10867-009-9174-8 (2010).Article 
    PubMed 

    Google Scholar 
    Josberger, E. E. et al. Proton conductivity in Ampullae of Lorenzini jelly. Sci. Adv. 2(5), e1600112. https://doi.org/10.1126/sciadv.1600112 (2016).ADS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Froese, R. and Pauly D. https://www.fishbase.de/ (2021).Sims, D. W. The biology, ecology and conservation of elasmobranchs: recent advances and new frontiers. J. Fish Biol. 87(6), 1265–1270. https://doi.org/10.1111/jfb.12861 (2015).CAS 
    Article 
    PubMed 

    Google Scholar 
    Heithaus, M. R., Frid, A., Wirsing, A. & Worm, B. Predicting ecological consequences of marine top predator declines. Trends Ecol. Evol. 23, 202–210. https://doi.org/10.1016/j.tree.2008.01.003 (2008).Article 
    PubMed 

    Google Scholar 
    Dymek, J., Muñoz, P., Mayo-Hernández, E., Kuciel, M. & Żuwała, K. Comparative analysis of the olfactory organs in selected species of marine sharks and freshwater batoids. Zool. Anz. 294, 50–61. https://doi.org/10.1016/j.jcz.2021.07.013 (2021).Article 

    Google Scholar 
    Bellono, N. W., Leitch, D. B. & Julius, D. Molecular tuning of electroreception in sharks and skates. Nature 558(7708), 122. https://doi.org/10.1038/s41586-018-0160-9 (2018).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Luchetti, E. A., Iglésias, S. P., & Sellos, D. Y. Chimaera opalescens n. sp., a new chimaeroid (Chondrichthyes: Holocephali) from the north‐eastern Atlantic Ocean. J. Fish Biol., 79(2), 399–417. https://doi.org/10.1111/j.1095-8649.2011.03027.x (2011).Marranzino, A. N. & Webb, J. F. Flow sensing in the deep sea: the lateral line system of stomiiform fishes. Zool. J. Linn. Soc. 183(4), 945–965. https://doi.org/10.1093/zoolinnean/zlx090 (2018).Article 

    Google Scholar 
    Yopak, K. E. & Montgomery, J. C. Brain organization and specialization in deep-sea chondrichthyans. Brain Behav. Evol. 71(4), 287–304. https://doi.org/10.1159/000127048 (2008).Article 
    PubMed 

    Google Scholar 
    Schneider, C. A., Rasband, W. S. & Eliceiri, K. W. NIH Image to ImageJ: 25 years of image analysis. Nat. Methods 9(7), 671–675. https://doi.org/10.1038/nmeth.2089 (2012).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    R Core Team, R. A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. https://www.R-project.org/ (2021).Wickham, H. ggplot2: Elegant Graphics for Data Analysis. Springer, New York (2016). More

  • in

    Climatic and tectonic drivers shaped the tropical distribution of coral reefs

    Spalding, M. D. & Grenfell, A. M. New estimates of global and regional coral reef areas. Coral Reefs 16, 225–230 (1997).Article 

    Google Scholar 
    Moberg, F. & Folke, C. Ecological goods and services of coral reef ecosystems. Ecol. Econ. 29, 215–233 (1999).Article 

    Google Scholar 
    Roberts, C. M. et al. Marine Biodiversity Hotspots and Conservation Priorities for Tropical Reefs. Science 295, 1280–1284 (2002).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Johannes, R., Wiebe, W., Crossland, C., Rimmer, D. & Smith, S. Latitudinal limits of coral reef growth. Mar. Ecol. Prog. Ser. 11, 105–111 (1983).ADS 
    Article 

    Google Scholar 
    Kleypas, J. A., Mcmanus, J. W. & Meñez, L. A. B. Environmental Limits to Coral Reef Development: Where Do We Draw the Line? Am. Zool. 39, 146–159 (1999).Article 

    Google Scholar 
    Yamano, H., Hori, K., Yamauchi, M., Yamagawa, O. & Ohmura, A. Highest-latitude coral reef at Iki Island, Japan. Coral Reefs 20, 9–12 (2001).Article 

    Google Scholar 
    Guan, Y., Hohn, S. & Merico, A. Suitable Environmental Ranges for Potential Coral Reef Habitats in the Tropical Ocean. PLOS ONE 10, e0128831 (2015).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Bellwood, D. R. & Hughes, T. P. Regional-Scale Assembly Rules and Biodiversity of Coral Reefs. Science 292, 1532–1535 (2001).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Connolly, S. R., Bellwood, D. R. & Hughes, T. P. Indo-Pacific Biodiversity of Coral Reefs: Deviations from a Mid-Domain Model. Ecology 84, 2178–2190 (2003).Article 

    Google Scholar 
    Bellwood, D. R., Hughes, T. P., Connolly, S. R. & Tanner, J. Environmental and geometric constraints on Indo‐Pacific coral reef biodiversity. Ecol. Lett. 8, 643–651 (2005).Article 

    Google Scholar 
    Kiessling, W., Simpson, C., Beck, B., Mewis, H. & Pandolfi, J. M. Equatorial decline of reef corals during the last Pleistocene interglacial. Proc. Natl Acad. Sci. 109, 21378–21383 (2012).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Veron, J. E. N. et al. Delineating the Coral Triangle. Galaxea. J. Coral Reef. Stud. 11, 91–100 (2009).Article 

    Google Scholar 
    Briggs, J. C. Marine Longitudinal Biodiversity: Causes and Conservation. Divers. Distrib. 13, 544–555 (2007).Article 

    Google Scholar 
    Renema, W. et al. Hopping Hotspots: Global Shifts in Marine Biodiversity. Science 321, 654–657 (2008).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Kiessling, W. Paleoclimatic significance of Phanerozoic reefs. Geology 29, 751–754 (2001).ADS 
    Article 

    Google Scholar 
    Wallace, C. & Rosen, B. Diverse staghorn corals (Acropora) in high-latitude Eocene assemblages: Implications for the evolution of modern diversity patterns of reef corals. Proc. Biol. Sci. 273, 975–982 (2006).PubMed 
    PubMed Central 

    Google Scholar 
    Perrin, C. & Kiessling, W. Latitudinal trends in Cenozoic reef patterns and their relationship to climate. Carbonate Syst. Oligocene–Miocene Clim. Transit. 17–33 (Wiley-Blackwell, 2010).Kiessling, W. Habitat effects and sampling bias on Phanerozoic reef distribution. Facies 51, 24–32 (2005).Article 

    Google Scholar 
    Kiessling, W. Reef expansion during the Triassic: Spread of photosymbiosis balancing climatic cooling. Palaeogeogr. Palaeoclimatol. Palaeoecol. 290, 11–19 (2010).Article 

    Google Scholar 
    Ziegler, A. M., Hulver, M. L., Lotts, A. L. & Schmachtenberg, W. F. Uniformitarianism and palaeoclimates: inferences from the distribution of carbonate rocks. In: Fossils and Climate (ed. Brenchley, P. J.), 3–25 (Wiley, Chichester, 1984).Crame, J. A. & Rosen, B. R. Cenozoic palaeogeography and the rise of modern biodiversity patterns. Geol. Soc. Lond. Spec. Publ. 194, 153–168 (2002).ADS 
    Article 

    Google Scholar 
    Leprieur, F. et al. Plate tectonics drive tropical reef biodiversity dynamics. Nat. Commun. 7, 1–8 (2016).Article 
    CAS 

    Google Scholar 
    Zaffos, A., Finnegan, S. & Peters, S. E. Plate tectonic regulation of global marine animal diversity. Proc. Natl Acad. Sci. U. S. A. 114, 5653–5658 (2017).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Roberts, G. G. & Mannion, P. D. Timing and periodicity of Phanerozoic marine biodiversity and environmental change. Sci. Rep. 9, 6116 (2019).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Valentine, J. W. & Moores, E. M. Global Tectonics and the Fossil Record. J. Geol. 80, 167–184 (1972).ADS 
    Article 

    Google Scholar 
    Pellissier, L., Heine, C., Rosauer, D. F. & Albouy, C. Are global hotspots of endemic richness shaped by plate tectonics? Biol. J. Linn. Soc. 123, 247–261 (2017).Article 

    Google Scholar 
    Chittaro, P. M. Species-area relationships for coral reef fish assemblages of St. Croix, US Virgin Islands. Mar. Ecol. Prog. Ser. 233, 253–261 (2002).ADS 
    Article 

    Google Scholar 
    Tittensor, D. P., Micheli, F., Nyström, M. & Worm, B. Human impacts on the species–area relationship in reef fish assemblages. Ecol. Lett. 10, 760–772 (2007).PubMed 
    Article 

    Google Scholar 
    Tittensor, D. P. et al. Global patterns and predictors of marine biodiversity across taxa. Nature 466, 1098–1101 (2010).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Huntington, B. E. & Lirman, D. Species-area relationships in coral communities: evaluating mechanisms for a commonly observed pattern. Coral Reefs 31, 929–938 (2012).ADS 
    Article 

    Google Scholar 
    Kiessling, W., Simpson, C. & Foote, M. Reefs as cradles of evolution and sources of biodiversity in the Phanerozoic. Science 327, 196–198 (2010).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Pandolfi, J. M. et al. Global Trajectories of the Long-Term Decline of Coral Reef Ecosystems. Science 301, 955–958 (2003).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Hoegh-Guldberg, O. Coral reef ecosystems and anthropogenic climate change. Reg. Environ. Change 11, 215–227 (2011).Article 

    Google Scholar 
    Hughes, T. P. et al. Global warming and recurrent mass bleaching of corals. Nature 543, 373–377 (2017).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Kim, S. W. et al. Refugia under threat: Mass bleaching of coral assemblages in high-latitude eastern Australia. Glob. Change Biol. 25, 3918–3931 (2019).ADS 
    Article 

    Google Scholar 
    Pörtner, H.-O. et al. IPCC special report on the ocean and cryosphere in a changing climate. IPCC Intergov. Panel Clim. Change Geneva Switz. 1, 1–755 (2019).Sully, S., Burkepile, D. E., Donovan, M. K., Hodgson, G. & van Woesik, R. A global analysis of coral bleaching over the past two decades. Nat. Commun. 10, 1264 (2019).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Couce, E., Ridgwell, A. & Hendy, E. J. Future habitat suitability for coral reef ecosystems under global warming and ocean acidification. Glob. Change Biol. 19, 3592–3606 (2013).ADS 
    Article 

    Google Scholar 
    Hoegh-Guldberg, O., Poloczanska, E. S., Skirving, W. & Dove, S. Coral Reef Ecosystems under Climate Change and Ocean Acidification. Front. Mar. Sci. 4, 1–20 (2017).O’Neill, B. C. et al. The Scenario Model Intercomparison Project (ScenarioMIP) for CMIP6. Geosci. Model Dev. 9, 3461–3482 (2016).ADS 
    Article 

    Google Scholar 
    Precht, W. F. & Aronson, R. B. Climate flickers and range shifts of reef corals. Front. Ecol. Environ. 2, 307–314 (2004).Article 

    Google Scholar 
    Greenstein, B. J. & Pandolfi, J. M. Escaping the heat: range shifts of reef coral taxa in coastal Western Australia. Glob. Change Biol. 14, 513–528 (2008).ADS 
    Article 

    Google Scholar 
    Pellissier, L. et al. Quaternary coral reef refugia preserved fish diversity. Science 344, 1016–1019 (2014).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Vilhena, D. A. & Smith, A. B. Spatial Bias in the Marine Fossil Record. PLoS ONE 8, 1–7 (2013).Article 
    CAS 

    Google Scholar 
    Close, R. A., Benson, R. B. J., Saupe, E. E., Clapham, M. E. & Butler, R. J. The spatial structure of Phanerozoic marine animal diversity. Science 368, 420–424 (2020).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Jones, L. A., Dean, C. D., Mannion, P. D., Farnsworth, A. & Allison, P. A. Spatial sampling heterogeneity limits the detectability of deep time latitudinal biodiversity gradients. Proc. R. Soc. B Biol. Sci. 288, 20202762 (2021).Article 

    Google Scholar 
    Jones, L. A. & Eichenseer, K. Uneven spatial sampling distorts reconstructions of Phanerozoic seawater temperature. Geology (2021) https://doi.org/10.1130/G49132.1.Stolarski, J. et al. The ancient evolutionary origins of Scleractinia revealed by azooxanthellate corals. BMC Evol. Biol. 11, 1–11 (2011).Article 

    Google Scholar 
    Frankowiak, K. et al. Photosymbiosis and the expansion of shallow-water corals. Sci. Adv. 2, e1601122 (2016).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

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

    Google Scholar 
    Phillips, S. J., Anderson, R. P. & Schapire, R. E. Maximum entropy modeling of species geographic distributions. Ecol. Model. 190, 231–259 (2006).Article 

    Google Scholar 
    Swets, J. A. Measuring the accuracy of diagnostic systems. Science 240, 1285–1293 (1988).ADS 
    MathSciNet 
    CAS 
    PubMed 
    MATH 
    Article 

    Google Scholar 
    Boyce, M. S., Vernier, P. R., Nielsen, S. E. & Schmiegelow, F. K. A. Evaluating resource selection functions. Ecol. Model. 157, 281–300 (2002).Article 

    Google Scholar 
    Hirzel, A. H., LeLay, 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 
    Elith, J., Kearney, M. & Phillips, S. The art of modelling range-shifting species. Methods Ecol. Evol. 1, 330–342 (2010).Article 

    Google Scholar 
    Miller, K. G. et al. The Phanerozoic Record of Global Sea-Level Change. Science 310, 1293–1298 (2005).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Hallam, A., Grose, J. A. & Ruffell, A. H. Palaeoclimatic significance of changes in clay mineralogy across the Jurassic-Cretaceous boundary in England and France. Palaeogeogr. Palaeoclimatol. Palaeoecol. 81, 173–187 (1991).Article 

    Google Scholar 
    Gröcke, D. R., Price, G. D., Ruffell, A. H., Mutterlose, J. & Baraboshkin, E. Isotopic evidence for Late Jurassic–Early Cretaceous climate change. Palaeogeogr. Palaeoclimatol. Palaeoecol. 202, 97–118 (2003).Article 

    Google Scholar 
    Royer, D. L., Berner, R. A., Montañez, I. P., Tabor, N. J. & Beerling, D. J. CO2 as a primary driver of Phanerozoic climate. GSA Today 14, 1–10 (2004).
    Google Scholar 
    Grabowski, J. et al. Magnetic susceptibility and spectral gamma logs in the Tithonian–Berriasian pelagic carbonates in the Tatra Mts (Western Carpathians, Poland): Palaeoenvironmental changes at the Jurassic/Cretaceous boundary. Cretac. Res. 43, 1–17 (2013).Article 

    Google Scholar 
    Vickers, M. L. et al. The duration and magnitude of Cretaceous cool events: Evidence from the northern high latitudes. GSA Bull. 131, 1979–1994 (2019).CAS 
    Article 

    Google Scholar 
    Hay, W. W. & Floegel, S. New thoughts about the Cretaceous climate and oceans. Earth-Sci. Rev. 115, 262–272 (2012).ADS 
    CAS 
    Article 

    Google Scholar 
    Tennant, J. P., Mannion, P. D. & Upchurch, P. Sea level regulated tetrapod diversity dynamics through the Jurassic/Cretaceous interval. Nat. Commun. 7, 12737 (2016).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Schouten, S. et al. Onset of long-term cooling of Greenland near the Eocene-Oligocene boundary as revealed by branched tetraether lipids. Geology 36, 147 (2008).ADS 
    Article 

    Google Scholar 
    Zachos, J. C., Dickens, G. R. & Zeebe, R. E. An early Cenozoic perspective on greenhouse warming and carbon-cycle dynamics. Nature 451, 279–283 (2008).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Crame, J. A. Taxonomic diversity gradients through geological time. Divers. Distrib. 7, 175–189 (2001).
    Google Scholar 
    Mannion, P. D., Upchurch, P., Benson, R. B. J. & Goswami, A. The latitudinal biodiversity gradient through deep time. Trends Ecol. Evol. 29, 42–50 (2014).PubMed 
    Article 

    Google Scholar 
    Fenton, I. S. et al. The impact of Cenozoic cooling on assemblage diversity in planktonic foraminifera. Philos. Trans. R. Soc. B Biol. Sci. 371, 1–12 (2016).Article 
    CAS 

    Google Scholar 
    Saupe, E. E. et al. Climatic shifts drove major contractions in avian latitudinal distributions throughout the Cenozoic. Proc. Natl Acad. Sci. 116, 12895–12900 (2019).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Hall, R. Cenozoic geological and plate tectonic evolution of SE Asia and the SW Pacific: computer-based reconstructions, model and animations. J. Asian Earth Sci. 20, 353–431 (2002).ADS 
    Article 

    Google Scholar 
    Hall, R. Southeast Asia’s changing palaeogeography. Blumea 54, 148–161 (2009).Article 

    Google Scholar 
    Gaboriau, T. et al. Ecological constraints coupled with deep-time habitat dynamics predict the latitudinal diversity gradient in reef fishes. Proc. R. Soc. B Biol. Sci. 286, 20191506 (2019).Article 

    Google Scholar 
    Saupe, E. E. et al. Extinction intensity during Ordovician and Cenozoic glaciations explained by cooling and palaeogeography. Nat. Geosci. 13, 65–70 (2020).ADS 
    Article 
    CAS 

    Google Scholar 
    Lunt, D. J. et al. DeepMIP: model intercomparison of early Eocene climatic optimum (EECO) large-scale climate features and comparison with proxy data. Clim 17, 203–227 (2021).ADS 

    Google Scholar 
    Freeman, L. A., Kleypas, J. A. & Miller, A. J. Coral Reef Habitat Response to Climate Change Scenarios. PLoS ONE 8, 1–14 (2013).
    Google Scholar 
    Foster, G. L., Royer, D. L. & Lunt, D. J. Future climate forcing potentially without precedent in the last 420 million years. Nat. Commun. 8, 1–8 (2017).ADS 
    Article 
    CAS 

    Google Scholar 
    Farnsworth, A. et al. Past East Asian monsoon evolution controlled by paleogeography, not CO2. Sci. Adv. 5, 1–13 (2019).Article 
    CAS 

    Google Scholar 
    Zhang, L. et al. Consensus Forecasting of Species Distributions: The Effects of Niche Model Performance and Niche Properties. PLoS ONE 10, 1–18 (2015).
    Google Scholar 
    Harrison, S. P. et al. Evaluation of CMIP5 palaeo-simulations to improve climate projections. Nat. Clim. Change 5, 735–743 (2015).ADS 
    Article 

    Google Scholar 
    Seo, C., Thorne, J. H., Hannah, L. & Thuiller, W. Scale effects in species distribution models: implications for conservation planning under climate change. Biol. Lett. 5, 39–43 (2009).PubMed 
    Article 

    Google Scholar 
    Couce, E., Ridgwell, A. & Hendy, E. J. Environmental controls on the global distribution of shallow-water coral reefs. J. Biogeogr. 39, 1508–1523 (2012).Article 

    Google Scholar 
    Laborel, J. West African reef corals: an hypothesis on their origin. in Proceedings of the Second International Coral Reef Symposium vol. 1 425–443 (Great Barrier Reef Committee Brisbane, 1974).Spalding, M., Spalding, M. D., Ravilious, C. & Green, E. P. World Atlas of Coral Reefs. (University of California Press, 2001).Block, S. et al. Where to Dig for Fossils: Combining Climate-Envelope, Taphonomy and Discovery Models. PLoS ONE 11, 1–16 (2016).Jones, L. A. et al. Coupling of palaeontological and neontological reef coral data improves forecasts of biodiversity responses under global climatic change. R. Soc. Open Sci. 6, 182111 (2019).ADS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Kusumoto, B. et al. Global distribution of coral diversity: Biodiversity knowledge gradients related to spatial resolution. Ecol. Res. 35, 315–326 (2020).Article 

    Google Scholar 
    Muir, P. R., Wallace, C. C., Done, T. & Aguirre, J. D. Limited scope for latitudinal extension of reef corals. Science 348, 1135–1138 (2015).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Sillero, N. & Barbosa, A. M. Common mistakes in ecological niche models. Int. J. Geogr. Inf. Sci. 35, 213–226 (2021).Article 

    Google Scholar 
    Valdes, P. J. et al. The BRIDGE HadCM3 family of climate models:HadCM3@Bristol v1.0. Geosci. Model Dev. 10, 3715–3743 (2017).ADS 
    CAS 
    Article 

    Google Scholar 
    Sheppard, C. R. C. Predicted recurrences of mass coral mortality in the Indian Ocean. Nature 425, 294–297 (2003).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Saupe, E. E. et al. Macroevolutionary consequences of profound climate change on niche evolution in marine molluscs over the past three million years. Proc. R. Soc. B Biol. Sci. 281, 1–9 (2014).
    Google Scholar 
    Haywood, A. M. et al. What can Palaeoclimate Modelling do for you? Earth Syst. Environ. 3, 1–18 (2019).Article 

    Google Scholar 
    Sellwood, B. W. & Valdes, P. J. Mesozoic climates: General circulation models and the rock record. Sediment. Geol. 190, 269–287 (2006).ADS 
    Article 

    Google Scholar 
    Waterson, A. M. et al. Modelling the climatic niche of turtles: a deep-time perspective. Proc. R. Soc. B Biol. Sci. 283, 1–9 (2016).
    Google Scholar 
    Chiarenza, A. A. et al. Ecological niche modelling does not support climatically-driven dinosaur diversity decline before the Cretaceous/Paleogene mass extinction. Nat. Commun. 10, 1–14 (2019).CAS 
    Article 

    Google Scholar 
    Dunne, E. M., Farnsworth, A., Greene, S. E., Lunt, D. J. & Butler, R. J. Climatic drivers of latitudinal variation in Late Triassic tetrapod diversity. Palaeontology 64, 101–117 (2020).Article 

    Google Scholar 
    Lyster, S. J., Whittaker, A. C., Allison, P. A., Lunt, D. J. & Farnsworth, A. Predicting sediment discharges and erosion rates in deep time—examples from the late Cretaceous North American continent. Basin Res. 1–27 (2020) https://doi.org/10.1111/bre.12442.Lunt, D. J. et al. Palaeogeographic controls on climate and proxy interpretation. Clim 12, 1181–1198 (2016).ADS 

    Google Scholar 
    Vasquez, V. L., de Lima, A. A., dos Santos, A. P. & Pinto, M. P. Influence of spatial extent on habitat suitability models for primate species of Atlantic Forest. Ecol. Inform. 61, 101179 (2021).Article 

    Google Scholar 
    Collins, D. S. et al. Controls on tidal sedimentation and preservation: Insights from numerical tidal modelling in the Late Oligocene–Miocene South China Sea, Southeast Asia. Sedimentology 65, 2468–2505 (2018).Article 

    Google Scholar 
    Dean, C. D., Collins, D. S., van Cappelle, M., Avdis, A. & Hampson, G. J. Regional-scale paleobathymetry controlled location, but not magnitude, of tidal dynamics in the Late Cretaceous Western Interior Seaway, USA. Geology 47, 1083–1087 (2019).ADS 
    CAS 
    Article 

    Google Scholar 
    Markwick, P. J. & Valdes, P. J. Palaeo-digital elevation models for use as boundary conditions in coupled ocean–atmosphere GCM experiments: a Maastrichtian (late Cretaceous) example. Palaeogeogr. Palaeoclimatol. Palaeoecol. 213, 37–63 (2004).Article 

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

    Google Scholar 
    Sillero, N. What does ecological modelling model? A proposed classification of ecological niche models based on their underlying methods. Ecol. Model. 222, 1343–1346 (2011).Article 

    Google Scholar 
    Guisan, A., Thuiller, W. & Zimmermann, N. E. Habitat suitability and distribution models: with applications in R. (Cambridge University Press, 2017).Kearney, M. R., Wintle, B. A. & Porter, W. P. Correlative and mechanistic models of species distribution provide congruent forecasts under climate change. Conserv. Lett. 3, 203–213 (2010).Article 

    Google Scholar 
    Owens, H. L. et al. Constraints on interpretation of ecological niche models by limited environmental ranges on calibration areas. Ecol. Model. 263, 10–18 (2013).Article 

    Google Scholar 
    Franklin, J. Mapping Species Distributions: Spatial Inference and Prediction. (Cambridge University Press, 2010). https://doi.org/10.1017/CBO9780511810602.Liu, C., White, M. & Newell, G. Selecting thresholds for the prediction of species occurrence with presence-only data. J. Biogeogr. 40, 778–789 (2013).Article 

    Google Scholar 
    Liu, C., Newell, G. & White, M. On the selection of thresholds for predicting species occurrence with presence-only data. Ecol. Evol. 6, 337–348 (2016).PubMed 
    Article 

    Google Scholar 
    Kiessling, W. & Krause, M. C. PARED—An online database of Phanerozoic reefs. https://www.paleo-reefs.pal.uni-erlangen.de/ (2021).Jones, L. A., Mannion, P. D., Farnsworth, A., Bragg, F. & Lunt, D. J. Code from ‘Climatic and tectonic drivers shaped the tropical distribution of coral reefs’. Zenodo (2022) https://doi.org/10.5281/zenodo.6458366. More

  • in

    Stocking density mediated stress modulates growth attributes in cage reared Labeo rohita (Hamilton) using multifarious biomarker approach

    Tolussi, C. E., Hilsdorf, A. W. S., Caneppele, D. & Moreira, R. G. The effects of stocking density in physiological parameters and growth of the endangered teleost species piabanha, Brycon insignis (Steindachner, 1877). Aquaculture 310, 221–228 (2010).
    Google Scholar 
    Wang, Y. et al. Effects of stocking density on growth, serum parameters, antioxidant status, liver and intestine histology and gene expression of largemouth bass (Micropterus salmoides) farmed in the in-pond raceway system. Aquac. Res. 51, 5228–5240 (2020).CAS 

    Google Scholar 
    Zahedi, S., Akbarzadeh, A., Mehrzad, J., Noori, A. & Harsij, M. Effect of stocking density on growth performance, plasma biochemistry and muscle gene expression in rainbow trout (Oncorhynchus mykiss). Aquaculture 498, 271–278 (2019).CAS 

    Google Scholar 
    Yousefi, M., Paktinat, M., Mahmoudi, N., Pérez-Jiménez, A. & Hoseini, S. M. Serum biochemical and non-specific immune responses of rainbow trout (Oncorhynchus mykiss) to dietary nucleotide and chronic stress. Fish Physiol. Biochem. 42, 1417–1425 (2016).CAS 
    PubMed 

    Google Scholar 
    Duan, Y., Dong, X., Zhang, X. & Miao, Z. Effects of dissolved oxygen concentration and stocking density on the growth, energy budget and body composition of juvenile Japanese flounder, Paralichthys olivaceus (Temminck et Schlegel). Aquac. Res. 42, 407–416 (2011).CAS 

    Google Scholar 
    Castillo-Vargasmachuca, S. et al. Effect of stocking density on growth performance and yield of subadult pacific red snapper cultured in floating sea cages. N. Am. J. Aquac. 74, 413–418 (2012).
    Google Scholar 
    Upadhyay, A. et al. Stocking density matters in open water cage culture: influence on growth, digestive enzymes, haemato-immuno and stress responses of Puntius sarana (Ham, 1822). Aquaculture 547, 737445 (2021).
    Google Scholar 
    Kumar, V. et al. Assessment of the effect of sub-lethal acute toxicity of Emamectin benzoate in Labeo rohita using multiple biomarker approach. Toxicol. Rep. 9, 102–110 (2022).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Rebl, A. et al. The synergistic interaction of thermal stress coupled with overstocking strongly modulates the transcriptomic activity and immune capacity of rainbow trout (Oncorhynchus mykiss). Sci. Rep. 10, 1–15 (2020).ADS 

    Google Scholar 
    Braun, N., de Lima, R. L., Baldisserotto, B., Dafre, A. L. & de Oliveira Nuñer, A. P. Growth, biochemical and physiological responses of Salminus brasiliensis with different stocking densities and handling. Aquaculture 301, 22–30 (2010).CAS 

    Google Scholar 
    Refaey, M. M., Tian, X., Tang, R. & Li, D. Changes in physiological responses, muscular composition and flesh quality of channel catfish Ictalurus punctatus suffering from transport stress. Aquaculture 478, 9–15 (2017).CAS 

    Google Scholar 
    Liu, G. et al. Influence of stocking density on growth, digestive enzyme activities, immune responses, antioxidant of Oreochromis niloticus fingerlings in biofloc systems. Fish Shellfish Immunol. 81, 416–422 (2018).CAS 
    PubMed 

    Google Scholar 
    Kumar, G. & Engle, C. R. Technological advances that led to growth of shrimp, salmon, and tilapia farming. Rev. Fish. Sci. Aquac. 24, 136–152 (2016).
    Google Scholar 
    Sundin, L. Hypoxia and blood flow control in fish gills. In Biology of tropical fishes (eds Val, A. L. & Almeida-Val, V. M. F.) 353–362 (Manaus INPA, 1999).
    Google Scholar 
    Beveridge, M. C. M. Cage Aquaculture Vol. 5 (John Wiley & Sons, 2008).
    Google Scholar 
    Valenti, W. C., Barros, H. P., Moraes-Valenti, P., Bueno, G. W. & Cavalli, R. O. Aquaculture in Brazil: past, present and future. Aquac. Rep. 19, 100611 (2021).
    Google Scholar 
    Das, A. K., Meena, D. K. & Sharma, A. P. Cage farming in an Indian Reservoir. World Aquac. 45, 56–59 (2014).
    Google Scholar 
    Sarkar, U. K. et al. Status, prospects, threats, and the way forward for sustainable management and enhancement of the tropical Indian reservoir fisheries: an overview. Rev. Fish. Sci. Aquac. 26, 155–175 (2018).
    Google Scholar 
    Singh, A. K. & Lakra, W. S. Culture of Pangasianodon hypophthalmus into India: impacts and present scenario. Pakistan J. Biol. Sci. 15, 19 (2012).CAS 

    Google Scholar 
    Jena, J. et al. Evaluation of growth performance of Labeo fimbriatus (Bloch), Labeo gonius (Hamilton) and Puntius gonionotus (Bleeker) in polyculture with Labeo rohita (Hamilton) during fingerlings rearing at varied densities. Aquaculture 319, 493–496 (2011).
    Google Scholar 
    Liu, B., Jia, R., Han, C., Huang, B. & Lei, J.-L. Effects of stocking density on antioxidant status, metabolism and immune response in juvenile turbot (Scophthalmus maximus). Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 190, 1–8 (2016).CAS 

    Google Scholar 
    Wu, F. et al. Effect of stocking density on growth performance, serum biochemical parameters, and muscle texture properties of genetically improved farm tilapia, Oreochromis niloticus. Aquac. Int. 26, 1247–1259 (2018).CAS 

    Google Scholar 
    Andrade, T. et al. Evaluation of different stocking densities in a Senegalese sole (Solea senegalensis) farm: implications for growth, humoral immune parameters and oxidative status. Aquaculture 438, 6–11 (2015).CAS 

    Google Scholar 
    Qi, C. et al. Effect of stocking density on growth, physiological responses, and body composition of juvenile blunt snout bream, Megalobrama amblycephala. J. World Aquac. Soc. 47, 358–368 (2016).CAS 

    Google Scholar 
    Shao, T. et al. Evaluation of the effects of different stocking densities on growth and stress responses of juvenile hybrid grouper♀ Epinephelus fuscoguttatus×♂ Epinephelus lanceolatus in recirculating aquaculture systems. J. Fish Biol. 95, 1022–1029 (2019).CAS 
    PubMed 

    Google Scholar 
    Adineh, H., Naderi, M., Hamidi, M. K. & Harsij, M. Biofloc technology improves growth, innate immune responses, oxidative status, and resistance to acute stress in common carp (Cyprinus carpio) under high stocking density. Fish Shellfish Immunol. 95, 440–448 (2019).CAS 
    PubMed 

    Google Scholar 
    Fazelan, Z., Vatnikov, Y. A., Kulikov, E. V., Plushikov, V. G. & Yousefi, M. Effects of dietary ginger (Zingiber officinale) administration on growth performance and stress, immunological, and antioxidant responses of common carp (Cyprinus carpio) reared under high stocking density. Aquaculture 518, 734833 (2020).CAS 

    Google Scholar 
    Hoseini, S. M., Yousefi, M., Hoseinifar, S. H. & Van Doan, H. Effects of dietary arginine supplementation on growth, biochemical, and immunological responses of common carp (Cyprinus carpio L.), stressed by stocking density. Aquaculture 503, 452–459 (2019).CAS 

    Google Scholar 
    Adineh, H., Naderi, M., Nazer, A., Yousefi, M. & Ahmadifar, E. Interactive effects of stocking density and dietary supplementation with nano selenium and garlic extract on growth, feed utilization, digestive enzymes, stress responses, and antioxidant capacity of grass carp, Ctenopharyngodon idella. J. World Aquac. Soc. 52, 789–804 (2021).CAS 

    Google Scholar 
    Zhao, H. et al. Transcriptome and physiological analysis reveal alterations in muscle metabolisms and immune responses of grass carp (Ctenopharyngodon idellus) cultured at different stocking densities. Aquaculture 503, 186–197 (2019).CAS 

    Google Scholar 
    Frisso, R. M., de Matos, F. T., Moro, G. V. & de Mattos, B. O. Stocking density of Amazon fish (Colossoma macropomum) farmed in a continental neotropical reservoir with a net cages system. Aquaculture 529, 735702 (2020).CAS 

    Google Scholar 
    Tammam, M. S., Wassef, E. A., Toutou, M. M. & El-Sayed, A.-F.M. Combined effects of surface area of periphyton substrates and stocking density on growth performance, health status, and immune response of Nile tilapia (Oreochromis niloticus) produced in cages. J. Appl. Phycol. 32, 3419–3428 (2020).CAS 

    Google Scholar 
    Zaki, M. A. A. et al. The impact of stocking density and dietary carbon sources on the growth, oxidative status and stress markers of Nile tilapia (Oreochromis niloticus) reared under biofloc conditions. Aquac. Reports 16, 100282 (2020).
    Google Scholar 
    Rowland, S. J., Mifsud, C., Nixon, M. & Boyd, P. Effects of stocking density on the performance of the Australian freshwater silver perch (Bidyanus bidyanus) in cages. Aquaculture 253, 301–308 (2006).
    Google Scholar 
    Mohler, J. W., King, M. K. & Farrell, P. R. Growth and survival of first-feeding and fingerling Atlantic sturgeon under culture conditions. N. Am. J. Aquac. 62, 174–183 (2000).
    Google Scholar 
    Mirghaed, A. T., Hoseini, S. M. & Ghelichpour, M. Effects of dietary 1, 8-cineole supplementation on physiological, immunological and antioxidant responses to crowding stress in rainbow trout (Oncorhynchus mykiss). Fish Shellfish Immunol. 81, 182–188 (2018).
    Google Scholar 
    Hoseini, S. M., Mirghaed, A. T., Iri, Y. & Ghelichpour, M. Effects of dietary cineole administration on growth performance, hematological and biochemical parameters of rainbow trout (Oncorhynchus mykiss). Aquaculture 495, 766–772 (2018).CAS 

    Google Scholar 
    Barton, B. A., Morgan, J. D. & Vijayan, M. M. Physiological and condition-related indicators of environmental stress in fish. In Biological Indicators of Aquatic Ecosystem Stress (ed. Adams, S. M.) 111–148 (American Fisheries Society, 2002).
    Google Scholar 
    Varela, J. L. et al. Dietary administration of probiotic Pdp11 promotes growth and improves stress tolerance to high stocking density in gilthead seabream Sparus auratus. Aquaculture 309, 265–271 (2010).CAS 

    Google Scholar 
    Costas, B., Aragão, C., Dias, J., Afonso, A. & Conceição, L. E. C. Interactive effects of a high-quality protein diet and high stocking density on the stress response and some innate immune parameters of Senegalese sole Solea senegalensis. Fish Physiol. Biochem. 39, 1141–1151 (2013).CAS 
    PubMed 

    Google Scholar 
    Long, L. et al. Effects of stocking density on growth, stress, and immune responses of juvenile Chinese sturgeon (Acipenser sinensis) in a recirculating aquaculture system. Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 219, 25–34 (2019).CAS 

    Google Scholar 
    Sadhu, N., Sharma, S. R. K., Joseph, S., Dube, P. & Philipose, K. K. Chronic stress due to high stocking density in open sea cage farming induces variation in biochemical and immunological functions in Asian seabass (Lates calcarifer, Bloch). Fish Physiol. Biochem. 40, 1105–1113 (2014).CAS 
    PubMed 

    Google Scholar 
    Zahran, E., Risha, E., AbdelHamid, F., Mahgoub, H. A. & Ibrahim, T. Effects of dietary Astragalus polysaccharides (APS) on growth performance, immunological parameters, digestive enzymes, and intestinal morphology of Nile tilapia (Oreochromis niloticus). Fish Shellfish Immunol. 38, 149–157 (2014).CAS 
    PubMed 

    Google Scholar 
    Aruoma, O. I. Free radicals, oxidative stress, and antioxidants in human health and disease. J. Am. Oil Chem. Soc. 75, 199–212 (1998).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Haridas, H. et al. Enhanced growth and immuno-physiological response of genetically improved farmed Tilapia in indoor biofloc units at different stocking densities. Aquac. Res. 48, 4346–4355 (2017).CAS 

    Google Scholar 
    Ruane, N. M., Carballo, E. C. & Komen, J. Increased stocking density influences the acute physiological stress response of common carp Cyprinus carpio (L.). Aquac. Res. 33, 777–784 (2002).
    Google Scholar 
    Wang, X. et al. Effects of stocking density on growth, nonspecific immune response, and antioxidant status in African catfish (Clarias gariepinus). (2013).Johnson, K. M. & Lema, S. C. Tissue-specific thyroid hormone regulation of gene transcripts encoding iodothyronine deiodinases and thyroid hormone receptors in striped parrotfish (Scarus iseri). Gen. Comp. Endocrinol. 172, 505–517 (2011).CAS 
    PubMed 

    Google Scholar 
    El-Khaldi, A. T. F. Effect of different stress factors on some physiological parameters of Nile tilapia (Oreochromis niloticus). Saudi J. Biol. Sci. 17, 241–246 (2010).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Sharma, A., Devi, S., Singh, K. & Prabhakar, P. K. Correlation of body mass index with thyroid-stimulating hormones in thyroid patient. Asian J. Pharm. Clin. Res. 11, 65–68 (2018).
    Google Scholar 
    Li, D., Liu, Z. & Xie, C. Effect of stocking density on growth and serum concentrations of thyroid hormones and cortisol in Amur sturgeon, Acipenser schrenckii. Fish Physiol. Biochem. 38, 511–520 (2012).CAS 
    PubMed 

    Google Scholar 
    Park, J.-W. et al. The thyroid endocrine disruptor perchlorate affects reproduction, growth, and survival of mosquitofish. Ecotoxicol. Environ. Saf. 63, 343–352 (2006).CAS 
    PubMed 

    Google Scholar 
    Refaey, M. M. et al. High stocking density alters growth performance, blood biochemistry, intestinal histology, and muscle quality of channel catfish Ictalurus punctatus. Aquaculture 492, 73–81 (2018).CAS 

    Google Scholar 
    Reinecke, M. et al. Growth hormone and insulin-like growth factors in fish: where we are and where to go. Gen. Comp. Endocrinol. 142, 20–24 (2005).CAS 
    PubMed 

    Google Scholar 
    Salas-Leiton, E. et al. Dexamethasone modulates expression of genes involved in the innate immune system, growth and stress and increases susceptibility to bacterial disease in Senegalese sole (Solea senegalensis Kaup, 1858). Fish Shellfish Immunol. 32, 769–778 (2012).CAS 
    PubMed 

    Google Scholar 
    Dyer, A. R. et al. Correlation of plasma IGF-I concentrations and growth rate in aquacultured finfish: a tool for assessing the potential of new diets. Aquaculture 236, 583–592 (2004).CAS 

    Google Scholar 
    Kajimura, S. et al. Dual mode of cortisol action on GH/IGF-I/IGF binding proteins in the tilapia, Oreochromis mossambicus. J. Endocrinol. 178, 91–99 (2003).CAS 
    PubMed 

    Google Scholar 
    Ren, Y., Wen, H., Li, Y. & Li, J. Stocking density affects the growth performance and metabolism of Amur sturgeon by regulating expression of genes in the GH/IGF axis. J. Oceanol. Limnol. 36, 956–972 (2018).ADS 
    CAS 

    Google Scholar 
    Salas-Leiton, E. et al. Effects of stocking density and feed ration on growth and gene expression in the Senegalese sole (Solea senegalensis): potential effects on the immune response. Fish Shellfish Immunol. 28, 296–302 (2010).CAS 
    PubMed 

    Google Scholar 
    Vijayan, M. M., Aluru, N. & Leatherland, J. F. Stress response and the role of cortisol. Fish Dis. Disord. 2, 182–201 (2010).
    Google Scholar 
    Hegazi, M. M., Attia, Z. I. & Ashour, O. A. Oxidative stress and antioxidant enzymes in liver and white muscle of Nile tilapia juveniles in chronic ammonia exposure. Aquat. Toxicol. 99, 118–125 (2010).CAS 
    PubMed 

    Google Scholar 
    Kpundeh, M. D., Xu, P., Yang, H., Qiang, J. & He, J. Stocking densities and chronic zero culture water exchange stress’ effects on biological performances, hematological and serum biochemical indices of GIFT tilapia juveniles (Oreochromis niloticus). J. Aquac. Res. Dev. 4, 2 (2013).
    Google Scholar 
    Tan, C. et al. Effects of stocking density on growth, body composition, digestive enzyme levels and blood biochemical parameters of Anguilla marmorata in a recirculating aquaculture system. Turk. J. Fish. Aquat. Sci. 18, 9–16 (2018).
    Google Scholar 
    Ni, M. et al. The physiological performance and immune responses of juvenile Amur sturgeon (Acipenser schrenckii) to stocking density and hypoxia stress. Fish Shellfish Immunol. 36, 325–335 (2014).CAS 
    PubMed 

    Google Scholar 
    Abdel-Tawwab, M. Effects of dietary protein levels and rearing density on growth performance and stress response of Nile tilapia, Oreochromis niloticus (L.). Int. Aquat. Res. 4, 1–13 (2012).
    Google Scholar 
    Chatterjee, N. et al. Effect of stocking density and journey length on the welfare of rohu (Labeo rohita Hamilton) fry. Aquac. Int. 18, 859–868 (2010).
    Google Scholar 
    Pakhira, C., Nagesh, T. S., Abraham, T. J., Dash, G. & Behera, S. Stress responses in rohu, Labeo rohita transported at different densities. Aquac. Rep. 2, 39–45 (2015).
    Google Scholar 
    Tahmasebi-Kohyani, A., Keyvanshokooh, S., Nematollahi, A., Mahmoudi, N. & Pasha-Zanoosi, H. Effects of dietary nucleotides supplementation on rainbow trout (Oncorhynchus mykiss) performance and acute stress response. Fish Physiol. Biochem. 38, 431–440 (2012).CAS 
    PubMed 

    Google Scholar 
    Montero, D. et al. Effect of vitamin E and C dietary supplementation on some immune parameters of gilthead seabream (Sparus aurata) juveniles subjected to crowding stress. Aquaculture 171, 269–278 (1999).CAS 

    Google Scholar 
    Urbinati, E. C., de Abreu, J. S., da Silva Camargo, A. C. & Parra, M. A. L. Loading and transport stress of juvenile matrinxã (Brycon cephalus, Characidae) at various densities. Aquaculture 229, 389–400 (2004).
    Google Scholar 
    Evans, D. H. Cell signaling and ion transport across the fish gill epithelium. J. Exp. Zool. 293, 336–347 (2002).CAS 
    PubMed 

    Google Scholar 
    McCormick, S. D. Endocrine control of osmoregulation in teleost fish. Am. Zool. 41, 781–794 (2001).CAS 

    Google Scholar 
    Postlethwaite, E. & McDonald, D. Mechanisms of Na+ and C-regulation in freshwater-adapted rainbow trout (Oncorhynchus mykiss) during exercise and stress. J. Exp. Biol. 198, 295–304 (1995).CAS 
    PubMed 

    Google Scholar 
    Liu, P., Du, Y., Meng, L., Li, X. & Liu, Y. Metabolic profiling in kidneys of Atlantic salmon infected with Aeromonas salmonicida based on 1H NMR. Fish Shellfish Immunol. 58, 292–301 (2016).CAS 
    PubMed 

    Google Scholar 
    Hosfeld, C. D., Hammer, J., Handeland, S. O., Fivelstad, S. & Stefansson, S. O. Effects of fish density on growth and smoltification in intensive production of Atlantic salmon (Salmo salar L.). Aquaculture 294, 236–241 (2009).
    Google Scholar 
    Wagner, E. I., Miller, S. A. & Bosakowski, T. Ammonia excretion by rainbow trout over a 24-hour period at two densities during oxygen injection. Progress. Fish-Culturist 57, 199–205 (1995).
    Google Scholar 
    Dong, J. et al. Effect of stocking density on growth performance, digestive enzyme activities, and nonspecific immune parameters of Palaemonetes sinensis. Fish Shellfish Immunol. 73, 37–41 (2018).CAS 
    PubMed 

    Google Scholar 
    Wang, Y. et al. Effects of stocking density on the growth performance, digestive enzyme activities, antioxidant resistance, and intestinal microflora of blunt snout bream (Megalobrama amblycephala) juveniles. Aquac. Res. 50, 236–246 (2019).CAS 

    Google Scholar 
    Trenzado, C. E. et al. Effect of dietary lipid content and stocking density on digestive enzymes profile and intestinal histology of rainbow trout (Oncorhynchus mykiss). Aquaculture 497, 10–16 (2018).CAS 

    Google Scholar 
    Li, X., Liu, Y. & Blancheton, J.-P. Effect of stocking density on performances of juvenile turbot (Scophthalmus maximus) in recirculating aquaculture systems. Chin. J. Oceanol. Limnol. 31, 514–522 (2013).ADS 
    CAS 

    Google Scholar 
    Ezhilmathi, S. et al. Effect of stocking density on growth performance, digestive enzyme activity, body composition and gene expression of Asian seabass reared in recirculating aquaculture system. Aquac. Res. https://doi.org/10.1111/are.15725 (2022).Article 

    Google Scholar 
    Bolasina, S., Tagawa, M., Yamashita, Y. & Tanaka, M. Effect of stocking density on growth, digestive enzyme activity and cortisol level in larvae and juveniles of Japanese flounder, Paralichthys olivaceus. Aquaculture 259, 432–443 (2006).CAS 

    Google Scholar 
    Hoseini, S. M., Hoseinifar, S. H. & Van Doan, H. Effect of dietary eucalyptol on stress markers, enzyme activities and immune indicators in serum and haematological characteristics of common carp (Cyprinus carpio) exposed to toxic concentration of ambient copper. Aquac. Res. 49, 3045–3054 (2018).CAS 

    Google Scholar 
    Ni, M. et al. Effects of stocking density on mortality, growth and physiology of juvenile Amur sturgeon (Acipenser schrenckii). Aquac. Res. 47, 1596–1604 (2016).CAS 

    Google Scholar 
    Abdel-Tawwab, M., Hagras, A. E., Elbaghdady, H. A. M. & Monier, M. N. Dissolved oxygen level and stocking density effects on growth, feed utilization, physiology, and innate immunity of Nile Tilapia, Oreochromis niloticus. J. Appl. Aquac. 26, 340–355 (2014).
    Google Scholar 
    Toko, I., Fiogbe, E. D., Koukpode, B. & Kestemont, P. Rearing of African catfish (Clarias gariepinus) and vundu catfish (Heterobranchus longifilis) in traditional fish ponds (whedos): effect of stocking density on growth, production and body composition. Aquaculture 262, 65–72 (2007).
    Google Scholar 
    Suárez, M. D. et al. Influence of dietary lipids and culture density on rainbow trout (Oncorhynchus mykiss) flesh composition and quality parameter. Aquac. Eng. 63, 16–24 (2014).
    Google Scholar 
    Santín, A., Grinyó, J., Bilan, M., Ambroso, S. & Puig, P. First report of the carnivorous sponge Lycopodina hypogea (Cladorhizidae) associated with marine debris, and its possible implications on deep-sea connectivity. Mar. Pollut. Bull. 159, 111501 (2020).PubMed 

    Google Scholar 
    Jørpeland, G., Imsland, A., Stien, L. H., Bleie, H. & Roth, B. Effects of filleting method, stress, storage and season on the quality of farmed Atlantic cod (Gadus morhua L.). Aquac. Res. 46, 1597–1607 (2015).
    Google Scholar 
    Bulow, F. J. RNA-DNA ratios as indicators of growth in fish: a review. In The Age and growth of fish (eds Summerfelt, R. C. & Hall, G. E.) 45–64 (Iowa State University Press, Ames, Iowa, 1987).
    Google Scholar 
    Regnault, M. & Luquet, P. Study by evolution of nucleic acid content of prepuberal growth in the shrimp Crangon vulgaris. Mar. Biol. 25, 291–298 (1974).CAS 

    Google Scholar 
    Tanaka, H. K. M. et al. High resolution imaging in the inhomogeneous crust with cosmic-ray muon radiography: the density structure below the volcanic crater floor of Mt. Asama, Japan. Earth Planet. Sci. Lett. 263, 104–113 (2007).ADS 
    CAS 

    Google Scholar 
    Gwak, W. S. & Tanaka, M. Developmental change in RNA: DNA ratios of fed and starved laboratory-reared Japanese flounder larvae and juveniles, and its application to assessment of nutritional condition for wild fish. J. Fish Biol. 59, 902–915 (2001).CAS 

    Google Scholar 
    Ali, M., Iqbal, R., Rana, S. A., Athar, M. & Iqbal, F. Effect of feed cycling on specific growth rate, condition factor and RNA/DNA ratio of Labeo rohita. African J. Biotechnol. 5, 1551–1556 (2006).CAS 

    Google Scholar 
    Zehra, S. & Khan, M. A. Dietary lysine requirement of fingerling Catla catla (Hamilton) based on growth, protein deposition, lysine retention efficiency, RNA/DNA ratio and carcass composition. Fish Physiol. Biochem. 39, 503–512 (2013).CAS 
    PubMed 

    Google Scholar 
    Misra, H. P. & Fridovich, I. The role of superoxide anion in the autoxidation of epinephrine and a simple assay for superoxide dismutase. J. Biol. Chem. 247, 3170–3175 (1972).CAS 
    PubMed 

    Google Scholar 
    Takahara, S. et al. Hypocatalasemia: a new genetic carrier state. J. Clin. Invest. 39, 610–619 (1960).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Rick, W. & Stegbauer, H. P. α-Amylase measurement of reducing groups. In Methods of Enzymatic Analysis (ed. Bergmeyer, H. S.) 885–890 (Elsevier, 1974).
    Google Scholar 
    Cherry, I. S. & Crandall, L. A. Jr. The specificity of pancreatic lipase: its appearance in the blood after pancreatic injury. Am. J. Physiol. Content 100, 266–273 (1932).CAS 

    Google Scholar 
    Drapeau, G. R. [38] Protease from Staphyloccus aureus. In Methods in Enzymology (eds Jura, N. & Murphy, J. M.) 469–475 (Elsevier, 1976).
    Google Scholar 
    AOAC. Official Methods of Analysis of AOAC International. (Association of Official Analytical Chemists Washington, DC, 2005).Bosworth, B. G., Small, B. C. & Mischke, C. Effects of transport water temperature, aerator type, and oxygen level on channel catfish Ictalurus punctatus fillet quality. J. World Aquac. Soc. 35, 412–419 (2004).
    Google Scholar 
    Ma, L. Q., Qi, C. L., Cao, J. J. & Li, D. P. Comparative study on muscle texture profile and nutritional value of channel catfish (Ictalurus punctatus) reared in ponds and reservoir cages. J. Fish. China 38, 532–537 (2014).
    Google Scholar 
    APHA. Standard Methods for the Examination of Water and Wastewater. (American Public Health Association, American Water Works Association, Water Environment Federation, 2012). More

  • in

    Aurochs roamed along the SW coast of Andalusia (Spain) during Late Pleistocene

    Theodor, J. M., Erfort, J. & Métais, G. The earliest artiodactyls: Diacodexeidae, Dichobunidae, Homacodontidae, Leptochoeridae and Raoellidae. in Evolution of Artiodactyls (eds. Prothero, D.R. & Foss, S. E.). 32–58. (Johns Hopkins University, 2007).Badiola, A. et al. The role of new Iberian finds in understanding European Eocene mammalian paleobiogeography. Geol. Acta. 7(1–2), 243–258 (2009).
    Google Scholar 
    Boivin, M. et al. New material of Diacodexis (Mammalia, Artiodactyla) from the early Eocene of Southern Europe. Geobios 51(4), 285–306 (2018).Article 

    Google Scholar 
    Ellenberger, P. Sur les empreintes de pas des gros mammiféres de l’Eocene supérieur de Garrigues-Ste-Eulalie (Gard). Palaeovertebr. Mém. Jubil. R. Lavocat. 13, 37–78 (1980).
    Google Scholar 
    Santamaría, R. L. G. & Casanovas-Cladellas, M. L. Nuevos yacimientos con icnitas de mamíferos del Oligoceno de los alrededores de Agramunt (Lleida, España). Paleont. Evol. 23, 141–152 (1990).
    Google Scholar 
    Sarjeant, W. A. S. & Langston, W. Jr. Vertebrate footprints and invertebrate traces from the Chadronian (Late Eocene) of Trans-Pecos. Texas. Mem. Mus. Bull. 36, 1–86 (1994).
    Google Scholar 
    Costeur, L., Balme, C. & Legal, S. Early Oligocene mammal tracks from southeastern France. Hist. Biol. 16(4), 257–267. https://doi.org/10.1080/10420940902953197 (2009).Article 

    Google Scholar 
    Wroblewski, A.F.-J. & Gulas-Wroblewski, B. E. Earliest evidence of marine habitat use by mammals. Sci. Rep. 11, 8846. https://doi.org/10.1038/s41598-021-88412-3 (2021).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Fornós, J. J. & Pons-Moya, J. Icnitas de Myotragus balearicus del yacimiento de Ses Piquetes (Santanyi, Mallorca). Bol. Soc. Hist. Nat. Balears 26, 135–144 (1982).
    Google Scholar 
    Flor, G. Estructuras de deformación por pisadas de cérvidos en la duna cementada de Gorliz (Vizcaya, N de España). Rev. Soc. Geol. Esp. 2(1–2), 23–29 (1989).
    Google Scholar 
    Fornós, J. J., Bromley, R. G., Clemmensen, L. B. & Rodríguez-Perea, A. Tracks and trackways of Myotragus balearicus Bate (Artiodactyla, Caprinae) in Pleistocene aeolianites from Mallorca (Balearic Islands, Western Mediterranean). Palaeogr. Palaeocl. Palaeoecol. 180, 277–313 (2002).ADS 
    Article 

    Google Scholar 
    Neto de Carvalho, C. Vertebrate tracksites from the Mid-Late Pleistocene eolianites of Portugal: The first record of elephant tracks in Europe. Geol. Q. 53(4), 407–414 (2009).
    Google Scholar 
    Neto de Carvalho, C., Saltão, S., Ramos, J. C. & Cachão, M. Pegadas de Cervus elaphus nos eolianitos plistocénicos da ilha do Pessegueiro (SW Alentejano, Portugal). Ciênc. Terra 5, 36–40 (2003).
    Google Scholar 
    Neto de Carvalho, C., Figueiredo, S. & Belo, J. Vertebrate tracks and trackways from the Pleistocene eolianites of SW Portugal. Commun. Geol. 103(1), 101–116 (2016).CAS 

    Google Scholar 
    Neto de Carvalho, C. et al. Paleoecological implications of large-sized wild boar tracks recorded during the Last Interglacial (MIS 5) at Huelva (SW Spain). Palaios https://doi.org/10.2110/palo.2020.058 (2020).Article 

    Google Scholar 
    Neto de Carvalho, C. et al. First vertebrate tracks and palaeoenvironment in a MIS-5 context in the Doñana National Park (Huelva, SW Spain). Quat. Sci. Rev. https://doi.org/10.1016/j.quascirev.2020.106508 (2020).Article 

    Google Scholar 
    Cardoso, J. L. Les grands mammifères du Pléistocène supérieur du Portugal. Essai de synthése. Geobios 29(2), 235–250 (1996).Article 

    Google Scholar 
    Sala, M. T. N., Pantoja, A., Arsuaga, J. L. & Algaba, M. Presencia de bisonte (Bison priscus Bojanus, 1827) y uro (Bos primigenius Bojanus, 1827) en las cuevas del Búho y de la Zarzamora (Segovia, España). Munibe 61, 43–55 (2010).
    Google Scholar 
    Figueiredo, S. D. & Sousa, M. F. O registo de bovídeos plistocénicos em Portugal. in Livro de Resumos das IV Jornadas de Arqueologia do Vale do Tejo. Vol. 10. (Centro Português de Geo-História e Pré-História, 2017).Barr, K. & Bell, M. Neolithic and Bronze age ungulate footprint-tracks of the Severn Estuary: Species, age, identification and the interpretation of husbandry practices. Environ. Archaeol. 22(1), 1–15 (2017).Article 

    Google Scholar 
    Bell, M. Making One’s Way in the World (Oxbow Books, 2020).Book 

    Google Scholar 
    Díaz-Martínez, I. et al. Multi-aged social behavior based on artiodactyl tracks in an early Miocene palustrine wetland (Ebro Basin, Spain). Sci. Rep. 10, 1099. https://doi.org/10.1038/s41598-020-57438-4 (2020).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Quintana, J. Descripción de un rastro de Myotragus e icnitas de Hypnomys del yacimiento cuaternario de Ses Penyes d’es Perico (Ciutadella de Menorca, Balears). Paleont. Evol. 26–27, 271–279 (1993).
    Google Scholar 
    Muñiz, F. et al. Following the last Neanderthals: Mammal tracks in Late Pleistocene coastal dunes of Gibraltar (S Iberian Peninsula). Quat. Sci. Rev. 217, 297–309 (2019).ADS 
    Article 

    Google Scholar 
    Altuna, J. Fauna de mamíferos de los yacimientos prehistóricos de Guipúzcoa. Con catálogo de los mamíferos cuaternarios del Cantábrico y del Pirineo occidental. Munibe 24, 1–464 (1972).
    Google Scholar 
    López González, F., Vila Taboada, M. & Grandal d’Anglade. Sobre los grandes bóvidos pleistocenos (Bovidae, Mammalia) en el NO de la Península Ibérica. Cad. Lab. Xeol. Laxe 24, 57–71 (1999).Sommer, R. S., Kalbe, J., Ekström, J., Benecke, N. & Liljengren, R. Range dynamics of the reindeer in Europe during the last 25,000 years. J. Biogeogr. 41, 298–306. https://doi.org/10.1111/jbi.12193 (2014).Article 

    Google Scholar 
    Whittle, A., Antoine, S., Gardiner, N., Milles, A. & Webster, A. Two Later Bronze Age occupations and an Iron Age channel on the Gwent foreshore. Bull. Board Celt. Stud. 36, 200–223 (1989).
    Google Scholar 
    Aldhouse-Green, S. et al. Prehistoric human footprints from the Severn Estuary at Uskmouth and Magor Pill, Gwent, Wales. Archae. Cambr. 141, 4–55 (1992).
    Google Scholar 
    Allen, J. R. L. Subfossil mammalian tracks (Flandrian) in the Severn Estuary, S.W. Britain: Mechanics of formation, preservation and distribution. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 352(1352), 481–518 (1997).ADS 
    PubMed Central 
    Article 

    Google Scholar 
    Bell, M. Prehistoric coastal communities: the Mesolithic in western Britain. in CBA Research Report. Vol. 149. (Council for British Archaeology, 2007).Bell, M. The Bronze Age in the Severn estuary. in Research Report. Vol. 172. (Council for British Archaeology, 2013).Scales, R. Footprint tracks of people and animals. in Prehistoric Coastal Communities: The Mesolithic in Western Britain (ed. Bell, M.). Vol. 149. 139–147. CBA Research Report 149. (Council of British Archaeology, 2007).Roberts, G. Ephemeral, subfossil mammalian, avian and hominid footprints within Flandrian sediment exposures at Formby Point, Sefton Coast, North West England. Ichnos 16, 33–48 (2009).Article 

    Google Scholar 
    Waddington, C. Low Hauxley, Northumberland: A review of archaeological interventions and site condition. Archael. Res. Serv. 2010/25 (2010).Eadie, G. & Waddington, C. Rescue recording of an eroding inter-tidal peat bed at Lower Hauxley, Northumberland (6109). Archael. Res. Serv. 2013/17 (2013).Burns, A. The prehistoric footprints at Formby. in Sefton Coast Landscape Partnership Scheme (2014).Pandolfi, L., Petronio, C. & Salari, L. Bos primigenius Bojanus, 1827 from the Early Late Pleistocene deposit of Avetrana (southern Italy) and the variation in size of the species in southern Europe: Preliminary report. J. Geol. Res. https://doi.org/10.1155/2011/245408 (2011).Article 

    Google Scholar 
    Currant, A. P. A review of the Quaternary mammals of Gibraltar. in Neanderthals on the Edge: 150th Anniversary Conference of the Forbes’ Quarry Discovery, Gibraltar (eds. Stringer, C. B., Barton, R. N. E. & Finlayson, J.C.). 201–206. (Oxbow, 2000).Penela, A. J. M. Los grandes mamíferos del yacimiento acheulense de la Solana del Zamborino, Fonelas (Granada, España). Antr. Paleoecol. Hum. 5, 29–187 (1988).
    Google Scholar 
    Bataille, G. Prehistoric Painting. Lascaux or the Birth of Art (MacMillan, 1980).
    Google Scholar 
    Zazo, C. et al. Palaeoenvironmental evolution of the Barbate-Trafalgar coast (Cadiz) during the last ~140 ka: Climate, sea-level interactions and tectonics. Geomorphology 100, 212–222 (2008).ADS 
    Article 

    Google Scholar 
    Zazo, C. et al. Landscape evolution and geodynamic controls in the Gulf of Cadiz (Huelva coast, SW Spain) during the Late Quaternary. Geomorphology 68, 269–290. https://doi.org/10.1016/j.geomorph.2004.11.022 (2005).ADS 
    Article 

    Google Scholar 
    García de Domingo, A., González Lastra, J., Hernaiz Huerta, P. P., Zazo Cardeña, C. & Goy Goy, J. L. Mapa Geológico de la Hoja No. 1073 (Vejer de la Frontera). Mapa Geológico de España a Escala 1:50.000. Segunda Serie (MAGNA). http://info.igme.es/cartografiadigital/geologica/Magna50Hoja.aspx?Id=1073&language=es (©Instituto Geológico y Minero de España (IGME), 1990).Demathieu, G., Ginsburg, L., Guérin, C. & Truc, G. Étude paléontologique, ichnologique et paléoécologique du gisêment oligocène de Saignon (bassin d’Apt, Vaucluse). Bull. Mus. Natl. Hist. Nat. 6(2), 153–183 (1984).
    Google Scholar 
    Bang, P. & Dahlstrøm, P. Animal Tracks and Signs (Oxford University Press, 2001).
    Google Scholar 
    Wright, E. The History of the European Aurochs (Bos primigenius) from the Middle Pleistocene to Its Extinction: An Archaeological Investigation of Its Evolution, Morphological Variability and Response to Human Exploitation. (PhD. Thesis, University of Sheffield, 2013).Koenigswald, W. V., Sander, P. M. & Walders, M. The Upper Pleistocene tracksite Bottrop-Welheim (Germany). Acta Zool. Cracov. 39(1), 235–244 (1996).
    Google Scholar 
    Martínez-Navarro, B., Rook, L., Papini, M. & Libsekal, Y. A new species of bull from the Early Pleistocene paleoanthropological site of Buia (Eritrea): Parallelism on the dispersal of the genus Bos and the Acheulian culture. Quat. Intern. 212(2), 169–175. https://doi.org/10.1016/j.quaint.2009.09.003 (2010).Article 

    Google Scholar 
    Van Vuure, C. Retracing the Aurochs: History, Morphology and Ecology of an Extinct Ox (Coronet Books, 2005).
    Google Scholar 
    Franks, J. W. Interglacial deposits at Trafalgar Square, London. N. Phytologist 59(2), 145–152 (1960).Article 

    Google Scholar 
    Estévez, J. & Saña, M. Auerochsenfunde auf der Iberischen Halbinsel. in Archäologie und Biologie des Auerochsen (ed. Weniger, G.-C.) (Neanderthal Museum, 1999).Mona, S. et al. Population dynamic of the extinct European aurochs: Genetic evidence of a north-south differentiation pattern and no evidence of post-glacial expansion. BMC Evol. Biol. 10, 1–13 (2010).Article 
    CAS 

    Google Scholar 
    Rodríguez-Vidal, J. et al. Undrowning a lost world—The Marine isotope stage 3 landscape of Gibraltar. Geomorphology 203, 105–114 (2013).ADS 
    Article 

    Google Scholar 
    Pfeiffer, T. Systematic relationship between the Bovini with special references to the fossil taxa Bos primigenius Bojanus and Bison priscus Bojanus. in Archäologie und Biologie des Auerochsen (ed. Weniger, G.-C.). 59–70. (Neanderthal Museum, 1999).Zazula, G. D. et al. A late Pleistocene steppe bison (Bison priscus) partial carcass from Tsiigehtchic, Northwest Territories, Canada. Quat. Sci. Rev. 28(25–26), 2734–2742 (2009).ADS 
    Article 

    Google Scholar 
    Boeskorov, G. G. et al. The Yukagir Bison: The exterior morphology of a complete frozen mummy of the extinct steppe bison, Bison priscus from the early Holocene of northern Yakutia, Russia. Quat. Intern. 406, 94–110. https://doi.org/10.1016/j.quaint.2015.11.084 (2016).Article 

    Google Scholar 
    Ekström, J. The Late Quaternary History of the Urus (Bos primigenius Bojanus 1827) in Sweden. PhD. Thesis. (Lund University, 1993).Grange, T. et al. The evolution and population diversity of Bison in Pleistocene and Holocene Eurasia: Sex matters. Diversity 10(3), 65. https://doi.org/10.3390/d10030065 (2018).Article 

    Google Scholar 
    Castaños, J., Castaños, P. & Murelaga, X. First complete skull of a Late Pleistocene Steppe Bison (Bison priscus) in the Iberian Peninsula. Ameghiniana 53(5), 543–551. https://doi.org/10.5710/amgh.03.06.2016.2995 (2016).Article 

    Google Scholar 
    Álvarez-Lao, D. J., Kahlke, R.-D., García, N. & Mol, D. The Padul mammoth finds: On the southernmost record of Mammuthus primigenius in Europe and its southern spread during the Late Pleistocene. Palaeogeogr. Palaeocl. Palaeoecol. 278(1–4), 57–70 (2009).ADS 
    Article 

    Google Scholar 
    Loope, D. B. Recognizing and utilizing vertebrate tracks in cross section: Cenozoic hoofprints from Nebraska. Palaios 1, 141–151 (1986).ADS 
    Article 

    Google Scholar 
    Albarella, U., Dobney, K. & Rowley-Conwy, P. Size and shape of the Eurasian wild boar (Sus scrofa), with a view to the reconstruction of its Holocene history. Environ. Archaeol. 14, 103–136 (2009).Article 

    Google Scholar 
    Davis, S. J. M. The effects of temperature change and domestication on the body size of Late Pleistocene to Holocene mammals of Israel. Palaeobiology 7, 101–114 (1981).Article 

    Google Scholar 
    Cerilli, E. & Petronio, C. Biometrical variations of Bos primigenius Bojanus 1827 from middle Pleistocene to Holocene. in Proceedings of the International Symposium on ‘Ongulés/Ungulates’, Toulouse. 37–42. (1991).Davis, S. J. M. & Mataloto, R. Animal remains from Chalcolithic of São Pedro (Redondo, Alentejo): Evidence for a crisis in the Mesolithic. Rev. Port. Arqueol. 15, 47–85 (2012).
    Google Scholar 
    Mariezkurrena, K. & Altuna, J. Biometría y diformismo sexual en el esqueleto de Cervus elaphus würmiense, postwürmiense y actual del Cantábrico. Munibe (Antr.-Arkeol.) 35, 203–246 (1983).
    Google Scholar 
    Davis, S. J. M. The mammals and birds from the Gruta do Caldeirão, Portugal. Rev. Port. Arqueol. 5, 29–98 (2002).CAS 

    Google Scholar 
    Barr, K. Prehistoric Avian, Mammalian and H. sapiens Footprint—Tracks from Intertidal Sediments as Evidence of Human Palaeoecology. PhD. Thesis. (University of Reading, 2018).Hall, J. G. A comparative analysis of the habitat of the extinct aurochs and other prehistoric mammals in Britain. Ecography 31, 187–190 (2008).Article 

    Google Scholar 
    Bicho, N. F., Gibaja, J. F., Stiner, M. & Manne, T. L. Paléolithique supérieur au sud du Portugal: Le site du Vale do Boi. L’antropologie 114, 48–67 (2010).
    Google Scholar 
    Bicho, N. & Haws, J. The Magdelian in central and southern Portugal: Human ecology at the end of the Pleistocene. Quatern. Int. 272–273, 6–16 (2012).Article 

    Google Scholar 
    Cortés-Sánchez, M. et al. Palaeoenvironmental and cultural dynamics of the coast of Málaga (Andalusia, Spain) during the Upper Pleistocene and early Holocene. Quatern. Sci. Rev. 27, 2176–2193 (2008).ADS 
    Article 

    Google Scholar 
    Bohórquez, A. M., Ruiz, C. B., Caparrós, M. & Moigne, A. M. Una aproximación a la compreensión de la fauna de macromamiferos de la Cueva de Zafarraya (Alcaucín, Málaga). Menga Rev. Prehist. Andalucía 3, 83–105 (2012).
    Google Scholar 
    Ripoll, M. P. & Maroto, J. L. fauna mediterránea durante el Pleistoceno superior del Mediterráneo Ibérico. Kobie Serie Anejo 18, 27–38 (2021).
    Google Scholar 
    Lazo, A. Ranging behaviour of feral cattle (Bos taurus) in Doñana National Park, S.W. Spain. J. Zool. 236(3), 359–369. https://doi.org/10.1111/j.1469-7998.1995.tb02718.x (1995).Article 

    Google Scholar 
    AliceVision. Meshroom: V2021.1.0. GNU-GPL. https://alicevision.org/ (2020).OpenDroneMap Authors ODM. A Command Line Toolkit to Generate Maps, Point Clouds, 3D Models and DEMs from Drone, Balloon or Kite Images. OpenDroneMap/ODM GitHub Page. https://github.com/OpenDroneMap/WebODM (2020).Cignoni, P., Callieri, M., Corsini, M., Dellepiane, M., Ganovelli, F. & Ranzuglia, G. MeshLab: an open-source mesh processing tool. in Sixth Eurographics Italian Chapter Conference. 129–136. MeshLab V. 2020.12. https://www.meshlab.net/ (2008).CloudCompare. V2.11.0. GNU-GPL. https://www.cloudcompare.org (2020).Zhukov, S., Iones, A. & Kronin, G. An ambient light illumination model. Render. Tech. 98, 45–55 (1998).Article 

    Google Scholar 
    Vergne, R., Pacanowski, R., Barla, P., Granier, X., & Schlick, C. Radiance scaling for versatile surface enhancement. in Proceedings of the 2010 ACMSIGGRAPH Symposium on Interactive 3D Graphics and Games.143–150. (2010). More

  • in

    Microbial isolates with Anti-Pseudogymnoascus destructans activities from Western Canadian bat wings

    Frick, W. F. et al. An emerging disease causes regional population collapse of a common North American bat species. Science 329, 679–682 (2010).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Froschauer, A. & Coleman, J. North American bat death toll exceeds 5.5 million from white-nose syndrome. Biol. Rep. US Fish Wildl. Serv. 2, 1–2 (2012).
    Google Scholar 
    Blehert, D. S. et al. Bat white-nose syndrome: An emerging fungal pathogen?. Science 323, 227 (2009).CAS 
    PubMed 
    Article 

    Google Scholar 
    Meteyer, C. U. et al. Histopathologic criteria to confirm white-nose syndrome in bats. J. Vet. Diagn. Invest. 21, 411–414 (2009).PubMed 
    Article 

    Google Scholar 
    O’Donoghue, A. J. et al. Destructin-1 is a collagen-degrading endopeptidase secreted by Pseudogymnoascus destructans, the causative agent of white-nose syndrome. Proc. Natl. Acad. Sci. USA. 112, 7478–7483 (2015).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Cryan, P. M., Meteyer, C. U., Boyles, J. G. & Blehert, D. S. Wing pathology of white-nose syndrome in bats suggests life-threatening disruption of physiology. BMC Biol. 8, 135 (2010).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Warnecke, L. et al. Pathophysiology of white-nose syndrome in bats: A mechanistic model linking wing damage to mortality. Biol. Lett. 9, 20130177 (2013).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Verant, M. L., Boyles, J. G., Waldrep, W., Wibbelt, G. & Blehert, D. S. Temperature-dependent growth of Geomyces destructans, the fungus that causes bat white-nose syndrome. PLoS ONE 7, e46280 (2012).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Field, K. A. et al. The white-nose syndrome transcriptome: Activation of anti-fungal host responses in wing tissue of hibernating little brown Myotis. PLoS Pathog. 11, e1005168 (2015).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Boyles, J. G. & Willis, C. K. R. Could localized warm areas inside cold caves reduce mortality of hibernating bats affected by white-nose syndrome?. Front. Ecol. Environ. 8, 92–98 (2010).Article 

    Google Scholar 
    Storm, J. J. & Boyles, J. G. Body temperature and body mass of hibernating little brown bats Myotis lucifugus in hibernacula affected by white-nose syndrome. Acta Theriol. 56, 123–127 (2011).Article 

    Google Scholar 
    Lorch, J. M. et al. First detection of bat white-nose syndrome in western North America. MSphere 1, 4 (2016).Article 
    CAS 

    Google Scholar 
    White-Nose Syndrome Response Team. Where is WNS Now? White-Nose Syndrome https://www.whitenosesyndrome.org/spreadmap (2021).Turner, G. G., Reeder, D. & Coleman, J. T. H. A five-year assessment of mortality and geographic spread of white-nose syndrome in north American bats, with a look at the future: update of white-nose syndrome in bats. Bat Res. News 52, 13 (2011).
    Google Scholar 
    Dzal, Y., McGuire, L. P., Veselka, N. & Fenton, M. B. Going, going, gone: The impact of white-nose syndrome on the summer activity of the little brown bat (Myotis lucifugus). Biol. Lett. 7, 392–394 (2011).PubMed 
    Article 

    Google Scholar 
    Ingersoll, T. E., Sewall, B. J. & Amelon, S. K. Improved analysis of long-term monitoring data demonstrates marked regional declines of bat populations in the eastern United States. PLoS ONE 8, e65907 (2013).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Vanderwolf, K. J. & McAlpine, D. F. Hibernacula microclimate and declines in overwintering bats during an outbreak of white-nose syndrome near the northern range limit of infection in North America. Ecol. Evol. 11, 2273–2288 (2021).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Boyles, J. G., Cryan, P. M., McCracken, G. F. & Kunz, T. H. Economic importance of bats in agriculture. Science 332, 41–42 (2011).ADS 
    PubMed 
    Article 

    Google Scholar 
    Kunz, T. H., de Torrez, E. B., Bauer, D., Lobova, T. & Fleming, T. H. Ecosystem services provided by bats. Ann. N. Y. Acad. Sci. 1223, 1–38 (2011).ADS 
    PubMed 
    Article 

    Google Scholar 
    Puig‐Montserrat, X. & Flaquer, C. Bats actively prey on mosquitoes and other deleterious insects in rice paddies: Potential impact on human health and agriculture. Pest Manag. Sci. (2020).Micalizzi, E. W. & Smith, M. L. Volatile organic compounds kill the white-nose syndrome fungus, Pseudogymnoascus destructans, in hibernaculum sediment. Can. J. Microbiol. 66, 593–599 (2020).CAS 
    PubMed 
    Article 

    Google Scholar 
    Padhi, S., Dias, I., Korn, V. & Bennett, J. Pseudogymnoascus destructans: Causative agent of white-nose syndrome in bats is inhibited by safe volatile organic compounds. Journal of Fungi 4, 48 (2018).PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Chaturvedi, S. et al. Antifungal testing and high-throughput screening of compound library against Geomyces destructans, the etiologic agent of geomycosis (WNS) in bats. PLoS ONE 6, e17032 (2011).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Cornelison, C. T. et al. A preliminary report on the contact-independent antagonism of Pseudogymnoascus destructans by Rhodococcus rhodochrous strain DAP96253. BMC Microbiol. 14, 246 (2014).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Boire, N. et al. Potent inhibition of Pseudogymnoascus destructans, the causative agent of white-nose syndrome in bats, by cold-pressed, terpeneless, Valencia orange oil. PLoS ONE 11, 1–10 (2016).Article 
    CAS 

    Google Scholar 
    Padhi, S., Dias, I. & Bennett, J. W. Two volatile-phase alcohols inhibit growth of Pseudogymnoascus destructans, causative agent of white-nose syndrome in bats. Mycology 8, 11–16 (2017).CAS 
    Article 

    Google Scholar 
    Raudabaugh, D. B. & Miller, A. N. Effect of Trans, trans-farnesol on Pseudogymnoascus destructans and several closely related species. Mycopathologia 180, 325–332 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    Kulhanek. The Application of Chitosan on an Experimental Infection of Pseudogymnoascus Destructans Increases Survival in Little Brown Bats. (Western Michigan University, 2016).Ghosh, S. et al. Evidence for Anti-Pseudogymnoascus destructans (Pd) activity of propolis. Antibiotics 7, 2 (2017).PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Bernard, R. F. & Grant, E. H. C. Identifying common decision problem elements for the management of emerging fungal diseases of wildlife. Soc. Nat. Resour. (2019).Haas, D. & Défago, G. Biological control of soil-borne pathogens by fluorescent pseudomonads. Nat. Rev. Microbiol. 3, 307–319 (2005).CAS 
    PubMed 
    Article 

    Google Scholar 
    Becker, M. H. & Harris, R. N. Cutaneous bacteria of the redback salamander prevent morbidity associated with a lethal disease. PLoS ONE 5, e10957 (2010).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Gerritsen, J., Smidt, H., Rijkers, G. T. & de Vos, W. M. Intestinal microbiota in human health and disease: The impact of probiotics. Genes Nutr. 6, 209–240 (2011).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Bletz, M. C. et al. Mitigating amphibian chytridiomycosis with bioaugmentation: Characteristics of effective probiotics and strategies for their selection and use. Ecol. Lett. 16, 807–820 (2013).PubMed 
    Article 

    Google Scholar 
    Becker, M. H. et al. Composition of symbiotic bacteria predicts survival in Panamanian golden frogs infected with a lethal fungus. Proc. Biol. Sci. 282, 2881 (2015).
    Google Scholar 
    Hamm, P. S. et al. Western bats as a reservoir of novel Streptomyces species with antifungal activity. Appl. Environ. Microbiol. 83, 1–10 (2017).Article 

    Google Scholar 
    Hoyt, J. R. et al. Bacteria isolated from bats inhibit the growth of Pseudogymnoascus destructans, the causative agent of white-nose syndrome. PLoS ONE https://doi.org/10.1371/journal.pone.0121329 (2015).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Cheng, T. L. et al. Efficacy of a probiotic bacterium to treat bats affected by the disease white-nose syndrome. J. Appl. Ecol. 54, 701–708 (2017).Article 

    Google Scholar 
    Berg, G. & Smalla, K. Plant species and soil type cooperatively shape the structure and function of microbial communities in the rhizosphere. FEMS Microbiol. Ecol. 68, 1–13 (2009).CAS 
    PubMed 
    Article 

    Google Scholar 
    Teplitski, M. & Ritchie, K. How feasible is the biological control of coral diseases?. Trends Ecol. Evol. 24, 378–385 (2009).PubMed 
    Article 

    Google Scholar 
    Clay, K. EDITORIAL: Defensive symbiosis: A microbial perspective. Funct. Ecol. 28, 293–298 (2014).Article 

    Google Scholar 
    Grice, E. A. & Segre, J. A. The skin microbiome. Nat. Rev. Microbiol. 9, 244–253 (2011).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Jani, A. J. & Briggs, C. J. The pathogen Batrachochytrium dendrobatidis disturbs the frog skin microbiome during a natural epidemic and experimental infection. Proc. Natl. Acad. Sci. USA. 111, E5049–E5058 (2014).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Lemieux-Labonté, V., Simard, A., Willis, C. K. R. & Lapointe, F.-J. Enrichment of beneficial bacteria in the skin microbiota of bats persisting with white-nose syndrome. Microbiome 5, 115 (2017).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Walke, J. B. et al. Most of the dominant members of amphibian skin bacterial communities can be readily cultured. Appl. Environ. Microbiol. 81, 6589–6600 (2015).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Avena, C. V. et al. Deconstructing the bat skin microbiome: Influences of the host and the environment. Front. Microbiol. 7, 1753 (2016).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Loudon, A. H. et al. Microbial community dynamics and effect of environmental microbial reservoirs on red-backed salamanders (Plethodon cinereus). ISME J. 8, 830–840 (2014).CAS 
    PubMed 
    Article 

    Google Scholar 
    Walke, J. B. et al. Amphibian skin may select for rare environmental microbes. ISME J. 8, 2207–2217 (2014).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Loudon, A. H. et al. Vertebrate hosts as islands: Dynamics of selection, immigration, loss, persistence, and potential function of bacteria on salamander skin. Front. Microbiol. 7, 333 (2016).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Winter, A. S. et al. Skin and fur bacterial diversity and community structure on American southwestern bats: Effects of habitat, geography and bat traits. PeerJ 5, e3944 (2017).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Perofsky, A. C., Lewis, R. J., Abondano, L. A., Di Fiore, A. & Meyers, L. A. Hierarchical social networks shape gut microbial composition in wild Verreaux’s sifaka. Proc. Biol. Sci. 284, 2274 (2017).
    Google Scholar 
    Raulo, A. et al. Social behaviour and gut microbiota in red-bellied lemurs (Eulemur rubriventer): In search of the role of immunity in the evolution of sociality. J. Anim. Ecol. 87, 388–399 (2018).PubMed 
    Article 

    Google Scholar 
    Tung, J. et al. Social networks predict gut microbiome composition in wild baboons. Elife 4, 5224 (2015).
    Google Scholar 
    Kolodny, O. et al. Coordinated change at the colony level in fruit bat fur microbiomes through time. Nat. Ecol. Evol. 3, 116–124 (2019).PubMed 
    Article 

    Google Scholar 
    Vuong, H. E., Yano, J. M., Fung, T. C. & Hsiao, E. Y. The microbiome and host behavior. Annu. Rev. Neurosci. 40, 21–49 (2017).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Lausen, C. L., Nagorsen, D. N., Brigham, R. M. & Hobbs, J. Bats of British Columbia 2nd edn. (Royal BC Museum, 2022).
    Google Scholar 
    Spring Cleaning: Why Wash a Bridge? https://www.tranbc.ca/2011/06/22/spring-cleaning-why-wash-a-bridge/ (2012).Maron, P.-A. et al. High microbial diversity promotes soil ecosystem functioning. Appl. Environ. Microbiol. 84, 9 (2018).Article 

    Google Scholar 
    Wagg, C., Bender, S. F., Widmer, F. & van der Heijden, M. G. A. Soil biodiversity and soil community composition determine ecosystem multifunctionality. Proc. Natl. Acad. Sci. USA. 111, 5266–5270 (2014).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Green, S. R. & Gray, P. P. A differential procedure for bacteriological studies useful in the fermentation industry. Arch. Biochem. Biophys. 32, 59–69 (1951).CAS 
    PubMed 
    Article 

    Google Scholar 
    Basu, S. et al. Evolution of bacterial and fungal growth media. Bioinformation 11, 182–184 (2015).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Medina, D. et al. Culture media and individual hosts affect the recovery of culturable bacterial diversity from amphibian skin. Front. Microbiol. 8, 1574 (2017).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Piovia-Scott, J. et al. Greater species richness of bacterial skin symbionts better suppresses the amphibian fungal pathogen Batrachochytrium dendrobatidis. Microb. Ecol. 74, 217–226 (2017).PubMed 
    Article 

    Google Scholar 
    Moeller, A. H. et al. Dispersal limitation promotes the diversification of the mammalian gut microbiota. Proc. Natl. Acad. Sci. USA. 114, 13768–13773 (2017).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Ingala, M. R. et al. Comparing microbiome sampling methods in a wild mammal: Fecal and intestinal samples record different signals of host ecology, evolution. Front. Microbiol. 9, 1–10 (2018).Article 

    Google Scholar 
    Lewis, S. E. Night roosting ecology of pallid bats (Antrozous pallidus) in oregon. Am. Midl. Nat. 132, 219–226 (1994).Article 

    Google Scholar 
    Hershey, O. S. & Barton, H. A. The microbial diversity of caves. Cave Ecol. 1, 69–90. https://doi.org/10.1007/978-3-319-98852-8_5 (2018).Article 

    Google Scholar 
    British Columbia Government Mineral Inventory. https://www2.gov.bc.ca/gov/content/industry/mineral-exploration-mining/british-columbia-geological-survey/mineralinventory (2018).Weller, T. J. et al. A review of bat hibernacula across the western United States: Implications for white-nose syndrome surveillance and management. PLoS ONE 13, e0205647 (2018).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Nagorsen, D. W., Brigham, R. M., Royal British Columbia Museum. Bats of British Columbia (UBC Press, 1993).
    Google Scholar 
    Fenton, M. B., Merriam, H. G. & Holroyd, G. L. Bats of Kootenay, Glacier, and Mount Revelstoke national parks in Canada: Identification by echolocation calls, distribution, and biology. Can. J. Zool. 61, 2503–2508 (1983).Article 

    Google Scholar 
    Bernard, R. F., Foster, J. T., Willcox, E. V., Parise, K. L. & McCracken, G. F. Molecular detection of the causative agent of white-nose syndrome on rafinesque’s big-eared bats (Corynorhinus rafinesquii) and two species of migratory bats in the Southeastern USA. J. Wildl. Dis. 51, 519–522 (2015).PubMed 
    Article 

    Google Scholar 
    Lutz, H. L. et al. Ecology and host identity outweigh evolutionary history in shaping the bat microbiome. MSystems 4, 1–10 (2019).
    Google Scholar 
    Gaona, O., Gómez-Acata, E. S., Cerqueda-García, D., Neri-Barrios, C. X. & Falcón, L. I. Fecal microbiota of different reproductive stages of the central population of the lesser-long nosed bat, Leptonycteris yerbabuenae. PLoS ONE 14, e0219982 (2019).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Voigt, C. C., Caspers, B. & Speck, S. Bats, bacteria, and bat smell: Sex-specific diversity of microbes in a sexually selected scent organ. J. Mammal. 86, 745–749 (2005).Article 

    Google Scholar 
    Gharout-Sait, A. et al. Occurrence of carbapenemase-producing Klebsiella pneumoniae in bat guano. Microb. Drug Resist. 25, 1057–1062 (2019).CAS 
    PubMed 
    Article 

    Google Scholar 
    Sánchez, C. et al. Contribution of citrate metabolism to the growth of Lactococcus lactis CRL264 at low pH. Appl. Environ. Microbiol. 74, 1136–1144 (2008).ADS 
    PubMed 
    Article 
    CAS 

    Google Scholar 
    Charyulu, E. M. & Gnanamani, A. Condition stabilization for Pseudomonas aeruginosa MTCC 5210 to yield high titers of extra cellular antimicrobial secondary metabolite using response surface methodology. Curr. Res. Bacteriol. 4, 197–213 (2010).Article 

    Google Scholar 
    Shen, Y. et al. Psychrobacillus lasiicapitis sp. nov., isolated from the head of an ant (Lasius fuliginosus). Int. J. Syst. Evol. Microbiol. 67, 4462–4467 (2017).CAS 
    PubMed 
    Article 

    Google Scholar 
    Rodríguez, M., Reina, J. C., Béjar, V. & Llamas, I. Psychrobacillus vulpis sp. nov., a new species isolated from faeces of a red fox in Spain. Int. J. Syst. Evol. Microbiol. 70, 882–888 (2020).PubMed 
    Article 
    CAS 

    Google Scholar 
    Pham, V. H. T., Jeong, S.-W. & Kim, J. Psychrobacillus soli sp. nov., capable of degrading oil, isolated from oil-contaminated soil. Int. J. Syst. Evol. Microbiol. 65, 3046–3052 (2015).CAS 
    PubMed 
    Article 

    Google Scholar 
    Kontro, M., Lignell, U., Hirvonen, M.-R. & Nevalainen, A. pH effects on 10 Streptomyces spp. growth and sporulation depend on nutrients. Lett. Appl. Microbiol. 41, 32–38 (2005).CAS 
    PubMed 
    Article 

    Google Scholar 
    Wodzinski, R. S., Umholtz, T. E., Rundle, J. R. & Beer, S. V. Mechanisms of inhibition of Erwinia amylovora by Erw. herbicola in vitro and in vivo. J. Appl. Bacteriol. 76, 22–29 (1994).Article 

    Google Scholar 
    Kuncharoen, N. et al. Achromobacter aloeverae sp. nov., isolated from the root of Aloe vera (L.) Burm. f. Int. J. Syst. Evol. Microbiol. 67, 37–41 (2017).CAS 
    PubMed 
    Article 

    Google Scholar 
    Aizawa, T. et al. Curtobacterium ammoniigenes sp. nov., an ammonia-producing bacterium isolated from plants inhabiting acidic swamps in actual acid sulfate soil areas of Vietnam. Int. J. Syst. Evol. Microbiol. 57, 1447–1452 (2007).CAS 
    PubMed 
    Article 

    Google Scholar 
    Kaira, G. S., Dhakar, K. & Pandey, A. A psychrotolerant strain of Serratia marcescens (MTCC 4822) produces laccase at wide temperature and pH range. AMB Express 5, 92 (2015).PubMed 
    Article 

    Google Scholar 
    Moon, J. & Kim, J. Isolation of Paenibacillus pinesoli sp. Nov. from forest soil in Gyeonggi-Do, Korea. J. Microbiol. 52, 273–277 (2014).CAS 
    PubMed 
    Article 

    Google Scholar 
    Heyrman, J. et al. Bacillus novalis sp. nov., Bacillus vireti sp. nov., Bacillus soli sp. nov., Bacillus bataviensis sp. nov. and Bacillus drentensis sp. nov., from the Drentse A grasslands. Int. J. Syst. Evol. Microbiol. 54, 47–57 (2004).CAS 
    PubMed 
    Article 

    Google Scholar 
    Hughes, K. L. & Sulaiman, I. The ecology of Rhodococcus equi and physicochemical influences on growth. Vet. Microbiol. 14, 241–250 (1987).CAS 
    PubMed 
    Article 

    Google Scholar 
    Schrempf, H. Recognition and degradation of chitin by streptomycetes. Antonie Van Leeuwenhoek 79, 285–289 (2001).CAS 
    PubMed 
    Article 

    Google Scholar 
    Seco, E. M., Cuesta, T., Fotso, S., Laatsch, H. & Malpartida, F. Two polyene amides produced by genetically modified Streptomyces diastaticus var. 108. Chem. Biol. 12, 535–543 (2005).CAS 
    PubMed 
    Article 

    Google Scholar 
    Kembel, S. W., Wu, M., Eisen, J. A. & Green, J. L. Incorporating 16S gene copy number information improves estimates of microbial diversity and abundance. PLoS Comput. Biol. 8, e1002743 (2012).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    León, M. et al. Antifungal activity of selected indigenous pseudomonas and bacillus from the soybean rhizosphere. Int. J. Microbiol. 2009, 572049 (2009).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Van Hai, N. & Fotedar, R. Comparison of the effects of the prebiotics (Bio-Mos® and β-1, 3-D-glucan) and the customised probiotics (Pseudomonas synxantha and P. aeruginosa) on the culture of juvenile western king prawns (Penaeus latisulcatus Kishinouye, 1896). Aquaculture 289, 310–316 (2009).Article 
    CAS 

    Google Scholar 
    Lauer, A., Simon, M. A., Banning, J. L., Lam, B. A. & Harris, R. N. Diversity of cutaneous bacteria with antifungal activity isolated from female four-toed salamanders. ISME J. 2, 145–157 (2008).CAS 
    PubMed 
    Article 

    Google Scholar 
    Ligon, J. M. et al. Natural products with antifungal activity fromPseudomonas biocontrol bacteria. Pest Manag. Sci. 56, 688–695 (2000).CAS 
    Article 

    Google Scholar 
    Scholz-Schroeder, B. K., Hutchison, M. L., Grgurina, I. & Gross, D. C. The contribution of syringopeptin and syringomycin to virulence of Pseudomonas syringae pv. syringae strain B301D on the basis of sypA and syrB1 biosynthesis mutant analysis. Mol. Plant Microb. Interact. 14, 336–348 (2001).CAS 
    Article 

    Google Scholar 
    Souza, J. T. & Raaijmakers, J. M. Polymorphisms within the prnD and pltC genes from pyrrolnitrin and pyoluteorin-producing Pseudomonas and Burkholderia spp. FEMS Microbiol. Ecol. 43, 21–34 (2003).PubMed 
    Article 

    Google Scholar 
    Mavrodi, D. V. et al. Functional analysis of genes for biosynthesis of pyocyanin and phenazine-1-carboxamide from Pseudomonas aeruginosa PAO1. J. Bacteriol. 183, 6454–6465 (2001).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Diby, P. et al. Mycolytic enzymes produced by Pseudomonas fluorescens and Trichoderma spp. against Phytophthora capsici, the foot rot pathogen of black pepper (Piper nigrum L.). Ann. Microbiol. 55, 129–133 (2005).CAS 

    Google Scholar 
    Vengust, M., Knapic, T. & Weese, J. S. The fecal bacterial microbiota of bats; Slovenia. PLoS ONE 13, e0196728 (2018).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Banskar, S., Mourya, D. T. & Shouche, Y. S. Bacterial diversity indicates dietary overlap among bats of different feeding habits. Microbiol. Res. 182, 99–108 (2016).PubMed 
    Article 

    Google Scholar 
    Wolkers-Rooijackers, J. C. M., Rebmann, K., Bosch, T. & Hazeleger, W. C. Fecal Bacterial Communities in Insectivorous Bats from the Netherlands and Their Role as a Possible Vector for Foodborne Diseases. Acta Chiropterol. 20, 475 (2019).Article 

    Google Scholar 
    Weller, T. J., Scott, S. A., Rodhouse, T. J., Ormsbee, P. C. & Zinck, J. M. Field identification of the cryptic vespertilionid bats, Myotis lucifugus and M. yumanensis. Acta Chiropt. 9, 133–147 (2007).Article 

    Google Scholar 
    Khankhet, J. et al. Clonal expansion of the Pseudogymnoascus destructans genotype in North America is accompanied by significant variation in phenotypic expression. PLoS ONE 9, e104625 (2014).Article 
    CAS 

    Google Scholar 
    McArthur, R. L., Ghosh, S. & Cheeptham, N. Improvement of protocols for the screening of biological control agents against white-nose syndrome. JEMI 2, 1–7 (2017).
    Google Scholar 
    Rajkumar, S. S. et al. Clonal genotype of Geomyces destructans among bats with white nose syndrome, New York, USA. Emerg. Infect. Dis. 17, 1273–1276 (2011).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Ren, P. et al. Clonal spread of Geomyces destructans among bats, Midwestern and Southern United States. Emerg. Infect. Dis. 18, 883–885 (2012).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Wilson, K. Genomc DNA extraction using the modified CTAB method. Curr. Protoc. Mol. Biol. 1, 1–2 (1997).
    Google Scholar 
    Edwards, U., Rogall, T., Blöcker, H., Emde, M. & Böttger, E. C. Isolation and direct complete nucleotide determination of entire genes: Characterization of a gene coding for 16S ribosomal RNA. Nucleic Acids Res. 17, 7843–7853 (1989).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Stackebrandt, E. & Liesack, W. Handbook of New Bacterial Systematics (Springer, 1993).
    Google Scholar 
    O’Leary, N. A. et al. Reference sequence (RefSeq) database at NCBI: Current status, taxonomic expansion, and functional annotation. Nucleic Acids Res. 44, D733–D745 (2016).PubMed 
    Article 
    CAS 

    Google Scholar 
    Camacho, C. et al. BLAST+: Architecture and applications. BMC Bioinformatics 10, 421 (2009).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Katoh, K. & Standley, D. M. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol. Biol. Evol. 30, 772–780 (2013).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Stamatakis, A. RAxML version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30, 1312–1313 (2014).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    R Core Team. R: A Language and Environment for Statistical Computing. (R Foundation for Statistical Computing, 2015).Venables, W. N. & Ripley, B. D. Modern applied statistics with S. Stat. Comput. https://doi.org/10.1007/978-0-387-21706-2 (2002).Article 
    MATH 

    Google Scholar 
    Bates, D., Mächler, M., Bolker, B. & Walker, S. Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1–48 (2015).Article 

    Google Scholar 
    Lenth, R. & Lenth, M. R. Package ‘lsmeans’. Am. Stat. 34, 216–221 (2018).
    Google Scholar 
    Kassambara, A. ggpubr:‘ggplot2’ based publication ready plots. R package version 0.1. 7 (2018). More

  • in

    Convergent evolution of a labile nutritional symbiosis in ants

    Genome characteristics of ancient obligate symbiontsWe first tested the hypothesis that each of the ant lineages sequenced in our study (Cardiocondyla, Formica, and Plagiolepis) hosts its own ancient strictly vertically transmitted symbiont that have co-speciated with its host, which has been shown previously in the Camponotus- Blochmannia symbiosis [17]. To address this aim, we compared the genomes of symbionts from 13 species of ants, 8 from our study combined with 5 previously published genomes, representing four independently evolved symbioses. This includes symbionts from three Formica, two Plagiolepis, and an additional three Cardiocondyla species that we sequenced, in addition to four previously published genomes from Blochmannia, the obligate symbiont of Camponotus ants, and the one pre-existing Westeberhardia genome from Cardiocondyla obscurior [8, 18,19,20,21].We found the gene order of single copy orthologs in symbionts is highly conserved in ant species belonging to the same genus (Fig. 1). This type of structural stability of genomes is typically found in symbionts that have been strictly vertically transmitted within a matriline [22] and has been documented in the obligate symbionts of whiteflies, psyllids, cockroaches, and aphids [23,24,25,26]. In contrast, genome structure differed substantially between symbionts from different ant genera (Fig. 1, Fig. S1). We also find that the host and symbiont phylogenies are in general concordance in Cardiocondyla (TreeMap: p = 0.00100 CI95% = [0.00000, 0.00424]), and in Formica the topologies suggest co-segregation, although there were too few nodes to confirm this statistically (Fig. S2). Together, this strongly suggests the symbioses in all four ant lineages are independently acquired ancient associations that have co-speciated with their hosts.Fig. 1: Structural stability of ant symbiont genomes.A Ant lineages known to host bacteriocyte-associated symbionts (red font) and lineages not known to (black font), based on [91]. Outgroup (grey font) not examined in this study. B Visualisation of symbiont genomes showing conservation of gene order in the symbionts of ant species that belong to the same genus. Blocks show the locations of single copy orthologs in the symbiont genome, lines connect shared single copy orthologs between genomes. All genomes and annotations were generated in this study except the Blochmannia symbionts and the Westeberhardia strain from C. obscurior [8, 18,19,20,21]. *Evidence of symbionts were detected in embryos of Anoplolepis [91] but it is unclear if they are localised in bacteriocytes in larvae and adults.Full size imageIn addition, our phylogenetic analysis reveals that all four symbiont lineages originate from a single clade, the Sodalis-allied bacteria (Fig. 2). This demonstrates that ant lineages that host bacteriocytes-associated symbionts have convergently acquired related bacteria, which differs from previous findings based on limited taxa and genes [27]. All of the symbionts have evidence of advanced genome reduction, which is characterized by reduced genome size, GC content, and number of coding sequences, similar to other ancient obligate symbionts of insects [4]. The three strains of Westeberhardia we analysed have extremely small (0.45–0.53 Mb) GC depleted genomes (22–26%) that are similar to the figures reported for the strain in Cardiocondyla obscurior [8]; confirming that they have some of the smallest genomes of any known gammaproteobacterial endosymbiont (Fig. 2). By comparison, the symbionts in Formica and Plagiolepis have genomes around twice the size (1.37–1.38 Mb) and GC content (~41%) of Westeberhardia (Fig. 2) raising the possibility that they are in an earlier stage of genome reduction than both Westeberhardia and Blochmannia. The Formica and Plagiolepis symbionts have a similar size, GC range, and number of coding sequences as known obligate symbionts such as Candidatus Doolittlea endobia [28], and several Serratia symbiotica lineages that are co-obligate symbionts in aphids [29].Fig. 2: Phylogenetic origins of the bacteriocyte-associated symbionts of ants.A pruned phylogeny of gammaproteobacterial endosymbionts based on Fig. S8. The phylogeny is based on a dayhoff6 recoded amino acid alignment of 72 genes analysed using phylobayes. Bar plots represent the size (in Mbp) and GC content of symbiont genomes. Bars are colour coded to represent hypothesised relationships between symbionts and hosts. Species names highlighted in red in the phylogeny indicate the four bacteriocyte-associated symbionts of ants. Genomes sequenced and assembled for this paper are referenced as ‘novel symbiont’ lineages. Full phylogenies with node support and branch lengths are available as Fig. S8 and Fig. S9, respectively.Full size imageBacteriocyte-associated endosymbiontsUsing fluorescent in situ hybridisation, we determine whether the Sodalis-allied symbionts we sequenced are localised in bacteriocytes to confirm they are the associations first observed by Lillienstern and Jungen in the early 1900’s [10, 11].Consistent with Lilienstern’s findings [11], we found the Sodalis symbiont in Formica ants is distributed in bacteriocytes surrounding the midgut in adult queens (Fig. 3A). The symbionts are also found in eggs and ovaries of adult queens, indicating they are vertically transmitted from queens to offspring (Fig. 3B–C). Sectioning of F. cinerea larvae shows the bacteriocytes to be arranged in a single layer of cells surrounding the midgut, as well as in clusters of bacteriocytes closely situated to the midgut (Fig. 3D–D’). In adult Plagiolepis queens, the symbiont was not present in bacteriocytes around the midgut, suggesting the symbiont may play a more substantive role in larval development or pupation and then migrates to the ovaries prior to or during metamorphosis. Apart from that, the localisation of the symbiont in Plagiolepis was the same as in Formica – symbionts in larval midgut bacteriocytes, ovaries and eggs (Fig. S3) – supporting Jungen’s cytological findings [10]. Bacteriocytes are also found surrounding the midgut in Camponotus and Cardiocondyla ants [8, 30, 31] indicating the symbionts are localised in a similar manner in all four ant lineages.Fig. 3: Anatomical localisation of symbiont in Formica ants.Fluorescent in situ hybridisation (FISH) generated images showing the localisation of symbionts in Formica ants. A–C Whole mount FISH of Formica fusca: queen gut (A, crop and proventriculus on the right, midgut in the middle, hindgut and Malpighian tubules on the left), ovaries (B) and egg (C). DAPI staining of host tissue in blue, symbiont stained in red. D–D’. FISH on transverse cytological sections of Formica cinerea larva midgut. DAPI staining only, showing host nuclei of bacteriocytes in a single layer surrounding the midgut (D), and a magnified region highlighting symbionts in red localised within bacteriocytes and in a bacteriome (D’). A FISH image of the symbiont-free midgut of a Formica lemani queen is available as Fig. S11.Full size imageConservation of metabolic functions in ant endosymbiontsDespite on-going genome reduction, obligate symbionts of insects typically retain gene networks required for maintaining the symbiosis with their host, such as pathways for synthesising essential nutrients. This has resulted in the symbionts of sap- and blood-feeding insects converging on genomes that have retained the same sets of metabolic pathways – to synthesise essential nutrients missing in their hosts’ diets [32, 33]. Here we compare the metabolic pathways retained in the reduced genomes of the four bacteriocyte-associated symbionts of ants to test the hypothesis that have been acquired to perform similar functions. For this, we assess whether they have consistently retained metabolic pathways to synthesise the same key nutrients. Two major patterns stand out.The first major pattern we find is that the four ant symbionts have all retained the shikimate pathway, which produces chorismate, along with most of the steps necessary to produce tyrosine from this precursor (Tables 1 and S2). Both the symbiont of Formica and Westeberhardia each lack one of the genes required to produce tyrosine. However, in Westeberhardia it is believed the host encodes the missing gene, supplying the enzyme to fulfil the final step of the pathway [8]. Intriguingly, we find that this gene is also present in the Formica ant genomes (Fig. S4). In addition, all symbionts except Westeberhardia can produce phenylalanine which is a precursor that can be converted to tyrosine by their hosts [5, 34, 35]. Tyrosine is important for insect development as it is used to produce L-DOPA, which is a key component of insect cuticles [5]. Tyrosine is also a precursor for melanin synthesis, which is important in protection against pathogens, and plays a fundamental role in neurotransmitters and hormone production [36, 37]. In several species of ants, weevils, and other beetles, symbionts are believed to provision hosts with tyrosine, and it has been shown experimentally in several of these species that removal or inhibition of their symbionts causes cuticle development to suffer [38,39,40,41,42,43,44,45,46,47]. A thicker cuticle has been shown to help symbiont-carrying grain beetles resist desiccation [43], and defend against natural enemies [48]. However, female reproduction is delayed at higher humidity, suggesting a metabolic cost to carrying their Bacteroidetes symbiont. Tyrosine provisioning is also the likely function of Westeberhardia in Cardiocondyla ants, as this is one of the few nutrient pathways retained in this symbiont. Our analysis confirms the shikimate pathway, and the symbiont portions of the tyrosine pathway, have been retained in Westeberhardia from three phylogenetically diverse Cardiocondyla lineages, providing additional support for this hypothesis. In addition to tyrosine, most of the symbionts have retained the capacity to produce vitamin B9 (tetrahydrofolate) and all can perform the single step conversions necessary to produce alanine and glycine. However, our gene enrichment analysis indicates that tyrosine, and the associated chorismate biosynthetic process, are the only enriched vitamin or amino acid pathways that are shared by all of the symbiont genomes (Table S1). This suggests that provisioning of tyrosine by symbionts, or tyrosine precursors, is of general importance across all bacteriocyte-associated symbioses of ants.Table 1 Comparison of the retention and losses of metabolic pathways for key nutrients across ant symbionts.Full size tableThe second major pattern emerging from our comparative analysis is that there are clear differences in the pathways lost or retained across symbionts (Tables 1 and S2). This is most evident when comparing Blochmannia with Westeberhardia, the latter of which has lost the capacity to synthesise most essential nutrients. The symbionts of Formica or Plagiolepis, in contrast, have retained the capacity to synthesise many of the same amino acids and B vitamins as Blochmannia, suggesting they may perform similar functions for their hosts. However, Blochmannia has retained more biosynthetic pathways, particularly those involved in the synthesis of essential amino acids. Previous experimental studies have confirmed that Blochmannia provisions hosts with essential amino acids [1]. The absence of several core essential amino acids in the Formica and Plagiolepis symbionts may reflect differences in the dietary ecology of the different ant genera. The retention of the full complement of essential amino acids biosynthetic pathways in the highly reduced genome of Blochmannia does however indicate it plays a more substantive nutrient-provisioning role for its hosts than the other ant symbionts we investigated.Previous work on the extracellular gut symbionts of several arboreal ant lineages identified nitrogen recycling via the urease operon as a function that may be of key importance for ant symbioses [1, 2, 49, 50]. However, we do not find any evidence that the symbionts of Formica, Plagiolepis, or Cardiocondyla play a role in nitrogen recycling via the urease operon (Table 1). This suggests that nitrogen recycling may play an important role for more strictly herbivorous ants, such as Cephalotes. Our results indicate tyrosine supplementation by symbionts may be universally required for essential physiological process across a broader range of ant lineages.The origins and losses of symbioses in Formica and Cardiocondyla
    We investigated the presence of the symbiont in phylogenetically diverse Formica and Cardiocondyla species to identify the evolutionary origins and losses of the symbiosis. Although the symbiont in Plagiolepis was present in P. pygmaea and two unknown Plagiolepis species we investigated, we did not have sufficient phylogenetic sampling to assess the origins of the symbiosis.In Formica, we find the symbiont is restricted to a single clade in the paraphyletic Serviformica group (Fig. 4A). The species in this clade are socially polymorphic, forming both multi-queen and single-queen colonies [51]. Based on a previously dated phylogeny of Formica ants, we estimate the symbiosis originated approximately 12–22 million years ago [52]. In Cardiocondyla, the symbiosis is widespread throughout the genus. The prevalence of the symbiont in Cardiocondyla, in combination with its highly reduced genome, suggests it is a very old association that likely dates back to the origins of the ant genus some 50–75 million years ago [53]. The symbiont was also absent in two clades, the argentea and palearctic groups (Fig. 4B). This may represent true evolutionary losses in these clades. It may be that these losses are linked to a notable change in social structure in these two Cardiocondyla clades, having gone from the ancestral state of maintaining multi-queen colonies to single-queen colonies [54], however it is not clear how this could impact the symbiosis.Fig. 4: Phylogenetic distribution of symbionts in queens of Formica and Cardiocondyla ants.Pie charts represent the proportion of Formica (A) and Cardiocondyla (B) queens sampled that carried the symbiont (red) and those that did not (grey). Numbers represent the number of queens positive for the symbiont over the total number of queens sampled (intracolony infection frequencies in Table S5). See the supplementary material for the statistical testing of differences in prevalence within Serviformica Clade 1. The Formica phylogeny is based on [81] and the Cardiocondyla phylogeny is based on [83], with major clades highlighted. Dashed lines indicate species added to the original source phylogeny based on additional published phylogenies (specified in the Taxonomic Analysis section of the methods). Starred names are provisional names of a recognised morphospecies to be described by B. Seifert.Full size imageEvidence of variation in colony-level dependence on symbiontsObservations from individual studies on F. cinerea and F. lemani [10, 11], as well as Cardiocondyla obscurior [8], reported rare cases of ant queens not harbouring their symbionts in nature. This called into question the degree to which these insects depend on symbionts for nutrients, and whether the symbiosis may be breaking down in certain host lineages. However, given the limited number of species and populations studied, it is unclear how often colonies are maintained with uninfected queens, and whether this varies across species, suggesting species may differ in their dependence on their symbiont. To answer this question, we assessed the presence of the symbionts in 838 samples from 147 colonies of phylogenetically diverse Formica and Cardiocondyla species collected across 8 countries.Our investigation reveals the natural occurrence of uninfected queens is a widespread phenomenon in many Formica and Cardiocondyla species (Fig. 4). We confirmed the absence of symbionts in queens, and that they have not been replaced with another bacterial or fungal symbiont, using multiple approaches including diagnostic PCR, metagenomic and deep-coverage amplicon sequencing (Tables S3,  S4, Figs. S5, S6). Wolbachia was high in relative abundance, especially in Formica ants, but was not sufficiently present across samples to be a feasible replacement. There was also clear evidence of variation across host species. In Formica, queens and workers of F. fusca always carried the symbiont, whereas queens and workers of F. lemani, F. cinerea, and F. selysi showed varying degrees of individuals not carrying the symbionts (Fig. 4A, Table S5). A similar pattern can be seen in Cardiocondyla, where queens of several species, such as C. obscurior, always carry the symbiont, compared to lower incidences in other species (Fig. 4B). Klein et al [8] identified a single C. obscurior colony with uninfected queens in Japan. However, queens of this species nearly always carry the symbiont in nature.The degradation and eventual loss of symbionts from bacteriocytes has been reported in males, and in sterile castes of aphids and ants [55], which do not transmit symbionts to offspring. In reproductive females, bacteriocytes may degrade as a female ages; however, symbionts are typically retained at high bacterial loads in the ovaries, as this is required to maintain the symbionts within the germline [31]. All of the symbiotic ant species we investigated maintain multi-queen colonies, and the vast majority had at least one queen, often more, within a colony that carried the symbiont (Table S5). We hypothesize that species that maintain colonies with uninfected queens may be able to retain sufficient colony-level fitness with only a fraction of queens harbouring the symbiont and receiving its nutritive benefits.Dependence on symbionts in a socioecological contextThe retention of symbionts in queens and workers of some species, but not others, suggests species either differ in their dependence on symbiont-derived nutrients, or that symbionts have lost the capacity to make nutrients in certain host lineages. Our analysis of symbiont genomes did not reveal any structural differences, such as the disruption of metabolic pathways, which could explain differences in symbiont retention between host species (Table S2). This suggests differences in the retention of symbionts may reflect differences in host ecologies.In ants, which occupy a wide range of feeding niches, reliance on symbiont-derived nutrients will largely depend on lineage-specific feeding ecologies. For example, several species of arboreal Camponotus ants have been shown to be predominantly herbivorous [56]. Blochmannia, in turn, has retained the capacity to synthesise key nutrients lacking in their plant-based diets, such as essential amino acids [1]. Blochmannia is also always present in queens and workers [31], which is a testament to the importance of these nutrients for the survival of its primarily herbivorous host [13]. In contrast, Formica and Cardiocondyla species are thought to have a more varied diet [14]. Diet flexibility and altered foraging efforts may therefore reduce their reliance on a limited number of symbiont-derived nutrients allowing colonies of some species to persist with uninfected queens in certain contexts. Silvanid beetles and grain weevils, for example, can survive in the absence of their tyrosine-provisioning symbionts [38, 57, 58] when provided nutritionally balanced diets, in the laboratory [57] or in cereal grain elevators [59, 60]. Similarly, studies on Cardiocondyla and Camponotus ants have shown they can maintain sufficient colony health in the absence of their symbionts, if provided a balanced diet [31, 61]. It would be interesting to know whether species of Formica and Cardiocondyla that always carry the symbiont in nature, such as F. fusca and C. obscurior, have more restricted diets with less access to nutrients such as tyrosine, as this may explain their dependence on their symbiont for nutrients and tendency to harbour them in queens.Although it is unusual for bacteriocyte-associated symbionts to be absent in reproductive females, the fact that it is simultaneously occurring in phylogenetically diverse species from many locations suggests the symbiosis may have persisted in this manner over evolutionary time. Perhaps through diet flexibility colonies can be maintained with uninfected queens in some contexts, however we expect them to be disadvantaged in other ecological scenarios. Fluctuating environmental conditions may therefore eventually purge asymbiotic queens from lineages, allowing the symbiosis to be retained over longer periods of evolutionary time. The multiple-queen colony lifestyle in all symbiotic Formica and Cardiocondyla species we investigated may also provide an additional social buffer that limits the costs to individual queens being asymbiotic. Workers will still nourish larvae and queens without symbionts and colony fitness may be maintained through the reproductive output of nestmate queens that carry the symbiont. There may also be an adaptive explanation for the losses if, for example, metabolic costs to maintain the symbiosis trade off in a context dependent manner [44, 62, 63]. Under this scenario, maintaining a mix of infected and uninfected queens may benefit a colony by allowing for optimal reproduction under a broader range of environmental scenarios.Our data suggest that symbiotic relationships can evolve to solve common problems but also rapidly break down if the symbiosis is no longer required, or potentially when costs are too high [44]. We have identified tyrosine provisioning as a possible unifying function across bacteriocyte-associated symbionts of ants. But we have also shown species can vary in how much they depend on symbionts for nutrients. Our results demonstrate that ants have a unique labile symbiotic system, allowing us to better understand the evolutionary forces that influence the persistence and breakdown of long-term endosymbiotic mutualisms.
    Candidatus Liliensternia hugann and Candidatus Jungenella plagiolepisWe propose the names Candidatus Liliensternia hugann for the Sodalis-allied symbiont found in Formica. The genus name is in honour of Margarete Lilienstern who first identified the symbiont [11]. The species name is derived from the combined first names of the first authors parents. Similarly, we propose the name of Candidatus Jungenella plagiolepis for the Plagiolepis-bound symbiont. The genus name is in honour of Hans Jungen who originally discovered the symbiont [10], and the species name is derived from Plagiolepis, the genus in which the symbiont can be found. More