More stories

  • in

    Social support correlates with glucocorticoid concentrations in wild African elephant orphans

    Wu, A. Social buffering of stress – Physiological and ethological perspectives. Appl. Anim. Behav. Sci. 239, 105325 (2021).
    Google Scholar 
    Hennessy, M. B., Kaiser, S. & Sachser, N. Social buffering of the stress response: diversity, mechanisms, and functions. Front. Neuroendocrinol. 30, 470–482 (2009).CAS 
    PubMed 

    Google Scholar 
    Young, C., Majolo, B., Heistermann, M., Schülke, O. & Ostner, J. Responses to social and environmental stress are attenuated by strong male bonds in wild macaques. Proc. Natl Acad. Sci. USA 111, 18195–18200 (2014).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Stanton, M. E., Patterson, J. M. & Levine, S. Social influences on conditioned cortisol secretion in the squirrel monkey. Psychoneuroendocrinology 10, 125–134 (1985).CAS 
    PubMed 

    Google Scholar 
    Caldji, C., Diorio, J. & Meaney, M. J. Variations in maternal care in infancy regulate the development of stress reactivity. Biol. Psychiatry 48, 1164–1174 (2000).CAS 
    PubMed 

    Google Scholar 
    Novak, M. A., Hamel, A. F., Kelly, B. J., Dettmer, A. M. & Meyer, J. S. Stress, the HPA axis, and nonhuman primate well-being: a review. Appl. Anim. Behav. Sci. 143, 135–149 (2013).PubMed 
    PubMed Central 

    Google Scholar 
    Sapolsky, R. M., Romero, L. M. & Munck, A. U. How do glucocorticoids influence stress responses? Integrating permissive, suppressive, stimulatory, and preparative actions. Endocr. Rev. 21, 55–89 (2000).CAS 
    PubMed 

    Google Scholar 
    Liu, D. et al. Maternal Care, hippocampal glucocorticoid receptors, and hypothalamic-pituitary-adrenal responses to stress. Sci. Ment. Heal. Stress Brain 9, 75–78 (1997).
    Google Scholar 
    Gjerstad, J. K., Lightman, S. L. & Spiga, F. Role of glucocorticoid negative feedback in the regulation of HPA axis pulsatility. Stress 21, 403–416 (2018).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Spiga, F., Walker, J. J., Terry, J. R. & Lightman, S. L. HPA axis-rhythms. Compr. Physiol. 4, 1273–1298 (2014).PubMed 

    Google Scholar 
    Sapolsky, R. M. Why Zebras Don’t Get Ulcers (Henry Holt and Company, LLC, 2004).Campos, F. A. et al. Glucocorticoid exposure predicts survival in female baboons. Sci. Adv. 7, 1–10 (2021).
    Google Scholar 
    Banerjee, S. B., Arterbery, A. S., Fergus, D. J. & Adkins-Regan, E. Deprivation of maternal care has long-lasting consequences for the hypothalamic-pituitary-adrenal axis of zebra finches. Proc. R. Soc. B Biol. Sci. 279, 759–766 (2012).
    Google Scholar 
    Hennessy, M. B., Nigh, C. K., Sims, M. L. & Long, S. J. Plasma cortisol and vocalization responses of postweaning age guinea pigs to maternal and sibling separation: evidence for filial attachment after weaning. Dev. Psychobiol. 28, 103–115 (1995).CAS 
    PubMed 

    Google Scholar 
    Hennessy, M. B., O’Leary, S. K., Hawke, J. L. & Wilson, S. E. Social influences on cortisol and behavioral responses of preweaning, periadolescent, and adult guinea pigs. Physiol. Behav. 76, 305–314 (2002).CAS 
    PubMed 

    Google Scholar 
    Wiener, S. G., Johnson, D. F. & Levine, S. Influence of postnatal rearing conditions on the response of squirrel monkey infants to brief perturbations in mother-infant relationships. Physiol. Behav. 39, 21–26 (1987).CAS 
    PubMed 

    Google Scholar 
    Girard-Buttoz, C. et al. Early maternal loss leads to short-but not long-term effects on diurnal cortisol slopes in wild chimpanzees. Elife 10, e64134 (2021).Rosenbaum, S. et al. Social bonds do not mediate the relationship between early adversity and adult glucocorticoids in wild baboons. Proc. Natl Acad. Sci. USA 117, 20052–20062 (2020).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Moss, C. Elephant Memories: Thirteen Years in the Life of an Elephant Family (Univ. Chicago Press, 1988).Douglas-Hamilton, I., Bhalla, S., Wittemyer, G. & Vollrath, F. Behavioural reactions of elephants towards a dying and deceased matriarch. Appl. Anim. Behav. Sci. 100, 87–102 (2006).
    Google Scholar 
    Shoshani, J., Kupsky, W. J. & Marchant, G. H. Elephant brain. Part I: gross morphology, functions, comparative anatomy, and evolution. Brain Res. Bull. 70, 124–157 (2006).PubMed 

    Google Scholar 
    Goldenberg, S. Z. & Wittemyer, G. Orphaned female elephant social bonds reflect lack of access to mature adults. Sci. Rep. 7, 14408 (2017).PubMed 
    PubMed Central 

    Google Scholar 
    Goldenberg, S. Z. & Wittemyer, G. Orphaning and natal group dispersal are associated with social costs in female elephants. Anim. Behav. 143, 1–8 (2018).
    Google Scholar 
    Lee, P. C. Allomothering among African elephants. Anim. Behav. 35, 278–291 (1987).
    Google Scholar 
    Parker, J. M. et al. Poaching of African elephants indirectly decreases population growth through lowered orphan survival. Curr. Biol. 31, 4156–4162.e5 (2021).Wittemyer, G. et al. Where sociality and relatedness diverge: the genetic basis for hierarchical social organization in African elephants. Proc. R. Soc. B Biol. Sci. 276, 3513–3521 (2009).
    Google Scholar 
    Goldenberg, S. Z., Douglas-Hamilton, I. & Wittemyer, G. Vertical transmission of social roles drives resilience to poaching in elephant metworks. Curr. Biol. 26, 75–79 (2016).CAS 
    PubMed 

    Google Scholar 
    Gobush, K. S., Mutayoba, B. M. & Wasser, S. K. Long-term impacts of poaching on relatedness, stress physiology, and reproductive output of adult female African elephants. Conserv. Biol. 22, 1590–1599 (2008).CAS 
    PubMed 

    Google Scholar 
    Gobush, K. S. et al. Loxodonta africana (African Savanna Elephant). Loxodonta africana: the IUCN red list of threatened species 2021 e.T181008073A181022663 https://doi.org/10.2305/IUCN.UK.2021-1.RLTS.T181008073A181022663.en (2021).Wittemyer, G. et al. Illegal killing for ivory drives global decline in African elephants. Proc. Natl Acad. Sci. USA 111, 13117–13121 (2014).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Wittemyer, G., Daballen, D. & Douglas-Hamilton, I. Comparative Demography of an At-Risk African Elephant Population. PLoS ONE 8, e53726 (2013).McCormick, S. D. & Romero, L. M. Conservation endocrinology. Bioscience 67, 429–442 (2017).
    Google Scholar 
    Wittemyer, G. The elephant population of Samburu and Buffalo Springs National Reserves, Kenya. Afr. J. Ecol. 39, 357–369 (2001).
    Google Scholar 
    Cockrem, J. F. Individual variation in glucocorticoid stress responses in animals. Gen. Comp. Endocrinol. 181, 45–58 (2013).CAS 
    PubMed 

    Google Scholar 
    Taff, C. C., Schoenle, L. A. & Vitousek, M. N. The repeatability of glucocorticoids: a review and meta-analysis. Gen. Comp. Endocrinol. 260, 136–145 (2018).CAS 
    PubMed 

    Google Scholar 
    Hooten, M. B. & Hobbs, N. T. A guide to Bayesian model selection for ecologists. Ecol. Monogr. 85, 3–28 (2015).
    Google Scholar 
    Wittemyer, G. & Getz, W. M. Hierarchical dominance structure and social organization in African elephants, Loxodonta africana. Anim. Behav. 73, 671–681 (2007).
    Google Scholar 
    Heim, C., Ehlert, U. & Hellhammer, D. H. The potential role of hypocortisolism in the pathophysiology of stress-related bodily disorders. Psychoneuroendocrinology 25, 1–35 (2000).CAS 
    PubMed 

    Google Scholar 
    Dickens, M. J. & Romero, L. M. A consensus endocrine profile for chronically stressed wild animals does not exist. Gen. Comp. Endocrinol. 191, 177–189 (2013).CAS 
    PubMed 

    Google Scholar 
    Ma, D., Serbin, L. A. & Stack, D. M. How children’s anxiety symptoms impact the functioning of the hypothalamus–pituitary–adrenal axis over time: a cross-lagged panel approach using hierarchical linear modeling. Dev. Psychopathol. 31, 1–15 (2018).
    Google Scholar 
    Blas, J., Bortolotti, G. R., Tella, J. L., Baos, R. & Marchant, T. A. Stress response during development predicts fitness in a wild, long lived vertebrate. Proc. Natl Acad. Sci. USA 104, 8880–8884 (2007).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Boonstra, R. Reality as the leading cause of stress: rethinking the impact of chronic stress in nature. Funct. Ecol. 27, 11–23 (2013).
    Google Scholar 
    Gunnar, M. R. & Vazquez, D. M. Low cortisol and a flattening of expected daytime rhythm: Potential indices of risk in human development. Dev. Psychopathol. 13, 515–538 (2001).CAS 
    PubMed 

    Google Scholar 
    Perry, R. E. et al. Corticosterone administration targeting a hypo-reactive HPA axis rescues a socially-avoidant phenotype in scarcity-adversity reared rats. Dev. Cogn. Neurosci. 40, 100716 (2019).Fries, E., Hesse, J., Hellhammer, J. & Hellhammer, D. H. A new view on hypocortisolism. Psychoneuroendocrinology 30, 1010–1016 (2005).CAS 
    PubMed 

    Google Scholar 
    Dorsey, C., Dennis, P., Guagnano, G., Wood, T. & Brown, J. L. Decreased baseline fecal glucocorticoid concentrations associated with skin and oral lesions in black rhinoceros (Diceros bicornis). J. Zoo. Wildl. Med. 41, 616–625 (2010).PubMed 

    Google Scholar 
    Pawluski, J. et al. Low plasma cortisol and fecal cortisol metabolite measures as indicators of compromised welfare in domestic horses (Equus caballus). PLoS ONE 12, 1–18 (2017).
    Google Scholar 
    Feng, X. et al. Maternal separation produces lasting changes in cortisol and behavior in rhesus monkeys. Proc. Natl Acad. Sci. USA 108, 14312–14317 (2011).CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    González Ramírez, C. et al. The NR3C1 gene expression is a potential surrogate biomarker for risk and diagnosis of posttraumatic stress disorder. Psychiatry Res. 284, 112797 (2020).PubMed 

    Google Scholar 
    Cluver, L., Fincham, D. S. & Seedat, S. Posttraumatic stress in AIDS-orphaned children exposed to high levels of trauma: the protective role of perceived social support. J. Trauma. Stress 22, 106–112 (2009).PubMed 

    Google Scholar 
    Bastille-Rousseau, G. et al. Landscape-scale habitat response of African elephants shows strong selection for foraging opportunities in a human dominated ecosystem. Ecography 43, 149–160 (2020).
    Google Scholar 
    Foley, C. A. H., Papageorge, S. & Wasser, S. K. Noninvasive stress and reproductive measures of social and ecological pressures in free-ranging African elephants. Conserv. Biol. 15, 1134–1142 (2001).
    Google Scholar 
    Wittemyer, G., Getz, W. M., Vollrath, F. & Douglas-Hamilton, I. Social dominance, seasonal movements, and spatial segregation in African elephants: a contribution to conservation behavior. Behav. Ecol. Sociobiol. 61, 1919–1931 (2007).
    Google Scholar 
    Wittemyer, G., Daballen, D. & Douglas‐Hamilton, I. Differential influence of human impacts on age‐specific demography underpins trends in an African elephant population. Ecosphere 12, e03720 (2021).Brown, J. L. et al. Individual and environmental risk factors associated with fecal glucocorticoid metabolite concentrations in zoo-housed Asian and African elephants. PLoS ONE 14, 1–18 (2019).
    Google Scholar 
    Goldenberg, S. Z. et al. Increasing conservation translocation success by building social functionality in released populations. Glob. Ecol. Conserv. 18, e00604 (2019).Dantzer, B., Fletcher, Q. E., Boonstra, R. & Sheriff, M. J. Measures of physiological stress: a transparent or opaque window into the status, management and conservation of species? Conserv. Physiol. 2, 1–18 (2014).
    Google Scholar 
    Kaisin, O., Fuzessy, L., Poncin, P., Brotcorne, F. & Culot, L. A meta-analysis of anthropogenic impacts on physiological stress in wild primates. Conserv. Biol. 0, 1–14 (2020).CAS 

    Google Scholar 
    Ganswindt, A., Rasmussen, H. B., Heistermann, M. & Hodges, J. K. The sexually active states of free-ranging male African elephants (Loxodonta africana): defining musth and non-musth using endocrinology, physical signals, and behavior. Horm. Behav. 47, 83–91 (2005).CAS 
    PubMed 

    Google Scholar 
    Santymire, R. M. et al. Using ACTH challenges to validate techniques for adrenocortical activity analysis in various African wildlife species. Int. J. Anim. Vet. Adv. 4, 99–108 (2012).CAS 

    Google Scholar 
    Watson, R. et al. Development of a versatile enzyme immunoassay for non-invasive assessment of glucocorticoid metabolites in a diversity of taxonomic species. Gen. Comp. Endocrinol. 186, 16–24 (2013).CAS 
    PubMed 

    Google Scholar 
    Oduor, S. et al. Differing physiological and behavioral responses to anthropogenic factors between resident and non-resident African elephants at Mpala Ranch, Laikipia County, Kenya. PeerJ 8, e10010 (2020).Brown, J. L., Kersey, D. C., Freeman, E. W. & Wagener, T. Assessment of diurnal urinary cortisol excretion in Asian and African elephants using different endocrine methods. Zoo. Biol. 29, 274–283 (2010).PubMed 

    Google Scholar 
    Justice, C. O. et al. The moderate resolution imaging spectroradiometer (MODIS): land remote sensing for global change research. IEEE Trans. Geosci. Remote Sens. 36, 1228–1249 (1998).
    Google Scholar 
    Lafferty, D. J. R., Zimova, M., Clontz, L., Hackländer, K. & Mills, L. S. Noninvasive measures of physiological stress are confounded by exposure. Sci. Rep. 9, 1–6 (2019).
    Google Scholar 
    O’Dwyer, K., Dargent, F., Forbes, M. R. & Koprivnikar, J. Parasite infection leads to widespread glucocorticoid hormone increases in vertebrate hosts: a meta-analysis. J. Anim. Ecol. 89, 519–529 (2020).PubMed 

    Google Scholar 
    Parker, J. M., Goldenberg, S. Z., Letitiya, D. & Wittemyer, G. Strongylid infection varies with age, sex, movement and social factors in wild African elephants. Parasitology 147, 348–359 (2020).PubMed 

    Google Scholar 
    Gibbons, L., Jacobs, D. E., Fox, M. T. & Hansen, J. The RVC/FAO guide to veterinary diagnostic parasitology. McMaster egg-counting technique. http://www.rvc.ac.uk/review/Parasitology/EggCount/Purpose.htm (2004)R Core Team. A language and environment for statistical computing. https://www.r-project.org/. (2020).Rstudio Team. RStudio: integrated development for R. http://www.rstudio.com/ (2020).Plummer, M. rjags: Bayesian graphical models using MCMC. https://cran.r-project.org/package=rjags (2019).Brooks, S. P. & Gelman, A. General methods for monitoring convergence of iterative simulations. J. Comput. Graph. Stat. 7, 434–455 (1998).
    Google Scholar 
    Gelman, A. & Rubin, D. B. Inference from iterative simulation using multiple sequences. Stat. Sci. 7, 457–511 (1992).
    Google Scholar 
    Wickham, H. ggplot2: elegant graphics for data analysis. https://ggplot2.tidyverse.org (2016).Youngflesh, C. MCMCvis: tools to visualize, manipulate, and summarize MCMC output. J. Open Source Softw. 3, 640 (2018).
    Google Scholar 
    Parker, J. M. The Physiological Condition of Orphaned African Elephants (Loxodonta africana). Doctoral dissertation, Colorado State University. (2021). More

  • in

    Plant rarity in fire-prone dry sclerophyll communities

    Mouillot, D. et al. Rare species support vulnerable functions in high-diversity ecosystems. PLoS Biol. 11, e1001569 (2013).CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Leitão, R. P. et al. Rare species contribute disproportionately to the functional structure of species assemblages. Proc. R Soc. B 283, 20160084 (2016).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Enquist, B. J. et al. The commonness of rarity: Global and future distribution of rarity across land plants. Sci. Adv. 5, eaaz0414 (2019).Bevill, R. L. & Louda, S. M. Comparisons of related rare and common species in the study of plant rarity. Conserv. Biol. 13, 493–498 (1999).Article 

    Google Scholar 
    Murray, B. R., Thrall, P. H., Gill, A. M. & Nicotra, A. B. How plant life-history and ecological traits relate to species rarity and commonness at varying spatial scales. Austral Ecol. 27, 291–310 (2002).Article 

    Google Scholar 
    Gaston, K. J. Common ecology. Bioscience 61, 354–362 (2011).Article 

    Google Scholar 
    Kraft, N. J. et al. Community assembly, coexistence and the environmental filtering metaphor. Funct. Ecol. 29, 592–599 (2015).Article 

    Google Scholar 
    Gaston, K. J. What is rarity? in Rarity 1–21 (Springer, 1994).Rabinowitz, D. Seven forms of rarity. in The biological aspects of rare plant conservation (ed. Synge, H.) 205–217 (John Wiley and Sons: Chichester, UK, 1981).Sykes, L., Santini, L., Etard, A. & Newbold, T. Effects of rarity form on species’ responses to land use. Conserv. Biol. 34, 688–696 (2019).PubMed 
    Article 

    Google Scholar 
    Patykowski, J. et al. The effect of prescribed burning on plant rarity in a temperate forest. Ecol. Evol. 8, 1714–1725 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Ames, G. M., Wall, W. A., Hohmann, M. G. & Wright, J. P. Trait space of rare plants in a fire-dependent ecosystem. Conserv. Biol. 31, 903–911 (2017).PubMed 
    Article 

    Google Scholar 
    Foster, C. N. et al. Effects of fire regime on plant species richness and composition differ among forest, woodland and heath vegetation. Appl. Veg. Sci. 21, 132–143 (2018).Article 

    Google Scholar 
    Fernández-García, V. et al. Fire regimes shape diversity and traits of vegetation under different climatic conditions. Sci. Total Environ. 716, 137137 (2020).ADS 
    PubMed 
    Article 
    CAS 

    Google Scholar 
    Bassett, M., Leonard, S. W. J., Chia, E. K., Clarke, M. F. & Bennett, A. F. Interacting effects of fire severity, time since fire and topography on vegetation structure after wildfire. For. Ecol. Manag. 396, 26–34 (2017).Article 

    Google Scholar 
    Miller, B. P., Symons, D. R. & Barrett, M. D. Persistence of rare species depends on rare events: Demography, fire response and phenology of two plant species endemic to a semiarid Banded Iron Formation range. Aust. J. Bot. 67, 268–280 (2019).Article 

    Google Scholar 
    Etchells, H., O’Donnell, A. J., Lachlan McCaw, W. & Grierson, P. F. Fire severity impacts on tree mortality and post-fire recruitment in tall eucalypt forests of southwest Australia. For. Ecol. Manag. 459, 117850 (2020).Article 

    Google Scholar 
    Bradstock, R. A., Tozer, M. G. & Keith, D. A. Effects of high frequency fire on floristic composition and abundance in a fire-prone heathland near Sydney. Aust. J. Bot. 45, 641–655 (1997).Article 

    Google Scholar 
    Penman, T. D., Binns, D. L., Brassil, T. E., Shiels, R. J. & Allen, R. M. Long-term changes in understorey vegetation in the absence of wildfire in south-east dry sclerophyll forests. Aust. J. Bot. 57, 533–540 (2010).Article 

    Google Scholar 
    Ooi, M. K. The importance of fire season when managing threatened plant species: A long-term case-study of a rare Leucopogon species (Ericaceae). J. Environ. Manage. 236, 17–24 (2019).PubMed 
    Article 

    Google Scholar 
    Pausas, J. G., Bradstock, R. A., Keith, D. A. & Keeley, J. E. Plant functional traits in relation to fire in crown-fire ecosystems. Ecology 85, 1085–1100 (2004).Article 

    Google Scholar 
    Australian Bureau of Meteorology. Climate Data Online. www.bom.gov.au (2019).Abell, R. S. Geoscience map of Jervis Bay Territory and Beecroft peninsula (1:25000 scale). Australian Geological Survey Organisation (1992).Taws, N. Vegetation survey and mapping of Jervis Bay Territory. (Taws Botanical Research, 1997).Taylor, G., Abell, R. & Paterson, I. Geology, geomorphology, soils and earth resources. in Jervis Bay (eds. Cho Arthur, G., Georges, Stoutjesdikj Richard, R., & Longmore) .-. (Australian Nature Conservation Agency, 1995).Keith, D. A. Ocean shores to desert dunes: the native vegetation of NSW and the ACT (Selected Extracts). (Department of Environment and Conservation (NSW), 2004).Keith, D. A. & Tozer, M. G. Vegetation dynamics in coastal heathlands of the Sydney basin. in Proceedings of the Linnean Society of New South Wales vol. 134 (2012).Lindenmayer, D. B. et al. Contrasting mammal responses to vegetation type and fire. Wildl. Res. 35, 395–408 (2008).Article 

    Google Scholar 
    Bradstock, R. A. & Kenny, B. J. An application of plant functional types to fire management in a conservation reserve in southeastern Australia. J. Veg. Sci. 14, 345–354 (2003).Article 

    Google Scholar 
    Bowd, E. J., Banks, S. C., Bissett, A., May, T. W. & Lindenmayer, D. B. Direct and indirect disturbance impacts in forests. Ecol. Lett. https://doi.org/10.1111/ele.13741 (2021).Article 
    PubMed 

    Google Scholar 
    Thompson, C. G., Kim, R. S., Aloe, A. M. & Becker, B. J. Extracting the variance inflation factor and other multicollinearity diagnostics from typical regression results. Basic Appl. Soc. Psychol. 39, 81–90 (2017).Article 

    Google Scholar 
    Fox, J. & Weisberg (Sage, 2019).
    Google Scholar 
    Venables, W. N. R., B. D. Modern Applied Statistics with S. Fourth Edition. (Springer, 2002).Morrison, D. A. et al. Effects of fire frequency on plant species composition of sandstone communities in the Sydney region: Inter-fire interval and time-since-fire. Aust. J. Ecol. 20, 239–247 (1995).ADS 
    Article 

    Google Scholar 
    Burnham, K. P., Anderson, D. R. & Huyvaert, K. P. AIC model selection and multimodel inference in behavioral ecology: Some background, observations, and comparisons. Behav. Ecol. Sociobiol. 65, 23–35 (2011).Article 

    Google Scholar 
    Hartig, F. DHARMa: Residual diagnostics for hierarchical (multi-Level / mixed) regression models. (2020).Falster, D. et al. AusTraits, a curated plant trait database for the Australian flora. Sci. Data 8, 254 (2021).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Tozer, M. G. & Bradstock, R. A. Fire-mediated effects of overstorey on plant species diversity and abundance in an eastern Australian heath. Plant Ecol. 164, 213–223 (2003).Article 

    Google Scholar 
    Gosper, C. R., Yates, C. J., Prober, S. M. & Parsons, B. C. Contrasting changes in vegetation structure and diversity with time since fire in two Australian Mediterranean-climate plant communities. Austral Ecol. 37, 164–174 (2012).Article 

    Google Scholar 
    Foster, C., Barton, P., Robinson, N., MacGregor, C. & Lindenmayer, D. B. Effects of a large wildfire on vegetation structure in a variable fire mosaic. Ecol. Appl. 27, 2369–2381 (2017).CAS 
    PubMed 
    Article 

    Google Scholar 
    Preston, F. W. The commonness, and rarity, of species. Ecology 29, 254–283 (1948).Article 

    Google Scholar 
    McGill, B. J. A renaissance in the study of abundance. Science 314, 770–772 (2006).CAS 
    PubMed 
    Article 

    Google Scholar 
    Silvertown, J. Plant coexistence and the niche. Trends Ecol. Evol. 19, 605–611 (2004).Article 

    Google Scholar 
    Lyons, K. G. & Schwartz, M. W. Rare species loss alters ecosystem function—invasion resistance. Ecol. Lett. 4, 358–365 (2001).Article 

    Google Scholar 
    Dee, L. E. et al. When do ecosystem services depend on rare species?. Trends Ecol. Evol. 34, 746–758 (2019).PubMed 
    Article 

    Google Scholar 
    Smith, M. D. & Knapp, A. K. Dominant species maintain ecosystem function with non-random species loss. Ecol. Lett. 6, 509–517 (2003).Article 

    Google Scholar 
    Lennon, J. J., Koleff, P., Greenwood, J. J. & Gaston, K. J. Contribution of rarity and commonness to patterns of species richness. Ecol. Lett. 7, 81–87 (2004).Article 

    Google Scholar 
    Foster, C. N. et al. Herbivory and fire interact to affect forest understory habitat, but not its use by small vertebrates. Anim. Conserv. 19, 15–25 (2016).Article 

    Google Scholar 
    Lamont, B. B., Enright, N. J. & He, T. Fitness and evolution of resprouters in relation to fire. Plant Ecol. 212, 1945–1957 (2011).Article 

    Google Scholar 
    Tolhurst, K. G. & Turvey, N. D. Effects of bracken (Pteridium esculentum (forst. f.) cockayne) on eucalypt regeneration in west-central Victoria. For. Ecol. Manag. 54, 45–67 (1992).Candeias, M. & Warren, R. J. Rareness starts early for disturbance-dependent grassland plant species. Biodivers. Conserv. 25, 2771–2785 (2016).Article 

    Google Scholar 
    Beadle, N. Soil phosphate and the delimitation of plant communities in eastern Australia. Ecology 35, 370–375 (1954).CAS 
    Article 

    Google Scholar 
    Orians, G. H. & Milewski, A. V. Ecology of Australia: The effects of nutrient-poor soils and intense fires. Biol. Rev. 82, 393–423 (2007).PubMed 
    Article 

    Google Scholar 
    Vesk, P. A. & Westoby, M. Funding the bud bank: A review of the costs of buds. Oikos 106, 200–208 (2004).Article 

    Google Scholar 
    Wilfahrt, P. et al. Temporal rarity is a better predictor of local extinction risk than spatial rarity. Ecology https://doi.org/10.1002/ecy.3504 (2021).Article 
    PubMed 

    Google Scholar 
    Miller, B. P. et al. Persistence of rare species depends on rare events: Demography, fire response and phenology of two plant species endemic to a semiarid Banded Iron Formation range. Aust. J. Bot. 67, 268–280 (2019).Article 

    Google Scholar 
    Gillespie, I. G. & Allen, E. B. Fire and competition in a southern California grassland: Impacts on the rare forb Erodium macrophyllum. J. Appl. Ecol. 41, 643–652 (2004).Article 

    Google Scholar 
    Maire, V. et al. Habitat filtering and niche differentiation jointly explain species relative abundance within grassland communities along fertility and disturbance gradients. New Phytol. 196, 497–509 (2012).PubMed 
    Article 

    Google Scholar 
    Yenni, G., Adler, P. B. & Ernest, S. M. Do persistent rare species experience stronger negative frequency dependence than common species?. Glob. Ecol. Biogeogr. 26, 513–523 (2017).Article 

    Google Scholar 
    Mayberry, R. J. & Elle, E. Conservation of a rare plant requires different methods in different habitats: Demographic lessons from Actaea elata. Oecologia 164, 1121–1130 (2010).ADS 
    PubMed 
    Article 

    Google Scholar 
    Rabinowitz, D. & Rapp, J. K. Dispersal abilities of seven sparse and common grasses froma Missouri prairie. Am. J. Bot. 68, 616–624 (1981).Article 

    Google Scholar 
    McIntyre, S. Comparison of a common, rare and declining plant species in the Asteraceae: Possible causes of rarity. Pac. Conserv. Biol. 2, 177–190 (1995).Article 

    Google Scholar 
    Hopfensperger, K. N. A review of similarity between seed bank and standing vegetation across ecosystems. Oikos 116, 1438–1448 (2007).Article 

    Google Scholar 
    Cross, A. T. et al. Defining the role of fire in alleviating seed dormancy in a rare Mediterranean endemic subshrub. AoB Plants 9, (2017). More

  • in

    Evaluating the temporal and spatio-temporal niche partitioning between carnivores by different analytical method in northeastern Japan

    Gause, G. F. Experimental analysis of Vito Volterra’s mathematical theory of the struggle for existence. Science 79, 16–17 (1934).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Amarasekare, P. Competitive coexistence in spatially structured environments: A synthesis. Ecol. Lett. 6, 1109–1122 (2003).Article 

    Google Scholar 
    HilleRisLambers, J., Adler, P. B., Harpole, W. S., Levine, J. M. & Mayfield, M. M. Rethinking community assembly through the lens of coexistence theory. Annu. Rev. Ecol. Evol. Syst. 43, 227–248 (2012).Article 

    Google Scholar 
    Wisz, M. S. et al. The role of biotic interactions in shaping distributions and realised assemblages of species: Implications for species distribution modelling. Biol. Rev. 88, 15–30 (2013).PubMed 
    Article 

    Google Scholar 
    Frey, S., Fisher, J. T., Burton, A. C. & Volpe, J. P. Investigating animal activity patterns and temporal niche partitioning using camera-trap data: Challenges and opportunities. Remote Sens. Ecol. Conserv. 3, 123–132 (2017).Article 

    Google Scholar 
    Davis, C. L. et al. Ecological correlates of the spatial co-occurrence of sympatric mammalian carnivores worldwide. Ecol. Lett. 21, 1401–1412 (2018).PubMed 
    Article 

    Google Scholar 
    Durant, S. M. Competition refuges and coexistence: An example from Serengeti carnivores. J. Anim. Ecol. 67, 370–386 (1998).Article 

    Google Scholar 
    Fedriani, J. M., Fuller, T. K., Sauvajot, R. M. & York, E. C. Competition and intraguild predation among three sympatric carnivores. Oecologia 125, 258–270 (2000).ADS 
    PubMed 
    Article 

    Google Scholar 
    Kamler, J. F., Ballard, W. B., Gilliland, R. L. & Mote, K. Spatial relationships between swift foxes and coyotes in northwestern Texas. Can. J. Zool. 81, 168–172 (2003).Article 

    Google Scholar 
    Vanak, A. T. et al. Moving to stay in place: Behavioral mechanisms for coexistence of African large carnivores. Ecology 94, 2619–2631 (2013).PubMed 
    Article 

    Google Scholar 
    Donadio, E. & Buskirk, S. W. Diet, morphology, and interspecific killing in carnivora. Am. Nat. 167, 524–536 (2006).PubMed 
    Article 

    Google Scholar 
    Tsunoda, H. et al. Food niche segregation between sympatric golden jackals and red foxes in central Bulgaria. J. Zool. 303, 64–71 (2017).Article 

    Google Scholar 
    Palomares, F. & Caro, T. M. Interspecific killing among mammalian carnivores. Am. Nat. 153, 492–508 (1999).CAS 
    PubMed 
    Article 

    Google Scholar 
    Linnell, J. D. C. & Strand, O. Interference interactions, co-existence and conservation of mammalian carnivores. Divers. Distrib. 6, 169–176 (2000).Article 

    Google Scholar 
    Kamler, J. F., Stenkewitz, U., Klare, U., Jacobsen, N. F. & MacDonald, D. W. Resource partitioning among cape foxes, bat-eared foxes, and black-backed jackals in South Africa. J. Wildl. Manag. 76, 1241–1253 (2012).Article 

    Google Scholar 
    Di Bitetti, M. S., Di Blanco, Y. E., Pereira, J. A., Paviolo, A. & Pírez, I. J. Time Partitioning favors the coexistence of sympatric crab-eating foxes (Cerdocyon thous) and Pampas Foxes (Lycalopex gymnocercus). J. Mammal. 90, 479–490 (2009).Article 

    Google Scholar 
    Lesmeister, D. B., Nielsen, C. K., Schauber, E. M. & Hellgren, E. C. Spatial and temporal structure of a mesocarnivore guild in Midwestern North America. Wildl. Monogr. 191, 1–61 (2015).Article 

    Google Scholar 
    Di Bitetti, M. S., De Angelo, C. D., Di Blanco, Y. E. & Paviolo, A. Niche partitioning and species coexistence in a Neotropical felid assemblage. Acta Oecologica 36, 403–412 (2010).ADS 
    Article 

    Google Scholar 
    Monterroso, P., Alves, P. C. & Ferreras, P. Plasticity in circadian activity patterns of mesocarnivores in southwestern Europe: Implications for species coexistence. Behav. Ecol. Sociobiol. 68, 1403–1417 (2014).Article 

    Google Scholar 
    Tsunoda, H., Ito, K., Peeva, S., Raichev, E. & Kaneko, Y. Spatial and temporal separation between the golden jackal and three sympatric carnivores in a human-modified landscape in central Bulgaria. Zool. Ecol. 28, 172–179 (2018).Article 

    Google Scholar 
    Tsunoda, H. et al. Spatio-temporal partitioning facilitates mesocarnivore sympatry in the Stara Planina Mountains, Bulgaria. Zoology 141, 125801 (2020).PubMed 
    Article 

    Google Scholar 
    Ramesh, T., Kalle, R., Sankar, K. & Qureshi, Q. Spatio-temporal partitioning among large carnivores in relation to major prey species in Western Ghats. J. Zool. 287, 269–275 (2012).Article 

    Google Scholar 
    Gómez-Ortiz, Y., Monroy-Vilchis, O. & Castro-Arellano, I. Temporal coexistence in a carnivore assemblage from central Mexico: Temporal-domain dependence. Mammal Res. 64, 333–342 (2019).Article 

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

    Google Scholar 
    Meredith, M. & Ridout, M. Overlap: Estimates of coefficient of overlapping for animal activity patterns. https://cran.r-project.org/web/packages/overlaphttps://cran.r-project.org/web/packages/overlap/index.html (2018).Marinho, P. H., Fonseca, C. R., Sarmento, P., Fonseca, C. & Venticinque, E. M. Temporal niche overlap among mesocarnivores in a Caatinga dry forest. Eur. J. Wildl. Res. 66, 1–13 (2020).Article 

    Google Scholar 
    Vilella, M., Ferrandiz-Rovira, M. & Sayol, F. Coexistence of predators in time: Effects of season and prey availability on species activity within a Mediterranean carnivore guild. Ecol. Evol. 10, 11408–11422 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Zhao, G. et al. Spatio-temporal coexistence of sympatric mesocarnivores with a single apex carnivore in a fine-scale landscape. Glob. Ecol. Conserv. 21, e00897 (2020).Article 

    Google Scholar 
    Farmer, M. J., Allen, M. L., Olson, E. R., Van Stappen, J. & Van Deelen, T. R. Agonistic interactions and island biogeography as drivers of carnivore spatial and temporal activity at multiple scales. Can. J. Zool. 99, 309–317 (2021).Article 

    Google Scholar 
    Watabe, R. & Saito, M. U. Diel activity patterns of three sympatric medium-sized carnivores during winter and spring in a heavy snowfall area in northeastern Japan. Mammal Study 46, 69–75 (2021).Article 

    Google Scholar 
    Lashley, M. A. et al. Estimating wildlife activity curves: comparison of methods and sample size. Sci. Rep. 8, 4173 (2018).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Niedballa, J., Wilting, A., Sollmann, R., Hofer, H. & Courtiol, A. Assessing analytical methods for detecting spatiotemporal interactions between species from camera trapping data. Remote Sens. Ecol. Conserv. 5, 272–285 (2019).Article 

    Google Scholar 
    Karanth, K. U. et al. Spatio-temporal interactions facilitate large carnivore sympatry across a resource gradient. Proc. R. Soc. B Biol. Sci. 284, 20161860 (2017).Article 

    Google Scholar 
    Cusack, J. J. et al. Revealing kleptoparasitic and predatory tendencies in an African mammal community using camera traps: A comparison of spatiotemporal approaches. Oikos 126, 812–822 (2017).Article 

    Google Scholar 
    Balme, G. et al. Big cats at large: density, structure, and spatio-temporal patterns of a leopard population free of anthropogenic mortality. Popul. Ecol. 61, 256–267 (2019).Article 

    Google Scholar 
    Li, Z. et al. Coexistence of two sympatric flagship carnivores in the human-dominated forest landscapes of Northeast Asia. Landsc. Ecol. 34, 291–305 (2019).Article 

    Google Scholar 
    Lahkar, D., Ahmed, M. F., Begum, R. H., Das, S. K. & Harihar, A. Inferring patterns of sympatry among large carnivores in Manas National Park: A prey-rich habitat influenced by anthropogenic disturbances. Anim. Conserv. 24, 589–601 (2021).Article 

    Google Scholar 
    Paúl, M. J., Layna, J. F., Monterroso, P. & Álvares, F. Resource partitioning of sympatric African Wolves (Canis lupaster) and side-striped jackals (Canis adustus) in an arid environment from West Africa. Diversity 12, 477 (2020).Article 

    Google Scholar 
    Prat-Guitart, M., Onorato, D. P., Hines, J. E. & Oli, M. K. Spatiotemporal pattern of interactions between an apex predator and sympatric species. J. Mammal. 101, 1279–1288 (2020).Article 

    Google Scholar 
    Stone, L. & Roberts, A. The checkerboard score and species distributions. Oecologia 85, 74–79 (1990).ADS 
    PubMed 
    Article 

    Google Scholar 
    Griffith, D. M., Veech, J. A. & Marsh, C. J. Cooccur: Probabilistic species co-occurrence analysis in r. J. Stat. Softw. 69, 1–17 (2016).Article 

    Google Scholar 
    Noor, A., Mir, Z. R., Veeraswami, G. G. & Habib, B. Activity patterns and spatial co-occurrence of sympatric mammals in the moist temperate forest of the Kashmir Himalaya, India. Folia Zool. 66, 231–241 (2017).Article 

    Google Scholar 
    de Satgé, J., Teichman, K. & Cristescu, B. Competition and coexistence in a small carnivore guild. Oecologia 184, 873–884 (2017).ADS 
    PubMed 
    Article 

    Google Scholar 
    Kass, J. M., Tingley, M. W., Tetsuya, T. & Koike, F. Co-occurrence of invasive and native carnivorans affects occupancy patterns across environmental gradients. Biol. Invasions 22, 2251–2266 (2020).Article 

    Google Scholar 
    Louppe, V., Herrel, A., Pisanu, B., Grouard, S. & Veron, G. Assessing occupancy and activity of two invasive carnivores in two Caribbean islands: implications for insular ecosystems. J. Zool. 313, 182–194 (2020).Article 

    Google Scholar 
    Proulx, G. et al. World distribution and status of the genus Martes in 20. In Martens and Fishers (Martes) in Human-Altered Environments (eds Harrison, D. J. et al.) 21–76 (Springer, Berlin, 2005). https://doi.org/10.1007/b99487.Chapter 

    Google Scholar 
    Ohdachi, S. D., Ishibashi, Y., Iwasa, M., Fukuki, D. & Saitoh, T. The Wild Mammals of Japan 2nd edn. (Shokadoh Book Seller, Kyoto, 2015).
    Google Scholar 
    Kauhala, K. & Saeki, M. Nyctereutes procyonoides. The IUCN Red List of Threatened Species. https://www.iucnredlist.org/species/14925/85658776 (2016).Yamamoto, Y. Comparative analyses on food habits of Japanese marten, red fox, badger and raccoon dog in the Mt. Nyugasa, Nagano Prefecture, Japan. Nat. Environ. Sci. Res. 7, 45–52 (1994) (in Japanese with English summary).
    Google Scholar 
    Hisano, M. et al. A comparison of visual and genetic techniques for identifying Japanese marten scats enabling diet examination in relation to seasonal food availability in a sub-alpine area of Japan. Zool. Sci. 34, 137–146 (2017).Article 

    Google Scholar 
    Lindstrom, E. R., Brainerd, S. M., Helldin, J. O. & Overskaug, K. Pine marten-red fox interactions: A case of intraguild predation?. Ann. Zool. Fenn. 32, 123–130 (1995).
    Google Scholar 
    Waggershauser, C. N., Ruffino, L., Kortland, K. & Lambin, X. Lethal interactions among forest-grouse predators are numerous, motivated by hunger and carcasses, and their impacts determined by the demographic value of the victims. Ecol. Evol. 11, 7164–7186 (2021).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Watabe, R., Saito, M. U., Enari, H. S. & Enari, H. Mammalian fauna of the Kaminagawa Experimental Forest of Yamagata University detected by camera traps. Tohoku J. For. Sci. 25, 37–40 (2020) (in Japanese).
    Google Scholar 
    Hofmeester, T. R., Rowcliffe, J. M. & Jansen, P. A. A simple method for estimating the effective detection distance of camera traps. Remote Sens. Ecol. Conserv. 3, 81–89 (2017).Article 

    Google Scholar 
    Di Bitetti, M. S., Paviolo, A. & De Angelo, C. Camera trap photographic rates on roads vs. off roads: Location does matter. Mastozoología Neotrop. 21, 37–46 (2014).
    Google Scholar 
    Borcard, D. & Legendre, P. Is the Mantel correlogram powerful enough to be useful in ecological analysis? A simulation study. Ecology 93, 1473–1481 (2012).PubMed 
    Article 

    Google Scholar 
    Oksanen, J. et al. Vegan: community ecology package. https://cran.r-project.org/web/packages/veganhttps://cran.r-project.org/web/packages/vegan/index.html (2019).R Core Team. R: a language environment for statistical computing. r foundation for statistical computing, Vienna, Austria. https://www.r-project.org/https://www.r-project.org/ (2021).Linkie, M. & Ridout, M. S. Assessing tiger-prey interactions in Sumatran rainforests. J. Zool. 284, 224–229 (2011).Article 

    Google Scholar 
    Watabe, R. & Saito, M. U. Effects of vehicle-passing frequency on forest roads on the activity patterns of carnivores. Landsc. Ecol. Eng. 17, 225–231 (2021).Article 

    Google Scholar 
    Furukawa, G. genkiFurukawa/rSetDayNightAttr documentation. https://rdrr.io/github/genkiFurukawa/rSetDayNightAhttps://rdrr.io/github/genkiFurukawa/rSetDayNightAttr/ (2019).Mielke, P. W., Berry, K. J. & Johnson, E. S. Multi-response permutation procedures for a priori classifications. Commun. Stat. Theory Methods 5, 1409–1424 (1976).MATH 
    Article 

    Google Scholar 
    Kronfeld-Schor, N. & Dayan, T. Partitioning of time as an ecological resource. Annu. Rev. Ecol. Evol. Syst. 34, 153–181 (2003).Article 

    Google Scholar 
    Monterroso, P., Alves, P. C. & Ferreras, P. Catch me if you can: Diel activity patterns of mammalian prey and predators. Ethology 119, 1044–1056 (2013).Article 

    Google Scholar 
    Hendrichsen, D. K. & Tyler, N. J. C. How the timing of weather events influences early development in a large mammal. Ecology 95, 1737–1745 (2014).CAS 
    PubMed 
    Article 

    Google Scholar 
    Herfindal, I. et al. Weather affects temporal niche partitioning between moose and livestock. Wildlife Biol. https://doi.org/10.2981/wlb.00275 (2017).Article 

    Google Scholar 
    Haswell, P. M., Jones, K. A., Kusak, J. & Hayward, M. W. Fear, foraging and olfaction: How mesopredators avoid costly interactions with apex predators. Oecologia 187, 573–583 (2018).ADS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Barrull, J. et al. Factors and mechanisms that explain coexistence in a Mediterranean carnivore assemblage: An integrated study based on camera trapping and diet. Mamm. Biol. 79, 123–131 (2014).Article 

    Google Scholar 
    Tattersall, E. R., Burgar, J. M., Fisher, J. T. & Burton, A. C. Boreal predator co-occurrences reveal shared use of seismic lines in a working landscape. Ecol. Evol. 10, 1678–1691 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Moll, R. J. et al. Humans and urban development mediate the sympatry of competing carnivores. Urban Ecosyst. 21, 765–778 (2018).Article 

    Google Scholar 
    McCreadie, J. W. & Bedwell, C. R. Patterns of co-occurrence of stream insects and an examination of a causal mechanism: Ecological checkerboard or habitat checkerboard?. Insect Conserv. Divers. 6, 105–113 (2013).Article 

    Google Scholar  More

  • in

    More than dollars: mega-review finds 50 ways to value nature

    Relatively few studies try to understand the value of cultural heritage sites such as Nachi Falls, which is also a pilgrimage route in Japan’s Kii mountain range.Credit: James Fichera/Getty

    There are more than 50 ways to value the environment, but most research and policymaking focuses on just a handful of methods. These include counting species and evaluating the cost of replacing a service provided by nature. Yet assessing nature in purely monetary terms can also be harmful to people and the environment, according to the world’s largest assessment of environmental valuation.“Policymaking largely disregards the multiple ways in which nature matters to people,” especially Indigenous people and low-income communities, says the report from the Intergovernmental Science-Policy Panel on Biodiversity and Ecosystem Services (IPBES).For example, in proposals for hydroelectric dams, the needs of affected communities are often seen as secondary to those of urban consumers — especially if communities are required to be displaced, resulting in people losing livelihoods and being compelled to change their way of life, the report finds.The world’s failure to properly value biodiversity has caused a long-term decline in a variety of services that the environment provides, said Anne Larigauderie, an ecologist who leads the IPBES secretariat, at the report’s launch on 11 July. “The capacity to pollinate crops, or regulate water, has been in decline for 50 years,” she said.There is strong evidence that valuing nature on the basis of market prices is contributing to the present biodiversity crisis, said Unai Pascual, an economist at the Basque Centre for Climate Change in Leioa, Spain, at the launch in Bonn, Germany. “Many other values are ignored in favour of short-term profit and economic growth,” added Pascual, who co-chaired the assessment.A summary for policymakers was approved by 139 governments on 8 July. The full assessment report is expected to be released ahead of the Conference of the Parties to the UN Convention on Biological Diversity, which takes place in Montreal in December. This meeting is expected to agree a new set of targets and indicators for biodiversity conservation.Studies of natureEighty-two researchers from around the world, with areas of expertise spanning the sciences, social sciences and humanities, identified 79,000 studies in environmental valuation, and found that their number has been increasing by 10% a year for four decades. But these studies also rarely lead to policy changes. The researchers selected 1,163 of the studies for in-depth review, and found that only for 5% of these cases were recommendations adopted by decision makers.Half of the studies selected for in-depth review used biophysical indicators, such as numbers of species, or quantity of forest biomass. Another 26% used monetary indicators, such as how much it would cost if pollination needed to be carried out by humans, or the amounts that governments pay farmers to conserve biodiversity on agricultural land.Only one-fifth of the studies valued biodiversity according to sociocultural criteria. Those that did included studies on the importance to people of a sacred site; and research on the value that someone attaches to the place where they grew up. Sociocultural values do not necessarily have a numerical quantity, or price tag. The value of sacred sites does not need to be turned into dollars, or euros, Sander Jacobs, one of the IPBES authors and an ecologist at the Research Institute of Nature and Forests in Brussels, said at the report’s launch.The report’s authors found that most studies do not consider multiple values, even when the evidence shows that doing so leads to better outcomes for the environment. The team found that few scientists consult or involve the people who live and work in regions of high biodiversity. Only 2% of the studies reviewed in depth reported having done so. And just 1% involved people in all the steps from designing a study to publishing it.“We need to build coalitions of scientists from different disciplines. But science needs allies too,” Pascual says. “Scientists need to be humble and invite those who represent other ways of knowing. Such a coalition could provide a solutions-oriented approach to the biodiversity and climate crises.” More

  • in

    Using PVA and captive breeding to balance trade-offs in the rescue of the island dibbler onto a new island ark

    Burbidge, A. A. & Abbott, I. Mammals on Western Australian islands: occurrence and preliminary analysis. Aust. J. Zool. 65, 183–195. https://doi.org/10.1071/zo17046 (2017).Article 

    Google Scholar 
    Fischer, J. & Lindenmayer, D. B. An assessment of the published results of animal relocations. Biol. Conserv. 96, 1–11. https://doi.org/10.1016/S0006-3207(00)00048-3 (2000).Article 

    Google Scholar 
    Legge, S. et al. Havens for threatened Australian mammals: the contributions of fenced areas and offshore islands to the protection of mammal species susceptible to introduced predators. Wildl. Res. 45, 627–644. https://doi.org/10.1071/wr17172 (2018).Article 

    Google Scholar 
    Morris, K. et al. Forty years of fauna translocations in Western Australia: lessons learned. In Advances in Reintroduction Biology of Australian and New Zealand Fauna (eds Armstrong, D. P. et al.) (CSIRO Publishing, 2015).
    Google Scholar 
    Seddon, P. J., Moro, D., Mitchell, N. J., Chauvenet, A. & Mawson, P. Proactive conservation or planned
    invasion? Past, current and future use of
    assisted colonisation. In Advances in Reintroduction Biology of Australian and New Zealand Fauna (eds Armstrong, D. P. et al.) (CSIRO Publishing, 2015).
    Google Scholar 
    Weeks, A. R. et al. Conserving and enhancing genetic
    diversity in translocation programmes. In Advances in Reintroduction Biology of Australian and New Zealand Fauna (eds Armstrong, D. P. et al.) (CSIRO Publishing, 2015).
    Google Scholar 
    IUCN/SSC. Guidelines for Reintroductions and Other Conservation Translocations. Report No. 1.0, viiii + 57 (Gland, Switzerland, 2013).Allendorf, F. W. & Ryman, N. The role of genetics in population viability analysis. In Population Viability Analysis (eds Beissinger, S. R. & McCullough, D. R.) 50–85 (University of Chicago Press, 2002).
    Google Scholar 
    Gilpin, M. E. & Soule, M. E. Minimum viable populations: process of species extinctions. In Conservation Biology: The Science of Scarcity and Diversity (ed Soule, M. E.) 19–34 (Sinauer, 1986).
    Google Scholar 
    Frankham, R. et al. Predicting the probability of outbreeding depression. Conserv. Biol. 25, 465–475. https://doi.org/10.1111/j.1523-1739.2011.01662.x (2011).Article 
    PubMed 

    Google Scholar 
    IUCN. IUCN Red List Categories and Criteria: Version 3.1. iv + 32 (Gland, Switzerland Cambridge, UK, 2012).Willoughby, J. R. et al. The reduction of genetic diversity in threatened vertebrates and new recommendations regarding IUCN conservation rankings. Biol. Conserv. 191, 495–503. https://doi.org/10.1016/j.biocon.2015.07.025 (2015).Article 

    Google Scholar 
    Allendorf, F. W. Genetic drift and the loss of alleles versus heterozygosity. Zoo Biol. 5, 181–190. https://doi.org/10.1002/zoo.1430050212 (1986).Article 

    Google Scholar 
    Frankham, R. Genetics and extinction. Biol. Conserv. 126, 131–140. https://doi.org/10.1016/j.biocon.2005.05.002 (2005).Article 

    Google Scholar 
    Easton, L. J., Bishop, P. J. & Whigham, P. A. Balancing act: modelling sustainable release numbers for translocations. Anim. Conserv. https://doi.org/10.1111/acv.12558 (2019).Article 

    Google Scholar 
    Allendorf, F. W., England, P. R., Luikart, G., Ritchie, P. A. & Ryman, N. Genetic effects of harvest on wild animal populations. Trends Ecol. Evol. 23, 327–337. https://doi.org/10.1016/j.tree.2008.02.008 (2008).Article 
    PubMed 

    Google Scholar 
    Snyder, N. F. R. & Snyder, H. The California Condor: A Saga of Natural History and Conservation 1st edn. (Princeton University Press, 2000).
    Google Scholar 
    Kuchling, G., Burbridge, A. A., Page, M. & Olejnik, C. Western Swamp Tortoise Pseudemydura umbrina: slow and steady wins the race. In Recovering Australian Threatened Species: A Book of Hope (eds Garnett, S. et al.) 217–226 (CSIRO, 2018).
    Google Scholar 
    Hogg, C. J. Preserving Australian native fauna: zoo-based breeding programs as part of a more unified strategic approach. Aust. J. Zool. 61, 101–108. https://doi.org/10.1071/zo13014 (2013).Article 

    Google Scholar 
    Snyder, N. F. R. et al. Limitations of captive breeding in endangered species recovery. Conserv. Biol. 10, 338–348. https://doi.org/10.1046/j.1523-1739.1996.10020338.x (1996).Article 

    Google Scholar 
    Frankham, R. Genetic rescue of small inbred populations: meta-analysis reveals large and consistent benefits of gene flow. Mol. Ecol. 24, 2610–2618. https://doi.org/10.1111/mec.13139 (2015).Article 
    PubMed 

    Google Scholar 
    Weeks, A. R. et al. Assessing the benefits and risks of translocations in changing environments: A genetic perspective. Evol. Appl. 4, 709–725. https://doi.org/10.1111/j.1752-4571.2011.00192.x (2011).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Coyne, J. A. & Orr, H. A. Speciation (Sinauer, 2004).
    Google Scholar 
    Edmands, S. Between a rock and a hard place: Evaluating the relative risks of inbreeding and outbreeding for conservation and management. Mol. Ecol. 16, 463–475. https://doi.org/10.1111/j.1365-294X.2006.03148.x (2007).Article 
    PubMed 

    Google Scholar 
    Armbruster, P., Bradshaw, W. E., Steiner, A. L. & Holzapfel, C. M. Evolutionary responses to environmental stress by the pitcher-plant mosquito, Wyeomyia smithii. Heredity 83, 509–519. https://doi.org/10.1038/sj.hdy.6886040 (1999).Article 
    PubMed 

    Google Scholar 
    Edmands, S. Heterosis and outbreeding depression in interpopulation crosses spanning a wide range of divergence. Evolution 53, 1757–1768. https://doi.org/10.2307/2640438 (1999).Article 
    PubMed 

    Google Scholar 
    Marr, A. B., Keller, L. F. & Arcese, P. Heterosis and outbreeding depression in descendants of natural immigrants to an inbred population of song sparrows (Melospiza melodia). Evolution 56, 131–142 (2002).Article 

    Google Scholar 
    Tymchuk, W. E., Sundstrom, L. F. & Devlin, R. H. Growth and survival trade-offs and outbreeding depression in rainbow trout (Oncorhynchus mykiss). Evolution 61, 1225–1237. https://doi.org/10.1111/j.1558-5646.2007.00102.x (2007).Article 
    PubMed 

    Google Scholar 
    Friend, J. A. Dibbler (Parantechinus apicalis) Recovery Plan July 2003-June 2013 (Department of Conserv. and Land Management, 2003).
    Google Scholar 
    Miller, S., Bencini, R., Mills, H. & Moro, D. Food availability for the dibbler (Parantechinus apicalis) on Boullanger and Whitlock Islands, Western Australia. Wildl. Res. 30, 649–654. https://doi.org/10.1071/wr01082 (2003).Article 

    Google Scholar 
    Mills, H. R. & Bencini, R. New evidence for facultative male die-off in island populations of dibblers, Parantechinus apicalis. Aust. J. Zool. 48, 501–510. https://doi.org/10.1071/zo00025 (2000).Article 

    Google Scholar 
    Mills, H. R., Moro, D. & Spencer, P. B. S. Conservation significance of island versus mainland populations: A case study of dibblers (Parantechinus apicalis) in Western Australia. Anim. Conserv. 7, 387–395. https://doi.org/10.1017/s1367943004001568 (2004).Article 

    Google Scholar 
    Woolley, P. A. Reproductive pattern of captive Boullanger Island dibblers, Parantechinus apicalis (Marsupialia, Dasyuridae). Wildl. Res. 18, 157–163. https://doi.org/10.1071/wr9910157 (1991).Article 

    Google Scholar 
    Burbridge, A. A. & Woinarski, J. C. Z. Parantechinus apicalis. The IUCN Red List of Threatened Species 2016: e.T16138A21944584. https://www.iucnredlist.org/species/16138/21944584 (2016).Friend, J. A. Island home: A new start for dibblers. Landscope 33, 39–42 (2017).
    Google Scholar 
    Moro, D. Translocation of captive-bred dibblers Parantechinus apicalis (Marsupialia: Dasyuridae) to Escape Island, Western Australia. Biol. Conserv. 111, 305–315. https://doi.org/10.1016/s0006-3207(02)00296-3 (2003).Article 

    Google Scholar 
    Thavornkanlapachai, R., Mills, H. R., Ottewell, K., Friend, J. A. & Kennington, W. J. Temporal variation in the genetic composition of an endangered marsupial reflects reintroduction history. Diversity https://doi.org/10.3390/d13060257 (2021).Article 

    Google Scholar 
    Morris, K., Page, M., Thomas, N. & Ottewell, K. A Strategic Framework for the Reconstruction and Conservation of the Vertebrate Fauna of Dirk Hartog Island 2016–2030. 26 (Department of Parks and Wildlife, 2017).
    Google Scholar 
    Thavornkanlapachai, R. Genetic Consequences of Genetic Mixing in Mammal Translocations in Western Australia Using Case Studies of Burrowing Bettongs and Dibblers. Doctor of Philosophy thesis, University of Western Australia (2016).Akcakaya, H. R. & Sjogren-Gulve, P. Population viability analyses in Conserv. planning: an overview. Ecol. Bull. 48, 9–21 (2000).
    Google Scholar 
    Beissinger, S. R. & McCullough, D. R. Population Viability Analysis (The University of Chicago Press, 2002).
    Google Scholar 
    Lindenmayer, D. B., Clark, T. W., Lacy, R. C. & Thomas, V. C. Population viability analysis as a tool in wildlife conservation policy—With reference to Australia. Environ. Manag. 17, 745–758. https://doi.org/10.1007/bf02393895 (1993).ADS 
    Article 

    Google Scholar 
    Pacioni, C., Wayne, A. F. & Page, M. Guidelines for genetic management in mammal translocation programs. Biol. Conserv. 237, 105–113. https://doi.org/10.1016/j.biocon.2019.06.019 (2019).Article 

    Google Scholar 
    White, D. J. et al. Genetic consequences of multiple translocations of the banded hare-wallaby in Western Australia. Diversity https://doi.org/10.3390/d12120448 (2020).Article 

    Google Scholar 
    Dickman, C. R. & Braithwaite, R. W. Postmating mortality of males in the Dasyurid marsupials, Dasyurus and Parantechinus. J. Mammal. 73, 143–147. https://doi.org/10.2307/1381875 (1992).Article 

    Google Scholar 
    Lambert, C. & Mills, H. Husbandry and breeding of the dibbler Parantechinus apicalis at Perth Zoo. Int. Zoo Yearb. 40, 290–301 (2006).Article 

    Google Scholar 
    Mills, H. R., Bradshaw, F. J., Lambert, C., Bradshaw, S. D. & Bencini, R. Reproduction in the marsupial dibbler, Parantechinus apicalis; differences between island and mainland populations. Gen. Comp. Endocrinol. 178, 347–354. https://doi.org/10.1016/j.ygcen.2012.06.013 (2012).CAS 
    Article 
    PubMed 

    Google Scholar 
    Fisher, D. O., Dickman, C. R., Jones, M. E. & Blomberg, S. P. Sperm competition drives the evolution of suicidal reproduction in mammals. Proc. Natl. Acad. Sci. USA 110, 17910–17914. https://doi.org/10.1073/pnas.1310691110 (2013).ADS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Stewart, A. Dibblers on the Jurien Islands: The Influence of Burrowing Seabirds and the Potential for Competition from Other Species. PhD thesis, University of Western Australia (2006).Sunnucks, P. & Hales, D. F. Numerous transposed sequences of mitochondrial cytochrome oxidase I-II in aphids of the genus Sitobion (Hemiptera: Aphididae). Mol. Biol. Evol. 13, 510–524. https://doi.org/10.1093/oxfordJ.s.molbev.a025612 (1996).CAS 
    Article 
    PubMed 

    Google Scholar 
    Van Oosterhout, C., Hutchinson, 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. https://doi.org/10.1111/j.1471-8286.2004.00684.x (2004).CAS 
    Article 

    Google Scholar 
    Goudet, J. FSTAT (Version 1.2): A computer program to calculate F-statistics. J. Heredity 86, 485–486. https://doi.org/10.1093/oxfordJ.s.jhered.a111627 (1995).Article 

    Google Scholar 
    Peakall, R. & Smouse, P. GenAlEx 6.5: Genetic analysis in Excel. Population genetic software for teaching and research—An update. Bioinformatics 28, 2537–2239. https://doi.org/10.1093/bioinformatics/bts460 (2012).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Peakall, R. & Smouse, P. E. GENALEX 6: Genetic analysis in Excel. Population genetic software for teaching and research. Mol. Ecol. Notes 6, 288–295. https://doi.org/10.1111/j.1471-8286.2005.01155.x (2006).Article 

    Google Scholar 
    R Core Team. R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing. https://www.R-project.org/ (2018).Pritchard, J. K., Stephens, M. & Donnelly, P. Inference of population structure using multilocus genotype data. Genetics 155, 945–959 (2000).CAS 
    Article 

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

    Google Scholar 
    Evanno, G., Regnaut, S. & Goudet, J. Detecting the number of clusters of individuals using the software STRUCTURE: A simulation study. Mol. Ecol. 14, 2611–2620. https://doi.org/10.1111/j.1365-294X.2005.02553.x (2005).CAS 
    Article 
    PubMed 

    Google Scholar 
    Do, C. et al. NEESTIMATOR v2: Re-implementation of software for the estimation of contemporary effective population size (Ne) from genetic data. Mol. Ecol. Resour. 14, 209–214. https://doi.org/10.1111/1755-0998.12157 (2014).CAS 
    Article 
    PubMed 

    Google Scholar 
    Waples, R. S. A bias correction for estimates of effective population size based on linkage disequilibrium at unlinked gene loci. Conserv. Genet. 7, 167–184. https://doi.org/10.1007/s10592-005-9100-y (2006).Article 

    Google Scholar 
    Cornuet, J. M. & Luikart, G. Description and power analysis of two tests for detecting recent population bottlenecks from allele frequency data. Genetics 144, 2001–2014 (1996).CAS 
    Article 

    Google Scholar 
    Piry, S., Luikart, G. & Cornuet, J. M. BOTTLENECK: A computer program for detecting recent reductions in the effective population size using allele frequency data. J. Heredity 90, 502–503. https://doi.org/10.1093/jhered/90.4.502 (1999).Article 

    Google Scholar 
    Queller, D. C. & Goodnight, K. F. Estimating relatedness using genetic markers. Evolution 43, 258–275. https://doi.org/10.2307/2409206 (1989).Article 
    PubMed 

    Google Scholar 
    Lacy, R. C. & Pollak, J. P. VORTEX: A Stochastic Simulation of the Extinction Process. Version 10.0 (Brookfield, Illinois, USA, 2014).Lacy, R. C. VORTEX—A computer simulation model for population viability analysis. Wildl. Res. 20, 45–65. https://doi.org/10.1071/wr9930045 (1993).Article 

    Google Scholar 
    Parrott, M. L., Ward, S. J., Temple-Smith, P. D. & Selwood, L. Effects of drought on weight, survival and breeding success of agile antechinus (Antechinus agilis), dusky antechinus (A. swainsonii) and bush rats (Rattus fuscipes). Wildl. Res. 34, 437–442. https://doi.org/10.1071/wr07071 (2007).Article 

    Google Scholar 
    Rhind, S. G. & Bradley, J. S. The effect of drought on body size, growth and abundance of wild brush-tailed phascogales (Phascogale tapoatafa) in south-western Australia. Wildl. Res. 29, 235–245. https://doi.org/10.1071/wr01014 (2002).Article 

    Google Scholar 
    Bureau of Meteorology. Monthly rainfall Jurien Bay. Australian Government. http://www.bom.gov.au/jsp/ncc/cdio/weatherData/av?p_nccObsCode=139&p_display_type=dataFile&p_startYear=&p_c=&p_stn_num=009131 (2020).McCarthy, M. A., Burgman, M. A. & Ferson, S. Sensitivity analysis for models of population viability. Biol. Conserv. 73, 93–100. https://doi.org/10.1016/0006-3207(95)00046-7 (1995).Article 

    Google Scholar 
    Waples, R. S. & Do, C. Linkage disequilibrium estimates of contemporary Ne using highly variable genetic markers: A largely untapped resource for applied conservation and evolution. Evol. Appl. 3, 244–262. https://doi.org/10.1111/j.1752-4571.2009.00104.x (2010).Article 
    PubMed 

    Google Scholar 
    Woinarski, J. C. Z., Burbidge, A. A. & Harrison, P. L. Ongoing unraveling of a continental fauna: Decline and extinction of Australian mammals since European settlement. Proc. Natl. Acad. Sci. USA 112, 4531–4540. https://doi.org/10.1073/pnas.1417301112 (2015).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Eldridge, M. D. B. et al. Unprecedented low levels of genetic variation and inbreeding depression in an island population of the black-footed rock-wallaby. Conserv. Biol. 13, 531–541. https://doi.org/10.1046/j.1523-1739.1999.98115.x (1999).Article 

    Google Scholar 
    Frankham, R. Do island populations have less genetic variation than mainland populations?. Heredity 78, 311–327. https://doi.org/10.1038/hdy.1997.46 (1997).Article 
    PubMed 

    Google Scholar 
    Wright, S. Evoluation in Mendelian populations. Genetics 16, 0097–0159 (1931).CAS 
    Article 

    Google Scholar 
    Wang, J. L. Estimation of effective population sizes from data on genetic markers. Philos. Trans. R. Soc. B. Sci. 360, 1395–1409. https://doi.org/10.1098/rstb.2005.1682 (2005).CAS 
    Article 

    Google Scholar 
    Frankham, R., Bradshaw, C. J. A. & Brook, B. W. Genetics in conservation management: Revised recommendations for the 50/500 rules, Red List criteria and population viability analyses. Biol. Conserv. 170, 56–63. https://doi.org/10.1016/j.biocon.2013.12.036 (2014).Article 

    Google Scholar 
    Kennington, W. J., Hevroy, T. H. & Johnson, M. S. Long-term genetic monitoring reveals contrasting changes in the genetic composition of newly established populations of the intertidal snail Bembicium vittatum. Mol. Ecol. 21, 3489–3500. https://doi.org/10.1111/j.1365-294X.2012.05636.x (2012).Article 

    Google Scholar 
    Olson, Z. H., Whittaker, D. G. & Rhodes, O. E. Translocation history and genetic diversity in reintroduced bighorn sheep. J. Wildl. Manag. 77, 1553–1563. https://doi.org/10.1002/jwmg.624 (2013).Article 

    Google Scholar 
    White, L. C., Moseby, K. E., Thomson, V. A., Donnellan, S. C. & Austin, J. J. Long-term genetic consequences of mammal reintroductions into an Australian conservation reserve. Biol. Conserv. 219, 1–11. https://doi.org/10.1016/j.biocon.2017.12.038 (2018).Article 

    Google Scholar 
    Di Fonzo, M. M. I., Pelletier, F., Clutton-Brock, T. H., Pemberton, J. M. & Coulson, T. The population growth consequences of variation in individual heterozygosity. PLoS ONE 6, e19667. https://doi.org/10.1371/J.pone.0019667 (2011).ADS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Foerster, K., Delhey, K., Johnsen, A., Lifjeld, J. T. & Kempenaers, B. Females increase offspring heterozygosity and fitness through extra-pair matings. Nature 425, 714–717. https://doi.org/10.1038/nature01969 (2003).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    Pujolar, J. M., Maes, G. E., Vancoillie, C. & Volckaert, F. A. M. Growth rate correlates to individual heterozygosity in the european eel, Anguilla anguilla L.. Evolution 59, 189–199 (2005).CAS 
    PubMed 

    Google Scholar 
    Wolfe, K. M., Robertson, H. & Bencini, R. The mating behaviour of the dibbler, Parantechinus apicalis, in captivity. Aust. J. Zool. 48, 541–550. https://doi.org/10.1071/zo00030 (2000).Article 

    Google Scholar 
    Hedrick, P. W., Robinson, J. A., Peterson, R. O. & Vucetich, J. A. Genetics and extinction and the example of Isle Royale wolves. Anim. Conserv. 22, 302–309. https://doi.org/10.1111/acv.12479 (2019).Article 

    Google Scholar 
    Huisman, J., Kruuk, L. E. B., Ellis, P. A., Clutton-Brock, T. & Pemberton, J. M. Inbreeding depression across the lifespan in a wild mammal population. Proc. Natl. Acad. Sci. USA 113, 3585–3590. https://doi.org/10.1073/pnas.1518046113 (2016).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Nunziata, S. O. & Weisrock, D. W. Estimation of contemporary effective population size and population declines using RAD sequence data. Heredity 120, 196–207. https://doi.org/10.1038/s41437-017-0037-y (2018).CAS 
    Article 
    PubMed 

    Google Scholar 
    Popa-Baez, A. D. et al. Genome-wide patterns of differentiation over space and time in the Queensland fruit fly. Sci. Rep. 10, 13. https://doi.org/10.1038/s41598-020-67397-5 (2020).CAS 
    Article 

    Google Scholar 
    Lacy, R. C. Importance of genetic variation to the viability of mammalian populations. J. Mammal. 78, 320–335. https://doi.org/10.2307/1382885 (1997).Article 

    Google Scholar 
    Lavery, T. H., Fisher, D. O., Flannery, T. F. & Leung, L. K. P. Higher extinction rates of dasyurids on Australo-Papuan continental shelf islands and the zoogeography of New Guinea mammals. J. Biogeogr. 40, 747–758. https://doi.org/10.1111/jbi.12072 (2013).Article 

    Google Scholar 
    Sigg, D. P. Reduced genetic diversity and significant genetic differentiation after translocation: Comparison of the remnant and translocated populations of bridled nailtail wallabies (Onychogalea fraenata). Conserv. Genet. 7, 577–589. https://doi.org/10.1007/s10592-005-9096-3 (2006).Article 

    Google Scholar 
    Burgman, M. A., Akcakaya, H. R. & Loew, S. S. The use of extinction models for species conservation. Biol. Conserv. 43, 9–25. https://doi.org/10.1016/0006-3207(88)90075-4 (1988).Article 

    Google Scholar 
    Frankham, R. Inbreeding and extinction: Island populations. Conserv. Biol. 12, 665–675. https://doi.org/10.1046/j.1523-1739.1998.96456.x (1998).Article 

    Google Scholar 
    Promislow, D. E. L. & Harvey, P. H. Living fast and dying young—A comparative-analysis of life-history variation among mammals. J. Zool. 220, 417–437. https://doi.org/10.1111/j.1469-7998.1990.tb04316.x (1990).Article 

    Google Scholar 
    CSIRO. State of the Climate 2018 https://www.csiro.au/en/Showcase/state-of-the-climate/ (2018).Harris, R. M. B. et al. Biological responses to the press and pulse of climate trends and extreme events. Nat. Clim. Change 8, 579–587. https://doi.org/10.1038/s41558-018-0187-9 (2018).ADS 
    Article 

    Google Scholar 
    Morita, K. & Yokota, A. Population viability of stream-resident salmonids after habitat fragmentation: A case study with white-spotted charr (Salvelinus leucomaenis) by an individual based model. Ecol. Model. 155, 85–94. https://doi.org/10.1016/s0304-3800(02)00128-x (2002).Article 

    Google Scholar 
    Ottewell, K. et al. Evaluating success of translocations in maintaining genetic diversity in a threatened mammal. Biol. Conserv. 171, 209–219. https://doi.org/10.1016/j.biocon.2014.01.012 (2014).Article 

    Google Scholar 
    Zeoli, L. F., Sayler, R. D. & Wielgus, R. Population viability analysis for captive breeding and reintroduction of the endangered Columbia basin pygmy rabbit. Anim. Conserv. 11, 504–512. https://doi.org/10.1111/j.1469-1795.2008.00208.x (2008).Article 

    Google Scholar 
    Mella, V. S. A., McArthur, C., Krockenberger, M. B., Frend, R. & Crowther, M. S. Needing a drink: Rainfall and temperature drive the use of free water by a threatened arboreal folivore. PLoS ONE 14, e0216964. https://doi.org/10.1371/journal.pone.0216964 (2019).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Smith, A. G. et al. Out on a limb: Habitat use of a specialist folivore, the koala, at the edge of its range in a modified semi-arid landscape. Landsc. Ecol. 28, 415–426. https://doi.org/10.1007/s10980-013-9846-4 (2013).Article 

    Google Scholar 
    Akesson, M. et al. Genetic rescue in a severely inbred wolf population. Mol. Ecol. 25, 4745–4756. https://doi.org/10.1111/mec.13797 (2016).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Heber, S. et al. The genetic rescue of two bottlenecked South Island robin populations using translocations of inbred donors. Proc. R. Soc. B. Sci. 280, 20122228. https://doi.org/10.1098/rspb.2012.2228 (2013).CAS 
    Article 

    Google Scholar 
    Hedrick, P. W. & Fredrickson, R. Genetic rescue guidelines with examples from Mexican wolves and Florida panthers. Conserv. Genet. 11, 615–626. https://doi.org/10.1007/s10592-009-9999-5 (2010).Article 

    Google Scholar 
    Weeks, A. R. et al. Genetic rescue increases fitness and aids rapid recovery of an endangered marsupial population. Nat. Commun. 8, 1071. https://doi.org/10.1038/s41467-017-01182-3 (2017).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Bell, D. A. et al. The exciting potential and remaining uncertainties of genetic rescue. Trends Ecol. Evol. 34, 1070–1079. https://doi.org/10.1016/j.tree.2019.06.006 (2019).Article 
    PubMed 

    Google Scholar 
    Ralls, K., Sunnucks, P., Lacy, R. C. & Frankham, R. Genetic rescue: A critique of the evidence supports maximizing genetic diversity rather than minimizing the introduction of putatively harmful genetic variation. Biol. Conserv. 251, 8. https://doi.org/10.1016/j.biocon.2020.108784 (2020).Article 

    Google Scholar 
    Ramsey, J., Bradshaw, H. D. & Schemske, D. W. Components of reproductive isolation between the monkeyflowers Mimulus lewisii and M. cardinalis (Phrymaceae). Evolution 57, 1520–1534 (2003).Article 

    Google Scholar 
    Skoracka, A. Reproductive barriers between populations of the cereal rust mite Abacarus hystrix confirm their host specialization. Evol. Ecol. 22, 607–616. https://doi.org/10.1007/s10682-007-9185-5 (2008).Article 

    Google Scholar 
    Vines, T. H. & Schluter, D. Strong assortative mating between allopatric sticklebacks as a by-product of adaptation to different environments. Proc. R. Soc. B. Sci. 273, 911–916. https://doi.org/10.1098/rspb.2005.3387 (2006).Article 

    Google Scholar 
    Keighery, G. J., Alford, J. J. & Longman, V. A vegetation survey of the islands of the Turquoise Coast from Dongara to Lancelin, south-western Australia. Conserv. Sci. West. Aust. 4, 13–62 (2002).
    Google Scholar  More

  • in

    Multi-objective optimization can balance trade-offs among boreal caribou, biodiversity, and climate change objectives when conservation hotspots do not overlap

    IPCC. Summary for policymakers in Climate Change 2021: The Physical Science Basis. Contribution of Working Group 1 to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change (eds. Masson-Delmotte, V. et al.) 3–32 (Cambridge University Press, 2021).Ceballos, G. et al. Accelerated modern human-induced species losses: entering the sixth mass extinction. Sci. Adv. 1, e1400253. https://doi.org/10.1126/sciadv.1400253 (2015).ADS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Barnosky, A. D. et al. Has the Earth’s sixth mass extinction already arrived?. Nature 471, 51–57 (2011).ADS 
    CAS 
    Article 

    Google Scholar 
    IPBES. Global assessment report of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services (IPBES Secretariat, 2019).United Nations. What is the United Nations Framework Convention on Climate Change? https://unfccc.int/process-and-meetings/the-convention/what-is-the-united-nations-framework-convention-on-climate-change (2021).United Nations. The Paris Agreement. https://unfccc.int/process-and-meetings/the-paris-agreement/the-paris-agreement (2022).United Nations. The Convention on Biological Diversity. https://www.cbd.int/convention/ (2021).UN environment programme. Aichi Target 11, Convention on Biological Diversity https://www.cbd.int/aichi-targets/target/11 (2021).Tagesson, T. et al. Recent divergence in the contributions of tropical and boreal forests to the terrestrial carbon sink. Nat. Ecol. Evol. 4, 202–209 (2020).Article 

    Google Scholar 
    Wells, J. V., Dawson, N., Culver, N., Reid, F. A. & Morgan Siegers, S. The state of conservation in North America’s boreal forest: issues and opportunities. Front. For. Glob. Chang. 3, 90 (2020).Article 

    Google Scholar 
    Bradshaw, C. J. A. & Warkentin, I. G. Global estimates of boreal forest carbon stocks and flux. Glob. Planet. Change 128, 24–30 (2015).ADS 
    Article 

    Google Scholar 
    Drever, C. R. et al. Natural climate solutions for Canada. Sci. Adv. 7, eabd6034. https://doi.org/10.1126/sciadv.abd6034 (2021).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Government of Canada. Species at Risk Act (S.C. 2002, c. 29) https://laws.justice.gc.ca/eng/acts/S-15.3/ (2021).SARA registry. Woodland caribou (Rangifer tarandus), boreal population: species summary. https://species-registry.canada.ca/index-en.html#/species/636-252 (2022).Brandt, J. P. The extent of the North American boreal zone. Environ. Rev. 17, 101–161 (2009).Article 

    Google Scholar 
    Environment and Climate Change Canada. Boreal caribou ranges – Canada https://open.canada.ca/data/en/dataset/4eb3e825-5b0f-45a3-8b8b-355188d24b71 (2016).Hebblewhite, M. Billion dollar boreal woodland caribou and the biodiversity impacts of the global oil and gas industry. Biol. Cons. 206, 102–111 (2017).Article 

    Google Scholar 
    Hebblewhite, M. & Fortin, D. Canada fails to protect its caribou. Science 358, 730 (2017).ADS 
    CAS 
    Article 

    Google Scholar 
    Boan, J. J., Malcolm, J. R., Vanier, M. D., Euler, D. L. & Moola, F. M. From climate to caribou: how manufactured uncertainty is affecting wildlife management. Wildl. Soc. Bull. 42, 366–381 (2018).Article 

    Google Scholar 
    Government of Canada. Overview of the Pan-Canadian approach to transforming species at risk conservation in Canada https://www.canada.ca/en/services/environment/wildlife-plants-species/species-risk/pan-canadian-approach.html (2020).Environment and Climate Change Canada. Pan-Canadian approach to transforming species at risk conservation in Canada (Environment and Climate Change Canada, 2018).Assembly of First Nations & David Suzuki Foundation. Cultural and ecological value of Boreal Woodland Caribou habitat https://davidsuzuki.org/science-learning-centre-article/cultural-ecological-value-boreal-woodland-caribou-habitat/ (2013).Royal Canadian Mint. A familiar face – the 25-cent coin. https://www.mint.ca/en/discover/canadian-circulation/25-cents (2022).Drever, C. R. et al. Conservation through co-occurrence: woodland caribou as a focal species for boreal biodiversity. Biol. Conserv. 232, 238–252 (2019).Article 

    Google Scholar 
    Johnson, C. A., Drever, C. R., Kirby, P., Neave, E. & Martin, A. E. Protecting boreal caribou habitat can help conserve biodiversity and safeguard large quantities of soil carbon in Canada. Sci. Rep. (in review).Government of Canada. Canadian Protected and Conserved Areas Database https://www.canada.ca/en/environment-climate-change/services/national-wildlife-areas/protected-conserved-areas-database.html (2022).Trudeau, J. Minister of Environment and Climate Change mandate letter https://pm.gc.ca/en/mandate-letters/2021/12/16/minister-environment-and-climate-change-mandate-letter (2021).Environment Canada. Scientific assessment to inform the identification of critical habitat for Woodland Caribou (Rangifer tarandus caribou), boreal population, in Canada: 2011 update (Environment Canada, 2011).Environment and Climate Change Canada. Amended recovery strategy of the Woodland Caribou (Rangifer tarandus caribou), boreal population, in Canada. Species at Risk Act Recovery Strategy Series (Environment and Climate Change Canada, 2020).Johnson, C. A. et al. Science to inform policy: linking population dynamics to habitat for a threatened species in Canada. J. Appl. Ecol. 57, 1314–1327 (2020).Article 

    Google Scholar 
    Mansuy, N. et al. Contrasting human influences and macro-environmental factors on fire activity inside and outside protected areas of North America. Environ. Res. Lett. 14, 064007. https://doi.org/10.1088/1748-9326/ab1bc5 (2019).ADS 
    Article 

    Google Scholar 
    Mitchell, M. G. E. et al. Identifying key ecosystem service providing areas to inform national-scale conservation planning. Environ. Res. Lett. 16, 014038. https://doi.org/10.1088/1748-9326/abc121 (2021).ADS 
    CAS 
    Article 

    Google Scholar 
    Kocsis, Á. T., Zhao, Q., Costello, M. J. & Kiessling, W. Not all biodiversity rich spots are climate refugia. Biogeosciences 18, 6567–6579 (2021).ADS 
    Article 

    Google Scholar 
    Barr, S. L., Larson, B. M. H., Beechey, T. J. & Scott, D. J. Assessing climate change adaptation progress in Canada’s protected areas. Can. Geogr. 65, 152–165 (2021).Article 

    Google Scholar 
    Groves, C. R. et al. Incorporating climate change into systematic conservation planning. Biodivers. Conserv. 21, 1651–1671 (2012).Article 

    Google Scholar 
    Reside, A. E., Butt, N. & Adams, V. M. Adapting systematic conservation planning for climate change. Biodivers. Conserv. 27, 1–29 (2018).Article 

    Google Scholar 
    Sothe, C. et al. Large soil carbon storage in terrestrial ecosystems of Canada. Global Biogeochem. Cycles 36, e2021GB007213. https://doi.org/10.1029/2021GB007213 (2022).ADS 
    CAS 
    Article 

    Google Scholar 
    Beyer, H. L., Dujardin, Y., Watts, M. E. & Possingham, H. P. Solving conservation planning problems with integer linear programming. Ecol. Modell. 328, 14–22 (2016).Article 

    Google Scholar 
    Hanson, J. O., Schuster, R., Strimas-Mackey, M. & Bennett, J. R. Optimality in prioritizing conservation projects. Methods Ecol. Evol. 10, 1655–1663 (2019).Article 

    Google Scholar 
    Schuster, R., Hanson, J. O., Strimas-Mackey, M. & Bennett, J. R. Exact integer linear programming solvers outperform simulated annealing for solving conservation planning problems. PeerJ 8, e9258. https://doi.org/10.7717/peerj.9258 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    McIntosh, E. J. et al. Absence of evidence for the conservation outcomes of systematic conservation planning around the globe: a systematic map. Environ. Evid. 7, 22. https://doi.org/10.1186/s13750-018-0134-2 (2018).Article 

    Google Scholar 
    Díaz-Yáñez, O., Pukkala, T., Packalen, P., Lexer, M. J. & Peltola, H. Multi-objective forestry increases the production of ecosystem services. For. Int. J. For. Res. 94, 386–394 (2021).
    Google Scholar 
    Coristine, L. E. et al. Informing Canada’s commitment to biodiversity conservation: a science-based framework to help guide protected areas designation through Target 1 and beyond. Facets 3, 531–562 (2018).Article 

    Google Scholar 
    Carroll, C. & Ray, J. C. Maximizing the effectiveness of national commitments to protected area expansion for conserving biodiversity and ecosystem carbon under climate change. Glob. Chang. Biol. 27, 3395–3414 (2021).Article 

    Google Scholar 
    Indigenous Circle of Experts. We rise together: achieving Pathway to Canada Target 1 through the creation of Indigenous Protected and Conserved Areas in the spirit and practice of reconciliation. (2018).Zurba, M., Beazley, K. F., English, E. & Buchmann-Duck, J. Indigenous Protected and Conserved Areas (IPCAs), Aichi Target 11 and Canada’s Pathway to Target 1: focusing conservation on reconciliation. Land 8, 10. https://doi.org/10.3390/land8010010 (2019).Article 

    Google Scholar 
    Schuster, R., Germain, R. R., Bennett, J. R., Reo, N. J. & Arcese, P. Vertebrate biodiversity on indigenous-managed lands in Australia, Brazil, and Canada equals that in protected areas. Environ. Sci. Policy 101, 1–6 (2019).Article 

    Google Scholar 
    Lee, P. & Boutin, S. Persistence and developmental transition of wide seismic lines in the western Boreal Plains of Canada. J. Environ. Manage. 78, 240–250 (2006).Article 

    Google Scholar 
    Ray, J. C. Defining habitat restoration for boreal caribou in the context of national recovery: a discussion paper (Wildlife Conservation Society Canada, 2014).Carwardine, J. et al. Avoiding costly conservation mistakes: the importance of defining actions and costs in spatial priority settings. PLoS ONE 3, e2586. https://doi.org/10.1371/journal.pone.0002586 (2008).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    DeCesare, N. J. et al. Estimating ungulate recruitment and growth rates using age ratios. J. Wildl. Manage. 76, 144–153 (2012).Article 

    Google Scholar 
    Cunningham, C. A., Thomas, C. D., Morecroft, M. D., Crick, H. Q. P. & Beale, C. M. The effectiveness of the protected area network of Great Britain. Biol. Conserv. 257, 109146. https://doi.org/10.1016/j.biocon.2021.109146 (2021).Article 

    Google Scholar 
    Olds, A. D., Connolly, R. M., Pitt, K. A. & Maxwell, P. S. Habitat connectivity improves reserve performance. Conserv. Lett. 5, 56–63 (2012).Article 

    Google Scholar 
    Gurd, D. B., Nudds, T. D. & Rivard, D. H. Conservation of mammals in eastern North American wildlife reserves: how small is too small? Conserv. Biol. 15, 1355–1363 (2001).Article 

    Google Scholar 
    Government of Canada. Canadian Protected and Conserved Areas Database, December 2019 CPCAD data https://www.canada.ca/en/environment-climate-change/services/national-wildlife-areas/protected-conserved-areas-database.html (2019).Environment Canada. Recovery strategy for the woodland caribou (Rangifer tarandus caribou), boreal population, in Canada. Species at Risk Act Recovery Strategy Series (Environment Canada, 2012).R Core Team. R: A language and environment for statistical computing. Version 4.0.4 (The R Foundation, 2021).Chung, N. C., Miasojedow, B., Startek, M. & Gambin, A. Jaccard/Tanimoto similarity test and estimation methods for biological presence-absence data. BMC Bioinf. 20, 644. https://doi.org/10.1186/s12859-019-3118-5 (2019).Article 

    Google Scholar 
    Chung, N. C., Miasojedow, B., Startek, M. & Gambin, A. jaccard: test similarity between binary data using Jaccard/Tanimoto coefficients. R package version 0.1.0. https://cran.r-project.org/package=jaccard (2018).Ralphs, T., Ladanyi, L., Guzelsoy, M. & Mahajan, A. Symphony. Zenodo https://doi.org/10.5281/zenodo.2576603/ (2019).Theußl, S., Schwendinger, F. & Hornik, K. ROI: an extensible R optimization infrastructure. J. Stat. Softw. 94, 1–64 (2020).Article 

    Google Scholar 
    Theussl, S. ROI.plugin.symphony: ‘SYMPHONY’ plug-in for the ‘R’ optimization interface. R package version 1.0–0 https://CRAN.R-project.org/package=ROI.plugin.symphony (2020).Environment and Climate Change Canada. 2015 – Anthropogenic disturbance footprint within boreal caribou ranges across Canada – as interpreted from 2015 Landsat satellite imagery https://open.canada.ca/data/en/dataset/a71ab99c-6756-4e56-9d2e-2a63246a5e94 (2019).Stralberg, D. Velocity-based macrorefugia for North American ecoregions. Zenodo https://doi.org/10.5281/zenodo.2579337 (2019).Hengl, T. et al. SoilGrids250m: global gridded soil information based on machine learning. PLoS ONE 12, e0169748. https://doi.org/10.1371/journal.pone.0169748 (2017).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar  More

  • in

    Amplified warming from physiological responses to carbon dioxide reduces the potential of vegetation for climate change mitigation

    Global vegetation physiological response to increasing atmospheric CO2 and its reduction of mitigation potentialWe calculate different effects of increasing CO2 on mean annual near-surface air temperature change over global vegetated land. We compare the direct atmospheric radiative effect (RAD)-induced climate warming to the temperature reductions caused by the BGC effect. These two global temperature changes are, in turn, then compared against our main focus of aggregated local PHY-induced temperature contributions (PHYall; Fig.1). The main finding is that the spatial aggregation of PHY feedbacks on temperature over global vegetated land (green bars) offsets a substantial amount of the cooling effect through enhanced terrestrial carbon storage because of the BGC effects (blue bars). Terrestrial carbon storage continuously increases with rising atmospheric CO2, and reaches a global total of 621 ± 260 Pg C under 4 × CO2 (Supplementary Fig. 1). This increased land carbon storage is equivalent to 293 ± 122 ppm of CO2 removal from the atmosphere and results in a temperature cooling of −1.24 ± 0.57 °C. The PHYall-induced land temperature increase is 0.83 ± 0.47 °C, from the ensemble of five ESMs we use (here we excluded bcc-csm1-1; see “Methods”), corresponding to a large offset of the cooling effect by terrestrial ecosystems through BGC. However, our estimated temperature cooling induced by the BGC effect is a transient response. This cooling effect will be larger for subsequently stabilised CO2 concentrations, since terrestrial ecosystems continue to fix carbon until reaching equilibrium. We note that BGC-induced cooling may be overestimated in the absence of land-cover change effect in the simulations, as the latter may reduce terrestrial carbon stores (Supplementary Fig. 2). However, inter-model differences, such as different parameterisation or different biogeochemistry module20, may prevent a definitive answer as to how land-cover change influence global temperature change through BGC (Supplementary Fig. 2–5). We also note that we focus our analysis on CMIP5 data mainly due to the fact that the parameters in Eqs. (1) and (2)21,22 (see “Methods”) for calculating BGC-induced cooling are only currently available for CMIP5 ESMs.Fig. 1: Climate warming mitigation potential of terrestrial ecosystems.Global mean land temperature change due to total CO2 physiological forcing (PHYall), increased terrestrial carbon storage (BGC) and CO2 radiative forcing (RAD). Note axes as coloured, and that the vertical blue axis is temperature cooling through BGC. It is straightforward to compare PHYall-, RAD- and BGC-induced temperature change when the same range and directions of bars are used. Three atmospheric CO2 horizons are selected here, for CO2 concentrations of 500 ppm, 800 ppm and 1032 ppm (4 × CO2). Each bar represents the global area-weighted average from the ensemble of five ESMs. The error bars indicate the standard error of these five models. Note, the bcc-csm1-1 model did not provide diagnostics of carbon storage data and so is not included in calculating the temperature change by carbon storage change.Full size imageThe RAD response to rising CO2 projects a global warming of 5.25 ± 0.65 °C, again under 4 × CO2, corresponding to roughly four times the magnitude of BGC-induced cooling. Hence, PHYall-induced temperature increase adds about 16 ± 8% to the RAD-induced warming globally. The relative magnitude of warming/cooling effects is similar for the lower CO2 levels of 500 ppm and 800 ppm (Fig. 1), illustrating the importance of accounting for PHYall-induced warming and how it affects the ability of terrestrial vegetation to mitigate global warming, irrespective of CO2 concentration. Our identified PHYall feedbacks are a combination of different altered land properties. The balance of changed land components will depend on location, and so spatial variations in the overall warming effect may suggest a reappraisal of some climate change adaptation measures. Hence to aid such assessments, we now consider in detail the global contributions of individual drivers of the PHY-based local warming, and then any geographical variations.In Fig. 2a, we show changes in global vegetated land air temperature associated with PHYall with increasing atmospheric CO2 from our ensemble of six ESMs (see “Methods”). PHYdir (see “Methods”; Supplementary Table 1) is based on the direct CO2 effect on vegetation physiology. PHYall represents all the vegetation physiology-related feedbacks, and captures any additional interactions between RAD and PHYdir, due to all effects not being a simple linear addition of RAD and PHY responses. The difference, PHYall minus PHYdir, is termed PHYint. Although inter-model difference exists, warming levels are projected to increase, by all the ESMs, as CO2 concentration rises, and for both PHYdir and PHYall (Fig. 2a). The interactive effects result in PHYall-induced temperature change being higher than PHYdir throughout the period. For the smaller increases in atmospheric CO2 of up to ~450 ppm, interactive effects dominate with PHYdir being almost zero up to that concentration. However, above the CO2 concentration of 450 ppm, PHYdir increases global warming reaching 0.17 ± 0.06 °C for CO2 of 500 ppm, and climbing to 0.62 ± 0.48 °C under quadrupled atmospheric CO2 (Fig. 2a). At that 4 × CO2 level, interaction term PHYint increases warming by approximately one-fifth of that induced by PHYdir. Hence, the overall physiological feedbacks described by PHYall produce a global temperature increase of 0.74 ± 0.47 °C under 4 × CO2 (again from multi-model mean of six ESMs; see “Methods”).Fig. 2: Change in global mean annual land air temperature and individual climate forcing induced by vegetation physiological response to increasing atmospheric CO2.a Global annual area-weighted temperature change of vegetated land induced by total CO2 physiological forcing (PHYall) and the direct CO2 physiological forcing (PHYdir) in response to increasing atmospheric CO2 concentration. Shaded areas are the standard errors of the six Earth System Models (ESMs) used, and the thick curves are their multi-model means. For each atmospheric CO2 concentration in panel (a), values are based on smoothing using a twenty-year running window (to match with the decomposition results in panel (b). The final temperature change induced by PHYall and PHYdir effects under 4 × CO2 are further marked on the righthand side, with the “+” markers indicating multi-model means. b PHYall-induced climate forcing associated with changes in albedo, aerodynamic resistance (ra), evapotranspiration (ET), downwelling shortwave radiation (SW) and near-surface air emissivity (ɛa). Again, the shaded areas are the standard errors of the models, and the mean is the thick continuous lines. The changes in these variables are calculated using a moving average with a 20-year window. The resulting values under 4 × CO2 are plotted on the righthand side.Full size imageTo better understand the factors influencing CO2 physiological drivers of temperature change, we decompose the global PHYall into individual biophysical components6 (see “Methods”). These five aspects are albedo, aerodynamic resistance (ra), evapotranspiration (ET), downwelling shortwave radiation (SW) and near-surface air emissivity (ɛa). These five-component changes result in climate forcings with different signs and magnitudes (Fig. 2b), and thus perturb the surface energy balance, where positive values correspond to an increase in temperature. Specifically, ET, SW, and albedo generates positive climate forcings that increase local temperature, while ra and ɛa produce negative effects and thus offset local temperature increase. The relative role of each biophysical component in influencing PHYall-induced temperature change remains largely invariant in the transition from low to high CO2 concentrations. In addition, the changes in these five quantities affecting PHYdir show similar variations with that of PHYall as atmospheric CO2 rises (Supplementary Fig. 6). Those results are generally valid for CMIP6 results but with a lower magnitude of change (Supplementary Fig. 7).The vegetation physiological response to rising CO2 causes changes in LAI and stomata closure, which adjust in parallel with more detailed attributes of the land surface. The ESM simulations of changes in LAI and transpiration compare moderately well with available field measurements23,24,25,26,27,28,29,30,31,32 (Supplementary Fig. 8). Here we focus on their effects on the near-surface thermal changes, expressed as climate forcings on near-surface energy fluxes (Fig. 2b). The LAI increase due to elevated CO2 (Supplementary Fig. 9a) leads to decreases in albedo11,33 (Supplementary Fig. 10a; Supplementary Table 2) and ra7 (Supplementary Fig. 10c). These changes are continuous with rising atmospheric CO2 and result in positive and negative effects on global land air temperature change, respectively (Fig. 2b). Specifically, the albedo reduction increases solar radiation absorption by the land surface11,33, and imposes a positive forcing of 0.30 ± 0.14 W m−2 for 500 ppm, and of 0.51 ± 0.45 W m−2 for a quadrupling of CO2 (Fig. 2b). The decreased ra favours the turbulent transport of heat from land to atmosphere7, and leads to a persistent surface cooling with increasing atmospheric CO234,35, which is −1.66 ± 0.44 W m−2 for 500 ppm and −3.15 ± 2.05 W m−2 for 4 × CO2. In contrast, ET decreases considerably (Supplementary Fig. 10e; Supplementary Table 2) due to decreased stomatal conductance responding to increasing atmospheric CO2. These reductions in ET reduce evaporative cooling15,17, and therefore result in a strong positive climate forcing on global warming (Fig. 2b), which increases with rising atmospheric CO2 and reaches 0.88 ± 0.25 W m−2 for atmospheric CO2 of 500 ppm and 2.88 ± 1.49 W m−2 with quadrupled atmospheric CO2 concentration (Fig. 2b).Moreover, ET reduction decreases the inflow of evaporative water to the atmosphere, reducing cloud fraction and water vapour content, thereby feeding back to impose indirect effects that influence temperature change. We quantify these indirect effects through their effects on the components of SW and ɛa, (Fig. 2b). The lower ET values reduce cloud fraction (Supplementary Table 2; Supplementary Fig. 9c) thereby increasing the amount of SW to the land surface6,35,36 (Supplementary Fig. 10g), and producing a moderate positive forcing of 0.67 ± 0.16 W m−2 for 500 ppm and 1.76 ± 1.26 W m−2 under 4 × CO2 (Fig. 2b). However, the reduced cloud fraction and water vapour content decreases ɛa (Supplementary Fig. 10i), which weakens the absorption of longwave radiation37 (Supplementary Fig. 9e). The ɛa changes result in a small negative forcing (−0.28 ± 0.11 W m−2 for 500 ppm and −0.57 ± 0.43 W m−2 for 4 × CO2), and thereby lowering global temperatures. Of particular note is that the direct changes in ET produce the strongest positive forcing for warming (Fig. 2b). Hence, ET changes have a dominant role in influencing the magnitude and sign of PHYall feedbacks on global temperature change. The changes in these five factors for PHYdir are very close in relative terms to those of PHYall (Supplementary Fig. 10), driving climate forcings in a similar magnitude and sign to influence PHYdir-induced global temperature change (Supplementary Fig. 6).Overall, as CO2 rises, these five biophysical factors cause vegetation to amplify warming locally, and when aggregated spatially act to raise planetary global warming. This additional warming effect is primarily driven by changes in ET, with smaller warming contributions from changes in SW and albedo, but also compensated with cooling effects from ra and ɛa changes. The substantial role of ET on biophysical climate feedbacks is consistent with a previous study6, which investigates the biophysical feedbacks because of vegetation greening. That study proposed that CO2-driven greening will enhance ET and thereby producing a net cooling effect. We build on that analysis by here additionally including the CO2-induced partial stomatal closure17,18. We find this inclusion overtakes the LAI influence on ET changes, resulting in large reductions in ET that will instead contribute to net warming as CO2 rises. Large inter-model differences in simulating ra dynamics (Supplementary Fig. 10c, d) mostly explains the substantial spread of ra-induced climate forcing (Fig. 2b). The different extents of LAI increase (Supplementary Fig. 9a, b) also contribute to such differences in simulated ra changes and effects on temperature change. Our noted large cooling through ra changes is also indicated in a recent study7, suggesting the importance of ra in affecting vegetation biophysical climate feedbacks, and the importance of constraining this factor in climate models.Spatial patterns and attributions of vegetation physiologically induced warmingWe present in Fig. 3 the area-weighted regional contributions to global temperature change of the physiological responses and BGC under 4 × CO2. PHYall-based warming and BGC-induced cooling both show larger values in East and Central North America (ENA and CNA), North and Central Europe (NEU and CEU), Amazon (AMZ) and North Asia (NAS). Whereas, the smallest changes are in the Sahel region (SAH) (Fig. 3a). PHYall-induced warming reduces large proportions of the temperature cooling through BGC in the northern mid-to-high latitudes (Fig. 3b). Furthermore, this cooling effect is fully offset by warming through vegetation physiological response in Alaska (ALA), Canada/Greenland/Iceland (CGI), NAS, NEU, SAH, Tibetan Plateau (TIB), Central (CAS) and West Asia (WAS), and West North America (WNA), resulting in slight warming in these regions. Presented as a map of the net effects of BGC and PHYall (Fig. 3b), we further illustrate this overall cooling effect from PHY and BGC for the tropical regions and South Hemisphere. Additionally, the balance of PHY and BGC to influence regional temperature change under relatively low atmospheric CO2 level of 500 ppm (Supplementary Fig. 11) is very close to that for 4 × CO2, suggesting consistency of vegetation biophysical and biogeochemical effects on climate irrespective of CO2 levels. In summary, the PHYall feedbacks offset the cooling benefits from ecosystem “draw-down” of CO2 to a large extent, in line with the global average values presented in Fig. 1. Northern mid-to-high latitudes may contribute less than expected to slow global warming (Fig. 3). However, as also noted for the global change values, the estimated temperature cooling by BGC is a transient effect that would keep increasing as the ecosystems approach their equilibrium. However, BGC-induced cooling may be overestimated without consideration of the effect of land-cover change (Supplementary Fig. 2), which is often associated with the deliberate removal of terrestrial carbon.Fig. 3: Regional contributions to temperature change by PHY and BGC under 4 × CO2.a Contribution of regional vegetation physiological responses (PHYall; green bars) and increased carbon storage (BGC; blue bars) to the overall global temperature change for each of the IPCC AR5 SREX regions. Bars represent area-weighted multi-model means, and the error bars indicate the standard errors of the models for each region. b Spatial distribution of the net effects of warming induced by PHYall and cooling through BGC.Full size imageThe regional pattern in Fig. 3b provides a motivation to investigate further the five climate forcings-driven PHYall contributions (Fig. 2a) to near-surface temperature change triggered by rising CO2. Here, we first examine the much finer spatial patterns and component contributions of PHYall, PHYdir and PHYint on warming for 4 × CO2. Focussing on PHYdir and typically for 4 × CO2, there is a larger local warming in the tropical forests (Fig. 4a), where vegetation shows higher potential to stabilise carbon than other ecosystems10 (Fig. 3). Within these tropical regions, the PHYdir-forced temperature increase is the highest in the Amazon forests (Fig. 4a), reaching approximately 30% of that induced by RAD (Supplementary Fig. 12b). Strong warming by PHYdir is also found in the northern mid-to-high latitudes ( >40°N). In contrast, smaller temperature increases due to PHYdir are seen in arid and semi-arid regions, such as Australia and Sahel (Fig. 4a). The spatial variations of PHYall-forced temperature change (Supplementary Fig. 13a) have strong similarities to those of PHYdir (Fig. 4a). These similarities again suggest a relatively small role of interactions on temperature change (Supplementary Fig. 13b) under quadrupled CO2. Additionally, the agreements across the six ESMs are reasonably high, with all the models agreeing that PHYall, PHYdir and PHYint result in local warming for 4 × CO2 across most of the global vegetated land (Supplementary Fig. 14).Fig. 4: Global patterns of local temperature change and climate forcings through vegetation physiological response to 4 × CO2.Spatial distribution of annual mean temperature change from multi-model ensemble induced by (a). direct CO2 physiological forcing (PHYdir) in response to a 4 × CO2 rise since pre-industrial level. The spatial patterns of the individual climate forcing contributing to PHYdir of panel (a) are as follows. In (b). albedo (α), c Aerodynamic resistance (ra), d Evapotranspiration (ET), e Downwelling shortwave radiation (SW) and (f). near-surface air emissivity (ɛa). In all panels, estimates use the mean of the final 20 years of the simulations (atmospheric CO2 at ~1032 ppm).Full size imageWe next analyse spatially the decomposition of PHYall- and PHYdir-induced temperature change into the five biophysical factors shown in Fig. 2b, and with findings shown in Fig. 4b–f. In response to quadrupled CO2, LAI shows increases over most global vegetated land (Supplementary Fig. 15b), leading to a positive climate forcing (and thus warming) through reducing albedo (Fig. 4b), and especially in the south Sahel and Tibetan Plateau. Moreover, a large albedo decrease occurs in the northern mid-to-high latitudes, such as for most of Siberia. This albedo decrease may be due to LAI increase combined with reduced snow cover promoted by PHY-driven warming17. The LAI increase also contributes to a strong negative forcing through ra decrease7. This ra decrease enhances the energy exchange between land and atmosphere, lowering the local warming effect in response to 4 × CO2, particularly in Australia, Sahel, South Africa and South Asia (Fig. 4c). In comparison, the projected forcing from ET reductions (Supplementary Fig. 15d) causes strongly positive warming, and especially in the tropical and boreal forests (Fig. 4d). The larger ET reductions in the tropical and boreal forests also lead to stronger cloud fraction decreases (Supplementary Fig. 15f). These feedbacks induce a positive climate forcing for additional warming by increasing SW reaching the land surface (Fig. 4e). Simultaneously, the decreased cloud cover and water vapour by the ET reductions generate a cooling effect (Fig. 4f) through decreasing net longwave radiation absorption in most locations. The PHYdir-induced warming and the associated climate forcings at an atmospheric CO2 concentration of 500 ppm (Supplementary Fig. 16) show similar spatial distributions, but smaller magnitude, compared with that adjustment for 4 × CO2 increase (Fig. 4). This result manifests that PHY continues to amplify global warming through the combined climate forcings from our identified five biophysical factors, irrespective of CO2 concentration.Substantial differences exist between PHYall- and PHYdir-induced temperature changes in northwestern Eurasia, implying pronounced PHYint-based feedbacks there under 4 × CO2 (Supplementary Fig. 13b). A relatively large change in SW, inducing a warming effect (Supplementary Fig. 13j), mainly contributes to the large PHYint-coupled changes in this region. Elsewhere, for the Amazon forests, the cooling effect in PHYint, through changes in ET (Supplementary Fig. 13h) and SW (Supplementary Fig. 13j), cancels out the warming due to changes in ra (Supplementary Fig. 13f) and ɛa (Supplementary Fig. 13l). This cancellation causes negligible PHYint feedback on temperature there (Supplementary Fig. 13b). In summary, local warming due to biophysical feedback of increasing CO2 is regulated primarily by the positive forcing from ET reductions (with smaller positive forcings from albedo and SW changes), which is partly compensated by reductions in ra (and a small contribution from ɛa).Strong physiologically induced warming in dense ecosystemsWe further investigate the finding presented in Fig. 4 that PHYdir-forced temperature increase is higher in the tropical forests, while lower in the arid and semi-arid ecosystems. This finding suggests that there may be a relationship between the forced temperature change and background baseline LAI, which we confirm in Fig. 5a (for both PHYall and PHYdir). We find widespread warming amplification with increasing LAI for lower background LAI levels. Such rates of increase in temperature per unit of LAI show evidence of flattening at higher LAI values38. That is, the CO2-fertilised LAI increase saturates in ecosystems with dense canopy cover such as tropical forests (Supplementary Fig. 17a). However, the stomatal closure-induced ET decrease varies almost linearly with baseline LAI, although the reduction breaks down when the baseline LAI approaches six (Supplementary Fig. 17b). Hence, we conclude that ET reductions caused by stomatal closure are the primary cause of PHY-induced temperature rises, although changes in ra because of LAI increases remain an important factor (Fig. 4; Supplementary Fig. 13). Taking all the components together, PHYdir-triggered temperature change increases almost linearly with baseline LAI gradients under 4 × CO2 (Fig. 5a). Of particular interest is that when expressing PHYdir-induced temperature increase as a fraction of warming induced by RAD, it also increases with the baseline LAI value (Fig. 5b). Hence, Fig. 5 provides strong evidence that background vegetation functional structure (LAI) influences the feedback of CO2 physiological forcing on warming in response to 4 × CO2. In terms of change, as the trend of global LAI increase slows down as atmospheric CO2 concentration increases to very high levels (Supplementary Fig. 9a, b), this suggests that the magnitude of the cooling effect (via ra) will decrease relative to the ET-based warming.Fig. 5: Effects of vegetation structure on CO2 physiological forcing in response to 4 × CO2.a Variations of local temperature change induced by total (PHYall) and direct (PHYdir) CO2 physiological forcing, for quadrupled CO2 and presented as a function of baseline pre-industrial leaf area index (LAI). b Variations of the ratio of PHYall- and PHYdir-induced temperature change relative to that of RAD warming, again for 4 × CO2 and as a function of baseline LAI. In both panels, the temperature changes along LAI gradients are smoothed using a five-bin running window for the 0.1 increments in LAI bin size. The shaded areas indicate the standard error among different models.Full size imageImplications and conclusionsTerrestrial ecosystems respond to increasing atmospheric CO2 not only through accumulating carbon (the “BGC” effect) but also through their physiological response (the “PHY” feedback), which can have opposing effects on local surface temperature. As such, to fully understand the net climate benefits of terrestrial ecosystems, comprehensive assessments need to account for both PHY and BGC effects39,40. We find that the warming through PHY feedback largely reduces the capacity of vegetation to slow global warming through BGC. In particular, vegetation transiently operates to amplify local warming in the northern mid-to-high latitudes, as PHY-induced warming is larger than BGC-induced cooling (Fig. 3). Tropical forests have net cooling effects, and forests and ecosystem restoration efforts there would have much higher net cooling benefit for mitigation than if they are placed in the northern mid-to-high latitudes. At the global scale, this PHY-induced warming can offset by up to 67% of the cooling gains from the transient BGC effect (Fig. 1). This is likely an upper-bound value, because the steady-state BGC effect is larger than the transient one. Thus, afforestation can still result in a net cooling effect, especially in the tropics, although the cooling achieved may be less than first expected. Of particular interest is that the PHY-based warming is generally stronger for high baseline LAI values (Fig. 5). This result further suggests the PHY produces extra warming through afforestation due to higher LAI for forests than non-forests. However, this does not contradict the finding for the tropics (where LAI is high) as possibly the most beneficial regions for afforestation, due to the compensating much larger BGC potential to lower temperature for such locations.Our analysis highlights the importance of including PHY feedbacks in any assessments of future levels of global warming. Such understanding may be especially important for the northern mid-to-high latitudes. We point out that only one-third of ESMs we use incorporate dynamic global vegetation schemes (DGVMs). Since land-cover change acts to influence PHY and BGC effects (Supplementary Fig. 2–5), future simulations with dynamic vegetation may allow refining these results. However, we note that inter-model structural differences (e.g. different ratio of transpiration to ET15,41) contribute to large uncertainties in these results. In particular, we note the potential implications of our results for climate mitigation policies with substantial reforestation as a mechanism to slow global warming through emissions offsetting. Reforestation may cause a near-surface cooling through the BGC effect. However, the PHY-based warming effect is stronger for higher LAI values as atmospheric CO2 rises (Fig. 5a). This must be accounted for in any major global reforestation plans with consideration of particular location of reforestation measures42. In general, the effects of anthropogenic land use and land cover change on vegetation biophysical and biogeochemical processes are important and complex with substantial spatial heterogeneity42,43,44,45,46,47. Given the lack of exploitable factorial simulations designed to separate land use and land-cover change effects, these are ignored in the idealised simulations, and may introduce a bias in our results. We note, however, that these idealized “4 × CO2” simulations broadly capture the sign and spatial distributions of land-atmosphere coupling effects on temperature change through comparisons to the ESM results under the RCP8.5 scenario (Supplementary Fig. 18) which include anthropogenic land use and land-cover change. We suggest that future climate projections should much more routinely account for the effect of anthropogenic land use and land-cover change together with PHY and BGC effects42,47, and so that this can be considered in reforestation climate mitigation strategies.An additional caveat of our research is that the PHY and RAD factorial ESM simulations that inform our analysis, are at present only available for the illustrative but potentially unrealistic exponential increase in atmospheric CO2 (1% per year). For this scenario, the systems can exhibit a unrealistic linear behaviour48. PHY and RAD simulations under other scenarios, such as abrupt 4 × CO2 or even historical forcing followed by a potential future scenario, would greatly help to improve projections of PHY feedbacks. We suggest this should be made a high priority of climate modelling exercises, especially if additionally including land use and land-cover change representation. We recognise that the newer CMIP6 ESMs contain improved representation of many processes (e.g. atmospheric aerosols, clouds, and land processes)49. We anticipate our findings to remain broadly valid with the CMIP6 models, although the magnitude of PHY-induced warming is lower for CMIP6 than CMIP5 ESMs18, especially in the northern high latitudes. We hope our analysis provides an incentive to others to undertake such analysis as all the calculations (BGC, PHY, and RAD) become available for that newer CMIP6 ensemble. In summary, our results illustrate that vegetation physiological response to increasing atmospheric CO2 has a substantial local warming effect, which requires consideration alongside the cooling effect vegetation offers by “drawing down” atmospheric carbon dioxide. More

  • in

    Case study of the convergent evolution in the color patterns in the freshwater bivalves

    Remarks on the residual color patterns in the Kitadani Freshwater BivalvesResidual color patterns in the form of visible pigmentation on fossil molluscan shells are generally uncommon2,3. In the Paleozoic to Mesozoic fossil records, the color patterns were limited to marine species3, which are preserved as black to dark-colored bands running on the shell surface as melanin pigments20,21. The black to dark-colored stripes on the shells of the Kitadani Freshwater Bivalves resemble the color patterns in some extant freshwater bivalves, suggesting that the dark bands are residual color patterns remaining as melanin pigments. Consequently, the Kitadani Freshwater Bivalves represents the oldest and second fossil record of residual color patterns among fossil freshwater bivalves.The residual color patterns of the Kitadani Freshwater Bivalves resemble the color patterns of extant freshwater bivalves in terms of width, number, and distribution of the colored bands. Both the Kitadani Freshwater Bivalves and extant freshwater bivalves examined in this study consist of two types of color patterns: stripes along the growth lines and radial rays tapered toward the umbo. Notably, the former pattern is similar among all the species examined, as it forms in the peripheries of prominent growth lines occurring periodically. In the latter pattern, however, the morphology and distribution of the bands are slightly different between the Kitadani Freshwater Bivalves and the extant species. The Kitadani Freshwater Bivalves exhibits relatively distinct and wide radial rays running roughly parallel to the lengths of the sculpture elements (radial plications and/or wrinkles), while the extant species bear obscure and fine radial rays running diagonally to the lengths of the sculpture elements. Nonetheless, the taxa with V-shaped sculpture elements (wrinkles, ribs or arranged nodules) lack or bear ambiguous radial rays, whether extant (e.g., Triplodon spp., Indochinella spp. and Tritogonia spp.)13,15,22 or extinct (†Trigonioides tetoriensis).Hypothesis I: phylogenetic constraintsThe resemblance of the color patterns between the Kitadani Freshwater Bivalves and the extant unionids possibly resulted from the phylogenetic constrains. Each of the three species of the Kitadani Freshwater Bivalves belongs to a separate family (†Trigonioides tetoriensis: †Trigonioididae, †Plicatounio naktongensis: †Plicatounionidae, and †Matsuomtoina matsumotoi: †Pseudohyriidae) in the order Trigoniida19. Trigoniida in turn, forms the subclass Palaeoheterodonta with Unionida23. This raises a possibility that the color patterns observed in the Kitadani Freshwater Bivalves and the extant unionids is inherited from their most recent common ancestor. In other words, these color patterns, stripes along the growth lines and radial rays tapered toward the umbo, may be the apomorphy for Palaeoheterodonta. In fact, some extant trigoniid species belonging to Neotrigonia exhibit color pattern similar to those in the Kitadani Freshwater Bivalves and extant unionids in this study (e.g. Neotrigonia margaritacea)1.Interestingly, the coloration of color patterns is quite different between unioniids (green to blue colorings) and trigoniids (red to yellow colorings), and the oldest known color patterns of the Palaeoheterodonta (Myophorella nodulosa, a marine species of Trigoniida from the Oxfordian of the Early Jurassic) appears different (concentric rows of patches)10 from those of the Kitadani Freshwater Bivalves or the extant unioniids. These observations suggest that colorations evolved independently, in contrast to the color patterns, between Trigoniida and Unionida, and that Trigoniida more diverse color patterns than Unionida did in the Palaeoheterodont evolutionary history. Although further examination of the fossil record for the residual colors and color patterns in Palaeoheterodonta is essential, it is plausible that the habitat differences may have caused such discrepancy in the colorations and color patterns between Trigoniida (mainly marine) and Unionida (freshwater) in spite of the phylogenetic constrains.Hypothesis II: convergent evolutionThe other possible interpretation of the color pattern similarity between the Kitadani Freshwater Bivalves and extant Unionida, is the convergent evolution. One potential factor that may have caused this convergent evolution of the color patterns is an adaptation to their habitats. In general, much of the convergent evolution in animals occurs through the morphological evolution in response to their habitats24. Similarly in mollusks, shell colors and their patterns are generally influenced by their habitats2,6,25. Considering marine mollusks, the shell colors and their patterns have great diversity due to varying habitat environments, especially in coral reeves that exhibit various colors and complex ecosystem2,6. Conversely, in the freshwater ecosystem, the environmental colors are relatively monotonous with rocks, sand, mud, and green algae8, and such habitat conditions are likely indifferent between the Mesozoic and Cenozoic. As a result, the freshwater bivalves retained simple and monotonous color patterns for adapting to such environments throughout their evolution.Another conceivable factor to explain the convergent evolution in the color patterns of the studied freshwater bivalves is the selection pressure by visual predators. In general, the shell colors and their patterns in bivalves act as camouflages against the predators2,7,8,26,27,28. Previous studies have demonstrated that extant freshwater bivalves are preyed upon by crayfish, fish, birds, reptiles, and mammals29,30. Because shell colors in freshwater bivalves tend to be greenish, such colors may be an adaptation against visual predators for blending into the freshwater sediments on which abundant greenish phytoplanktons occur2,8. Therefore, the evolutionary conservatism in color patterns of freshwater bivalves may result from camouflages into freshwater microenvironments, which has been advantageous against visual predators since the late Early Cretaceous.The above discussion assumes that the visual predators of freshwater bivalves remained similar for at least 120 million years. Which animals could have been potential threads to the Kitadani Freshwater Bivalves, and, in turn, the Early Cretaceous freshwater bivalves? Among the extant visual predators of the freshwater bivalves, those whose lineages were present in the Early Cretaceous include crustaceans (especially brachyuran decapoda31), fish, lizards, turtles, crocodiles, birds, and mammals. Among them, the fossil record of durophagous lizards and mammals can be traced back only to the Late Cretaceous32,33. Conversely, lines of fossil evidence suggest that some fish34,35, turtles36, and crocodiles35 fed on molluscan invertebrates during the Early Cretaceous, and the Kitadani Freshwater Bivalves indeed occurs with abundant lepisosteiform scales, testudinate shells and crocodile teeth. Additionally, at least one Early Cretaceous avian species with crustacean gut contents can be attributed to the durophagous diet37, and the Kitadani Formation has yielded avialan skeletal remains38, and footprints39,40. Therefore, fish, turtles, crocodiles, and birds are likely candidates for visual predators of the Early Cretaceous freshwater bivalves, and have remained so until present. Additionally, while crustaceans have not been identified in the Kitadani Formation, they flourished in the Early Cretaceous and their remains occur with the fossil freshwater bivalves of the time elsewhere31. Thus, crustaceans may have also played a role as visual predators of the freshwater bivalves since the Early Cretaceous.In addition to the crustaceans, fishes, turtles, crocodiles and birds, the visual predators of the Early Cretaceous freshwater bivalves likely include extinct lineages. For example, some pliosauroid plesiosaurs are suggested as being durophagous34, although the freshwater members of the group are considered endemic41 and less likely to be a major thread to the Early Cretaceous freshwater bivalves. Another extinct candidate is non-avian dinosaurs. Ornithischians are suggested to have possessed a dietary flexibility including the durophagy. For instance, well-preserved hadrosaurid coprolites from the Late Cretaceous of Montana, U.S.A. include sizeable crustaceans and mollusks, possibly suggesting that the Cretaceous freshwater mollusks were consumed by these herbivorous dinosaurs42. In addition, some basal ceratopsian psittacosaurids are hypothesized for the durophagy based on the predicted large bite force in the caudal portion of the toothrow43. Among saurischians, some oviraptorosaurian theropods are indicated to consume mollusks with hard shells based on their mandibular features44. While hadrosaurids, psittacosaurids, and oviraptorosaurians have not been identified in the Kitadani Formation, psittacosaurids, and oviraptorosaurians are common elsewhere in the Early Cretaceous of East Asia45,46, and hadrosauroid Koshisaurus is present in the formation47. Because dinosaurs occupied a niche of large terrestrial predators throughout the Mesozoic, they may have acted as one of major mollusk consumers in absence of large lizards and mammals in the Early Cretaceous ecosystem. Thus, the predation pressure by visual predators to the freshwater bivalves in the Early Cretaceous is likely similar to that in the present. Consequently, one of evolutionary adaptations of the freshwater bivalves against such pressure has remained to camouflage in the phytoplankton-rich sediments, leading to the long-term evolutionary conservatism of their color patterns. More