More stories

  • in

    Hinfluences severe disease-mediated population declines in two of the most common garden bird species in Great Britain

    Gregory, R. D. & van Strien, A. Wild bird indicators: Using composite population trends of birds as measures of environmental health. Ornithol. Sci. 9, 3–22 (2010).Article 

    Google Scholar 
    Cox, D. T. C. & Gaston, K. J. Urban bird feeding: Connecting people with nature. PLoS ONE 11, e0158717 (2016).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Anderson, R. M. & May, R. M. Population biology of infectious diseases: Part I. Nature 280, 361–367 (1979).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Keesing, F. et al. Impacts of biodiversity on the emergence and transmission of infectious diseases. Nature 468, 647–652 (2010).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Smith, K. F., Acevedo-Whitehouse, K. & Pedersen, A. B. The role of infectious diseases in biological conservation. Anim. Conserv. 12, 1–12 (2009).Article 

    Google Scholar 
    Han, B. A., Kramer, A. M. & Drake, J. M. Global patterns of zoonotic disease in mammals. Trends Parasitol. 32, 565–577 (2016).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Estrada-Peña, A., Ostfeld, R. S., Peterson, A. T., Poulin, R. & de la Fuente, J. Effects of environmental change on zoonotic disease risk: An ecological primer. Trends Parasitol. 30, 205–214 (2014).PubMed 
    Article 

    Google Scholar 
    Daszak, P., Cunningham, A. A. & Hyatt, A. D. Emerging infectious diseases of wildlife–threats to biodiversity and human health. Science 287(5452), 443–449 (2000).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Pedersen, A. B., Jones, K. E., Nunn, C. L. & Altizer, S. Infectious diseases and extinction risk in wild mammals. Conserv. Biol. 21, 1269–1279 (2007).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Atkinson, C. T. & Samuel, M. D. Avian malaria Plasmodium relictum in native Hawaiian forest birds: Epizootiology and demographic impacts on àapapane Himatione sanguinea. J. Avian Biol. 41, 357–366 (2010).Article 

    Google Scholar 
    George, T. L. et al. Persistent impacts of West Nile virus on North American bird populations. Proc. Natl. Acad. Sci. USA. 112, 14290–14294 (2015).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Dhondt, A. A., Tessaglia, D. L. & Slothower, R. L. Epidemic mycoplasmal conjunctivitis in house finches from Eastern North America. J. Wildl. Dis. 34, 265–280 (1998).CAS 
    PubMed 
    Article 

    Google Scholar 
    Monterroso, P. et al. Disease-mediated bottom-up regulation: An emergent virus affects a keystone prey, and alters the dynamics of trophic webs. Sci. Rep. 6, 36072 (2016).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Cheng, T. L. et al. The scope and severity of white-nose syndrome on hibernating bats in North America. Conserv. Biol. 35, 1586–1597 (2021).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Rushton, S. P. et al. Disease threats posed by alien species: The role of a poxvirus in the decline of the native red squirrel in Britain. Epidemiol. Infect. 134, 521–533 (2006).CAS 
    PubMed 
    Article 

    Google Scholar 
    Scheele, B. C. et al. Amphibian fungal panzootic causes catastrophic and ongoing loss of biodiversity. Science 363(6434), 1459–1463 (2019).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Bradley, C. A. & Altizer, S. Urbanization and the ecology of wildlife diseases. Trends Ecol. Evol. 22, 95–102 (2007).PubMed 
    Article 

    Google Scholar 
    Murray, M. H. et al. City sicker? A meta-analysis of wildlife health and urbanization. Front. Ecol. Environ. 17, 575–583 (2019).Article 

    Google Scholar 
    Giraudeau, M., Mousel, M., Earl, S. & McGraw, K. Parasites in the city: Degree of urbanization predicts poxvirus and coccidian infections in house finches (Haemorhous mexicanus). PLoS ONE 9, e86747 (2014).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Shutt, J. D. & Lees, A. C. Killing with kindness: Does widespread generalised provisioning of wildlife help or hinder biodiversity conservation efforts? Biol. Conserv. 261, 109295 (2021).Article 

    Google Scholar 
    Van Doren, B. M. et al. Human activity shapes the wintering ecology of a migratory bird. Glob. Chang. Biol. 27, 2715–2727 (2021).PubMed 
    Article 
    CAS 

    Google Scholar 
    Plummer, K. E., Risely, K., Toms, M. P. & Siriwardena, G. M. The composition of British bird communities is associated with long-term garden bird feeding. Nat. Commun. 10, 2088 (2019).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Lawson, B. et al. Health hazards to wild birds and risk factors associated with anthropogenic food provisioning. Philos. Trans. R. Soc. B Biol. Sci. 373, 20170091 (2018).Galbraith, J. A., Stanley, M. C., Jones, D. N. & Beggs, J. R. Experimental feeding regime influences urban bird disease dynamics. J. Avian Biol. 48, 700–713 (2017).Article 

    Google Scholar 
    Siriwardena, G. M. et al. The effect of supplementary winter seed food on breeding populations of farmland birds: Evidence from two large-scale experiments. J. Appl. Ecol. 44, 920–932 (2007).Article 

    Google Scholar 
    Kubasiewicz, L. M., Bunnefeld, N., Tulloch, A. I. T., Quine, C. P. & Park, K. J. Diversionary feeding: An effective management strategy for conservation conflict? Biodivers. Conserv. 25, 1–22 (2016).Article 

    Google Scholar 
    Lawson, B. et al. A clonal strain of Trichomonas gallinae is the aetiologic agent of an emerging avian epidemic disease. Infect. Genet. Evol. 11, 1638–1645 (2011).PubMed 
    Article 

    Google Scholar 
    Robinson, R. A. et al. Emerging infectious disease leads to rapid population declines of common British birds. PLoS ONE 5, e12215 (2010).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Forrester, D. J. & Foster, G. W. Trichomonosis. In: Parasitic Diseases of Wild Birds 120–153 (Wiley-Blackwell, 2008).Lawson, B. et al. Evidence of spread of the emerging infectious disease, finch trichomonosis, by migrating birds. EcoHealth 8, 143–153 (2011).PubMed 
    Article 

    Google Scholar 
    Lawson, B. et al. The emergence and spread of finch trichomonosis in the British Isles. Philos. Trans. R. Soc. B Biol. Sci. 367, 2852–2863 (2012).Article 

    Google Scholar 
    Woodward, I. D. et al. BirdTrends 2020: Trends in numbers, breeding success and survival for UK breeding birds. Research Report 732. BTO, Thetford. (2020).Enoksson, B. Age- and sex-related differences in dominance and foraging behaviour of nuthatches Sitta europaea. Anim. Behav. 36, 231–238 (1988).Article 

    Google Scholar 
    Tarvin, K. A. & Woolfenden, G. E. Patterns of dominance and aggressive behavior in blue jays at a feeder. Condor 99, 434–444 (1997).Article 

    Google Scholar 
    Brittingham, M. C. & Temple, S. A. Use of winter feeders by black-capped chickadees. Wildl. Soc. 56, 103–110 (1992).
    Google Scholar 
    Woodward, I. et al. Population estimates of birds in Great Britain and the United Kingdom. Br. Birds 113, 69–104 (2020).
    Google Scholar 
    Musgrove, A. J. et al. Population estimates of birds in Great Britain and the United Kingdom. Br. Birds 106, 64–100 (2013).
    Google Scholar 
    Wernham, C. et al. The Migration Atlas: Movements of the Birds of Britain and Ireland. (T & AD Poyser, 2002).Main, I. G. The partial migration of Fennoscandian Greenfinches Carduelis chloris. Ringing Migr. 20, 167–180 (2000).Article 

    Google Scholar 
    Lack, P. C. The Atlas of Wintering Birds in Britain and Ireland. (T. & A.D. Poyser, 1986).Robinson, R. A. BirdFacts: profiles of birds occurring in Britain & Ireland. BTO, Thetford (2005). Available at: http://www.bto.org/birdfacts. Accessed: 15 May 2022.Tratalos, J. et al. Bird densities are associated with household densities. Glob. Chang. Biol. 13, 1685–1695 (2007).ADS 
    Article 

    Google Scholar 
    Gregory, R. D. Broad-scale habitat use of sparrows, finches and buntings in Britain. Die Vogelwelt 120, 47–57 (1999).
    Google Scholar 
    Newton, I. Finches. New Naturalist Series, Volume: 55. (HarperCollins, 1972).Robinson, R. A., Baillie, S. R. & Crick, H. Q. P. Weather-dependent survival: Implications of climate change for passerine population processes. Ibis. 149, 357–364 (2007).Article 

    Google Scholar 
    Crick, H. Q. P. A bird-habitat coding system for use in Britain and Ireland incorporating aspects of land-management and human activity. Bird Study 39, 1–12 (1992).Article 

    Google Scholar 
    Davies, Z. G. et al. A national scale inventory of resource provision for biodiversity within domestic gardens. Biol. Conserv. 142, 761–771 (2009).Article 

    Google Scholar 
    Balmer, D. E. et al. Bird Atlas 2007–11: The breeding and wintering birds of Britain and Ireland. (BTO Books, 2013).Lawson, B. et al. Epidemiology of salmonellosis in garden birds in England and Wales, 1993 to 2003. EcoHealth 7, 294–306 (2010).CAS 
    PubMed 
    Article 

    Google Scholar 
    Svensson, L. Identification guide to European passerines, 4th edition. (BTO, 1992).Jenni, L. & Winkler, R. Moult and ageing of European passerines, 2nd edition. (Helm, 2020).Baillie, S. R. The contribution of ringing to the conservation and management of bird populations: A review. Ardea 89, 167–184 (2001).
    Google Scholar 
    Kéry, M. & Schaub, M. Bayesian Population Analysis using WinBUGS: A hierarchical perspective (Academic Press, 2012).
    Google Scholar 
    R Core Team. R: A language and environment for statistical computing. (2020).Plummer, M. JAGS: A program for analysis of Bayesian graphical models using Gibbs sampling. in Proceedings of the 3rd International Workshop on Distributed Statistical Computing (DSC 2003) (eds. Hornik, K., Leisch, F. & Zeileis, A.) (2003).Su, Y.-S. & Yajima, M. R2jags: Using R to Run ‘JAGS’. R package version 0.6–1. (2020).Robinson, R. A., Morrison, C. A. & Baillie, S. R. Integrating demographic data: Towards a framework for monitoring wildlife populations at large spatial scales. Methods Ecol. Evol. 5, 1361–1372 (2014).Article 

    Google Scholar 
    Newson, S. E., Evans, K. L., Noble, D. G., Greenwood, J. J. D. & Gaston, K. J. Use of distance sampling to improve estimates of national population sizes for common and widespread breeding birds in the UK. J. Appl. Ecol. 45, 1330–1338 (2008).Article 

    Google Scholar 
    Newson, S. E., Massimino, D., Johnston, A., Baillie, S. R. & Pearce-Higgins, J. W. Should we account for detectability in population trends? Bird Study 60, 384–390 (2013).Article 

    Google Scholar 
    Crick, H. Q. P., Baillie, S. R. & Leech, D. I. The UK Nest Record Scheme: its value for science and conservation. Bird Study 50, 254–270 (2003).Article 

    Google Scholar 
    Abadi, F., Gimenez, O., Arlettaz, R. & Schaub, M. An assessment of integrated population models: Bias, accuracy, and violation of the assumption of independence. Ecology 91, 7–14 (2010).PubMed 
    Article 

    Google Scholar 
    Plard, F., Turek, D., Grüebler, M. U. & Schaub, M. IPM2: Toward better understanding and forecasting of population dynamics. Ecol. Monogr. 89, e01364 (2019).Article 

    Google Scholar 
    Weegman, M. D., Arnold, T. W., Clark, R. G. & Schaub, M. Partial and complete dependency among data sets has minimal consequence on estimates from integrated population models. Ecol. Appl. 31, e02258 (2021).Article 

    Google Scholar 
    Koons, D. N., Iles, D. T., Schaub, M. & Caswell, H. A life-history perspective on the demographic drivers of structured population dynamics in changing environments. Ecol. Lett. 19, 1023–1031 (2016).PubMed 
    Article 

    Google Scholar 
    Koons, D. N., Arnold, T. W. & Schaub, M. Understanding the demographic drivers of realized population growth rates. Ecol Appl. 27, 2102–2115 (2017).PubMed 
    Article 

    Google Scholar 
    Caswell, H. Matrix population models: Construction, analysis and interpretation. (Sinauer Associates, 2001).Stubben, C. & Milligan, B. Estimating and analyzing demographic models using the popbio package in R. J. Stat. Softw. 22, 1–23 (2007).Article 

    Google Scholar 
    Stanbury, A. et al. The status of our bird populations: The fifth Birds of Conservation Concern in the United Kingdom, Channel Islands and Isle of Man and second IUCN Red List assessment of extinction risk for Great Britain. Br. Birds 114, 723–747 (2021).
    Google Scholar 
    Lehikoinen, A., Lehikoinen, E., Valkama, J., Väisänen, R. A. & Isomursu, M. Impacts of trichomonosis epidemics on greenfinch Chloris chloris and chaffinch Fringilla coelebs populations in Finland. Ibis 155, 357–366 (2013).Article 

    Google Scholar 
    PECBMS. EBCC/BirdLife/RSPB/CSO’ Pan-European Common Bird Monitoring Scheme. (2021). Available at: https://pecbms.info/. (Accessed: 14th July 2022)Keller, V. et al. European Breeding Bird Atlas 2: Distribution, Abundance and Change. (European Bird Census Council and Lynx Edicions, 2020).Rijks, J. M. et al. Trichomonosis in greenfinches (Chloris chloris) in the Netherlands 2009–2017: A concealed threat. Front. Vet. Sci. 6, 425 (2019).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Boele, A. et al. Broedvogels in Nederland in 2020. Sovonrapport 2022/05. (Sovon Vogelonderzoek Nederland, Nijmegen., 2022).Jones, D. The Birds at My Table: Why We Feed Wild Birds and Why It Matters. (Cornell University Press, 2018).Pennycott, T. W. et al. Causes of death of wild birds of the family fringillidae in Britain. Vet. Rec. 143, 155–158 (1998).CAS 
    PubMed 
    Article 

    Google Scholar 
    Bouwman, K. M. & Hawley, D. M. Sickness behaviour acting as an evolutionary trap? Male house finches preferentially feed near diseased conspecifics. Biol. Lett. 6, 462–465 (2010).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Lawson, B. et al. Acute necrotising pneumonitis associated with Suttonella ornithocola infection in tits (Paridae). Vet. J. 188, 96–100 (2011).PubMed 
    Article 

    Google Scholar 
    Clewley, G. D., Robinson, R. A. & Clark, J. A. Estimating mortality rates among passerines caught for ringing with mist nets using data from previously ringed birds. Ecol. Evol. 8, 5164–5172 (2018).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Francis, M. L. et al. Effects of supplementary feeding on interspecific dominance hierarchies in garden birds. PLoS ONE 13, e0202152 (2018).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Wojczulanis-Jakubas, K., Kulpińska, M. & Minias, P. Who bullies whom at a garden feeder? Interspecific agonistic interactions of small passerines during a cold winter. J. Ethol. 33, 159–163 (2015).Article 

    Google Scholar 
    Cramp, S. Handbook of the Birds of Europe, the Middle East and North Africa. Volume VIII: Crows to Finches. (Oxford University Press, 1994).Brook, B. W. & Bradshaw, C. J. A. Strength of evidence for density dependence in abundance time series of 1198 species. Ecology 87, 1445–1451 (2006).PubMed 
    Article 

    Google Scholar 
    Hochachka, W. M. & Dhondt, A. A. Density-dependent decline of host abundance resulting from a new infectious disease. Proc. Natl. Acad. Sci. USA. 97, 5303–5306 (2000).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Hochachka, W. M., Dobson, A. P., Hawley, D. M. & Dhondt, A. A. Host population dynamics in the face of an evolving pathogen. J. Anim. Ecol. 90, 1480–1491 (2021).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Chi, J. F. et al. The finch epidemic strain of Trichomonas gallinae is predominant in British non-passerines. Parasitology 140, 1234–1245 (2013).PubMed 
    Article 

    Google Scholar 
    Orros, M. E. & Fellowes, M. D. E. Wild bird feeding in an urban area: Intensity, economics and numbers of individuals supported. Acta Ornithol. 50, 43–58 (2015).Article 

    Google Scholar 
    Dirren, S., Borel, S., Wolfrum, N. & Korner-Nievergelt, F. Trichomonas gallinae infections in the naïve host Montifringilla nivalis subsp nivalis. J. Ornithol. 163, 333–337 (2022).Article 

    Google Scholar 
    Tulloch, A. I. T., Possingham, H. P., Joseph, L. N., Szabo, J. & Martin, T. G. Realising the full potential of citizen science monitoring programs. Biol. Conserv. 165, 128–138 (2013).Article 

    Google Scholar 
    Silvertown, J., Buesching, C., Jacobson, S. & Rebelo, T. Citizen science and nature conservation. in Key Topics in Conservation Biology 2 (eds. Macdonald, D. W. & Willis, K. J.) 127–142 (John Wiley & Sons, 2013).Dickinson, J. L., Zuckerberg, B. & Bonter, D. N. Citizen science as an ecological research tool: Challenges and benefits. Annu. Rev. Ecol. Evol. Syst. 41, 149–172 (2010).Article 

    Google Scholar 
    Baillie, S. R., Wernham, C. V. & Clark, J. A. Development of the British and Irish ringing scheme and its role in conservation biology. Ringing Migr. 19, S5–S19 (1999).Article 

    Google Scholar 
    Greenwood, J. J. D. Citizens, science and bird conservation. J. Ornithol. 148, S77–S124 (2007).Article 

    Google Scholar 
    Horns, J. J., Adler, F. R. & Şekercioğlu, Ç. H. Using opportunistic citizen science data to estimate avian population trends. Biol. Conserv. 221, 151–159 (2018).Article 

    Google Scholar 
    Ryan, R. L., Kaplan, R. & Grese, R. E. Predicting volunteer commitment in environmental stewardship programmes. J. Environ. Plan. Manag. 44, 629–648 (2001).Article 

    Google Scholar 
    Maund, P. R. et al. What motivates the masses: Understanding why people contribute to conservation citizen science projects. Biol. Conserv. 246, 108587 (2020).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Martin, V. Y. & Greig, E. I. Young adults’ motivations to feed wild birds and influences on their potential participation in citizen science: An exploratory study. Biol. Conserv. 235, 295–307 (2019).Article 

    Google Scholar 
    Cox, D. T. C. & Gaston, K. J. Human–nature interactions and the consequences and drivers of provisioning wildlife. Philos.Trans. R. Soc. B Biol. Sci. 373, 20170092 (2018).Article 

    Google Scholar 
    Murray, M. H., Becker, D. J., Hall, R. J. & Hernandez, S. M. Wildlife health and supplemental feeding: A review and management recommendations. Biol. Conserv. 204, 163–174 (2016).Article 

    Google Scholar 
    Rocha, G. & Quillfeldt, P. Effect of supplementary food on age ratios of European turtle doves (Streptopelia turtur L.). Anim. Biodivers. Conserv. 38, 11–21 (2015).Article 

    Google Scholar  More

  • in

    Increased drought effects on the phenology of autumn leaf senescence

    Richardson, A. D. et al. Terrestrial biosphere models need better representation of vegetation phenology: results from the North American Carbon Program Site Synthesis. Glob. Change Biol. 18, 566–584 (2012).Article 

    Google Scholar 
    Keenan, T. F. et al. Net carbon uptake has increased through warming-induced changes in temperate forest phenology. Nat. Clim. Change 4, 598–604 (2014).CAS 
    Article 

    Google Scholar 
    Piao, S. L. et al. Leaf onset in the northern hemisphere triggered by daytime temperature. Nat. Commun. 6, 6911 (2015).CAS 
    Article 

    Google Scholar 
    Penuelas, J., Rutishauser, T. & Filella, I. Phenology feedbacks on climate change. Science 324, 887–888 (2009).CAS 
    Article 

    Google Scholar 
    Garonna, I. et al. Strong contribution of autumn phenology to changes in satellite-derived growing season length estimates across Europe (1982–2011). Glob. Change Biol. 20, 3457–3470 (2014).Article 

    Google Scholar 
    Piao, S. L. et al. Net carbon dioxide losses of northern ecosystems in response to autumn warming. Nature 451, 49–52 (2008).CAS 
    Article 

    Google Scholar 
    Zhao, Y. et al. ABA receptor PYL9 promotes drought resistance and leaf senescence. Proc. Natl Acad. Sci. USA 113, 1949–1954 (2016).CAS 
    Article 

    Google Scholar 
    Keskitalo, J., Bergquist, G., Gardestrom, P. & Jansson, S. A cellular timetable of autumn senescence. Plant Physiol. 139, 1635–1648 (2005).CAS 
    Article 

    Google Scholar 
    Liu, Q. et al. Delayed autumn phenology in the Northern Hemisphere is related to change in both climate and spring phenology. Glob. Change Biol. 22, 3702–3711 (2016).Article 

    Google Scholar 
    Wu, C. Y. et al. Contrasting responses of autumn-leaf senescence to daytime and night-time warming. Nat. Clim. Change 8, 1092–1096 (2018).CAS 
    Article 

    Google Scholar 
    Zani, D., Crowther, T. W., Mo, L., Renner, S. S. & Zohner, C. M. Increased growing-season productivity drives earlier autumn leaf senescence in temperate trees. Science 370, 1066–1071 (2020).CAS 
    Article 

    Google Scholar 
    Zhang, Y., Parazoo, N. C., Williams, A. P., Zhou, S. & Gentine, P. Large and projected strengthening moisture limitation on end-of-season photosynthesis. Proc. Natl Acad. Sci. USA 117, 9216–9222 (2020).CAS 
    Article 

    Google Scholar 
    Grossiord, C. et al. Plant responses to rising vapor pressure deficit. New Phytol. 226, 1550–1566 (2020).Article 

    Google Scholar 
    Ciais, P. et al. Europe-wide reduction in primary productivity caused by the heat and drought in 2003. Nature 437, 529–533 (2005).CAS 
    Article 

    Google Scholar 
    Keenan, T. F. & Richardson, A. D. The timing of autumn senescence is affected by the timing of spring phenology: implications for predictive models. Glob. Change Biol. 21, 2634–2641 (2015).Article 

    Google Scholar 
    Liu, L. B. et al. Soil moisture dominates dryness stress on ecosystem production globally. Nat. Commun. 11, 4892 (2020).CAS 
    Article 

    Google Scholar 
    Delpierre, N. et al. Modelling interannual and spatial variability of leaf senescence for three deciduous tree species in France. Agric. For. Meteorol. 149, 938–948 (2009).Article 

    Google Scholar 
    Piao, S. L. et al. Weakening temperature control on the interannual variations of spring carbon uptake across northern lands. Nat. Clim. Change 7, 359–363 (2017).CAS 
    Article 

    Google Scholar 
    Fu, Y. S. H. et al. Declining global warming effects on the phenology of spring leaf unfolding. Nature 526, 104–107 (2015).CAS 
    Article 

    Google Scholar 
    Seastedt, T. R. & Knapp, A. K. Consequences of nonequilibrium resource availability across multiple time scales: the transient maxima hypothesis. Am. Nat. 141, 621–633 (1993).CAS 
    Article 

    Google Scholar 
    Korner, C. Paradigm shift in plant growth control. Curr. Opin. Plant Biol. 25, 107–114 (2015).CAS 
    Article 

    Google Scholar 
    Huxman, T. E. et al. Convergence across biomes to a common rain-use efficiency. Nature 429, 651–654 (2004).CAS 
    Article 

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

    Google Scholar 
    Nolan, R. H. et al. Differences in osmotic adjustment, foliar abscisic acid dynamics, and stomatal regulation between an isohydric and anisohydric woody angiosperm during drought. Plant Cell Environ. 40, 3122–3134 (2017).CAS 
    Article 

    Google Scholar 
    Fan, Y., Miguez-Macho, G., Jobbágy, E. G., Jackson, R. B. & Otero-Casal, C. Hydrologic regulation of plant rooting depth. Proc. Natl Acad. Sci. USA 114, 10572–10577 (2017).CAS 
    Article 

    Google Scholar 
    Choat, B. et al. Triggers of tree mortality under drought. Nature 558, 531–539 (2018).CAS 
    Article 

    Google Scholar 
    Giardina, F. et al. Tall Amazonian forests are less sensitive to precipitation variability. Nat. Geosci. 11, 405–409 (2018).CAS 
    Article 

    Google Scholar 
    Kannenberg, S. A., Driscoll, A. W., Szejner, P., Anderegg, W. R. L. & Ehleringer, J. R. Rapid increases in shrubland and forest intrinsic water-use efficiency during an ongoing megadrought. Proc. Natl Acad. Sci. USA https://doi.org/10.1073/pnas.2118052118 (2021).Liu, Q. et al. Extension of the growing season increases vegetation exposure to frost. Nat. Commun. https://doi.org/10.1038/s41467-017-02690-y (2018).Schuur, E. A. G. et al. Climate change and the permafrost carbon feedback. Nature 520, 171–179 (2015).CAS 
    Article 

    Google Scholar 
    Samaniego, L. et al. Anthropogenic warming exacerbates European soil moisture droughts. Nat. Clim. Change 8, 421–426 (2018).Article 

    Google Scholar 
    Templ, B. et al. Pan European Phenological database (PEP725): a single point of access for European data. Int. J. Biometeorol. 62, 1109–1113 (2018).Article 

    Google Scholar 
    Shen, M. et al. Increasing altitudinal gradient of spring vegetation phenology during the last decade on the Qinghai-Tibetan Plateau. Agric. For. Meteorol. 189, 71–80 (2014).Article 

    Google Scholar 
    Zhang, X. Y. Reconstruction of a complete global time series of daily vegetation index trajectory from long-term AVHRR data. Remote Sens. Environ. 156, 457–472 (2015).Article 

    Google Scholar 
    Chen, J. et al. A simple method for reconstructing a high-quality NDVI time-series data set based on the Savitzky–Golay filter. Remote Sens. Environ. 91, 332–344 (2004).Article 

    Google Scholar 
    White, M. A. et al. Intercomparison, interpretation, and assessment of spring phenology in North America estimated from remote sensing for 1982–2006. Glob. Change Biol. 15, 2335–2359 (2009).Article 

    Google Scholar 
    Zhang, X. et al. Monitoring vegetation phenology using MODIS. Remote Sens. Environ. 84, 471–475 (2003).Article 

    Google Scholar 
    Gonsamo, A., Chen, J. M., Price, D. T., Kurz, W. A. & Wu, C. Y. Land surface phenology from optical satellite measurement and CO2 eddy covariance technique. J. Geophys. Res. 117, G03032 (2012).
    Google Scholar 
    Muñoz, S. ERA5-Land Monthly Averaged Data from 1981 to Present (C3S CDS, date accessed:10-8-2021); https://doi.org/10.24381/cds.68d2bb30Abatzoglou, J. T., Dobrowski, S. Z., Parks, S. A. & Hegewisch, K. C. TerraClimate, a high-resolution global dataset of monthly climate and climatic water balance from 1958–2015. Sci. Data 5, 170191 (2018).Article 

    Google Scholar 
    Müller, W. A. et al. A Higher-resolution version of the Max Planck Institute Earth System Model (MPI-ESM1.2-HR). J. Adv. Model. Earth Syst. 10, 1383–1413 (2018).Article 

    Google Scholar 
    Vicente-Serrano, S. M. et al. Response of vegetation to drought time-scales across global land biomes. Proc. Natl Acad. Sci. USA 110, 52–57 (2013).CAS 
    Article 

    Google Scholar 
    Allen, R. G., Smith, M., Pereira, L. S. & Perrier, A. An update for the calculation of reference evapotranspiration. ICID Bull. 43, 64–92 (1994).
    Google Scholar 
    Gampe, D. et al. Increasing impact of warm droughts on northern ecosystem productivity over recent decades. Nat. Clim. Change https://doi.org/10.1038/s41558-021-01112-8 (2021).Sheffield, J., Wood, E. F. & Roderick, M. L. Little change in global drought over the past 60 years. Nature 491, 435–438 (2012).CAS 
    Article 

    Google Scholar 
    Peng, J., Wu, C. Y., Zhang, X. Y., Wang, X. Y. & Gonsamo, A. Satellite detection of cumulative and lagged effects of drought on autumn leaf senescence over the Northern Hemisphere. Glob. Change Biol. 25, 2174–2188 (2019).Article 

    Google Scholar 
    Harris, I., Jones, P. D., Osborn, T. J. & Lister, D. H. Updated high-resolution grids of monthly climatic observations—the CRU TS3.10 Dataset. Int. J. Climatol. 34, 623–642 (2014).Article 

    Google Scholar 
    Beaudoing, H., Rodell, M. & NASA/GSFC/HSL. GLDAS Noah Land Surface Model L4 3 Hourly 0.25 × 0.25 Degree Version 2.0 (GES DISC, 2015); https://doi.org/10.5067/342OHQM9AK6QBeaudoing, H., Rodell, M. & NASA/GSFC/HSL. GLDAS Noah Land Surface Model L4 3 Hourly 0.25 ×0.25 Degree Version 2.1 (GES DISC, 2016); https://doi.org/10.5067/E7TYRXPJKWOQZheng, Y. et al. Improved estimate of global gross primary production for reproducing its long-term variation, 1982–2017. Earth Syst. Sci. Data 12, 2725–2746 (2020).Article 

    Google Scholar 
    Zhang, K. et al. Vegetation greening and climate change promote multidecadal rises of global land evapotranspiration. Sci. Rep. https://doi.org/10.1038/srep15956 (2015).Li, Y. et al. Estimating global ecosystem isohydry/anisohydry using active and passive microwave satellite data. J. Geophys. Res. 122, 3306–3321 (2017).Article 

    Google Scholar 
    Moesinger, L. et al. The global long-term microwave Vegetation Optical Depth Climate Archive (VODCA). Earth Syst. Sci. Data 12, 177–196 (2020).Gupta, H. V., Kling, H., Yilmaz, K. K. & Martinez, G. F. Decomposition of the mean squared error and NSE performance criteria: implications for improving hydrological modelling. J. Hydrol. 377, 80–91 (2009).Article 

    Google Scholar 
    Botta, A., Viovy, N., Ciais, P., Friedlingstein, P. & Monfray, P. A global prognostic scheme of leaf onset using satellite data. Glob. Change Biol. 6, 709–725 (2000).Article 

    Google Scholar  More

  • in

    A Cryptochrome adopts distinct moon- and sunlight states and functions as sun- versus moonlight interpreter in monthly oscillator entrainment

    l-cry mutants show higher spawning synchrony than wild-type animals under non-natural light conditionsIn order to test for a functional involvement of L-Cry in monthly oscillator function, we generated two l-cry mutant alleles (Δ34 and Δ11bp) (Fig. 1a) using TALENs28. In parallel, we generated a monoclonal antibody against Platynereis L-Cry. By testing mutant versus wildtype worms with the anti-L-Cry antibody in Western blots (Fig. 1b) and immunohistochemistry (Fig. 1e–j), we verified the absence of L-Cry protein in mutants. Furthermore, we confirmed that the staining of the antibody in wildtype worms (Fig. 1e–h) matches the regions where l-cry mRNA is expressed (Fig. 1d). These tests confirmed that the engineered l-cry mutations result in loss-of-function alleles. In turn, they validate the specificity of the raised anti-L-Cry antibody.Fig. 1: l-cry–/– mutants are loss-of-function alleles.a Overview of the l-cry genomic locus for wt and mutants. Both mutant alleles result in an early frameshift and premature stop codons. The Δ34 allele has an additional 9 bp deletion in exon 3. b Western Blots of P. dumerilii heads probed with anti-L-Cry antibody. In the context of further investigations such Western blots of mutant versus wild types have been performed more than 10 times with highly consistent results. Also see further analyses in this manuscript and ref. 36. c overview of P. dumerilii. d whole mount in situ hybridization against l-cry mRNA on worm head. ae, anterior eye; pe, posterior eye. e–j Immunohistochemistry of premature wild-type (e–h) and mutant (i, j) worm heads sampled at zt19/20 using anti-L-Cry antibody (green) and Hoechst staining (magenta), dorsal views, anterior up. e, f: z-stack images (maximal projections of 50 layers, 1.28 µm each) in the area highlighted by the rectangle in (d), whereas (g–j) are single layer images of the area highlighted by the white rectangles in (e, f). In the context of further investigations such stainings of mutant versus wild types have been performed more than 10 times with highly consistent results. Also see further analyses in this manuscript and ref. 36.Full size imageWe next assessed the circalunar maturation timing of wild types and l-cry mutant populations in conventional culture conditions, i.e. worms grown under typical indoor room lighting (named here artificial sun- and moonlight, Supplementary Fig. 1b).We expected either no phenotype (if L-Cry was not involved in circalunar clock entrainment) or a decreased spawning precision (if L-Cry was functioning as moonlight receptor in circalunar clock entrainment). Instead we observed an increased precision of the entrained worm population:We analyzed the maturation data using two statistical approaches, linear and circular statistics. We used the classical linear plots5 and statistics to compare the monthly spawning data distribution (Fig. 2a–c, i). This revealed a clear difference between mutant animals, which exhibited a stronger spawning peak at the beginning of the NM phase, compared to their wildtype and heterozygous counterparts (Fig. 2a–c, Kolmogorov–Smirnov test on overall data distribution, Fig. 2i).Fig. 2: L-Cry shields the circalunar clock from light that is not naturalistic moonlight.a–d, j Spawning of l-cry +/+ (a), l-cry +/– (Δ34) (b) and l-cry −/−(Δ34/ Δ34) (c) animals over the lunar month in the lab with 8 nights of artificial moonlight (a–c), under natural conditions in the sea (d, replotted from ref. 34,50,) and in the lab using naturalistic sun- and moonlight (j, 8 nights moonlight). e–h, k Data as in (a–d, j) as circular plot. 360° correspond to 30 days of the lunar month. The arrow represents the mean vector, characterized by the direction angle µ and r (length of µ). r indicates phase coherence (measure of population synchrony). p-values inside the plots: result of Rayleigh Tests. Significance indicates non-random distribution of data points. The inner circle represents the Rayleigh critical value (p = 0.05). i–l Results of two-sided multisample statistics on spawning data shown in (a–h, j, k). The phase differences in days can be calculated from the angle between the two mean vectors (i.e. 12°= 1 day).Full size imageWe then analyzed the same data using circular statistics (as the monthly cycle is repeating, see details in Methods section), which allowed us to describe the data with the mean vector (defined by the direction angle µ and its length r, shown as arrows in Fig. 2e–g). The phase coherence r (ranging from 0 to 1) serves as a measure for synchrony of the population data. As expected for entrained populations, all genotypes distributed their spawning across a lunar month significantly different from random (Fig. 2e–g, p values in circles, Rayleigh’s Uniformity test29). In line with the observed higher spawning peak of the l-cry−/− mutants in the linear plots, the circular analysis revealed a significant difference in spawning distribution (Mardia–Watson-Wheeler test, for details see Methods section) and higher spawning synchrony of mutants (r = 0.614) than in wild types and heterozygotes (r = 0.295 and r = 0.222) (Fig. 2i). The specificity of this phenotype of higher spawning precision for l-cry homozygous mutants was confirmed by analyses on trans-heterozygous l-cry (Δ34/Δ11) mutants (Supplementary Fig. 2), and by the fact that such a phenotype is not detectable in any other light receptor mutant available in Platynereis (r-opsin130: Supplementary Fig. 3a, b, e, f, i; c-opsin131: Supplementary Fig. 3c, d, g, h, i, Go-opsin: refs. 32, 33).The higher spawning synchrony of l-cry mutants under artificial light mimics the spawning precision of wild-type at its natural habitatThis increased spawning precision of l-cry mutants under artificial (but conventional indoor) laboratory light conditions let us wonder about the actual population synchrony of the worms under truly natural conditions. The lunar spawning synchrony of P. dumerilii at the Bay of Naples (the origin of our lab culture) has been worked on for more than 100 y. This allowed us to re-investigate very detailed spawning data records from the worms’ natural habitat published prior to environmental/light pollution. For better accessibility and comparability we combined all months and replotted the data published in 192934 (Fig. 2d, h, I; see details in Methods section; r = 0.631). This analysis revealed that the higher spawning synchrony in l-cry–/– worms mimics the actual spawning synchrony of P. dumerillii populations in their natural habitat34 (compare Fig. 2c, g with 2d, h.)Given that recent, non-inbred isolates from the same habitat as our lab inbred strains (which is the same habitat as the data collected in ref. 34) exhibit a broad spawning distribution under standard worm culture light conditions (which includes the bright artificial moonlight)35, we hypothesized that the difference in spawning synchrony between wildtype laboratory cultures and populations in their natural habitat is caused by the rather bright nocturnal light stimulus typically used for the standard laboratory culture (Supplementary Fig. 1a vs. b).Lunar spawning precision of wild-type animals depends on naturalistic moonlight conditionsWe next tested the resulting prediction that naturalistic moonlight should increase the spawning precision of the wildtype population, using naturalistic sun- and moonlight devices we specifically designed based on light measurements at the natural habitat of P. dumerilii31 (Supplementary Fig. 1a, c). We assessed the impact of the naturalistic sun- and moonlight (Supplementary Fig. 1a, c) on wildtype animals, maintaining the temporal aspects of the lab light regime (i.e. 8 nights of “full moon”). Indeed, merely adjusting the light intensity to naturalistic conditions increased the precision and phase coherence of population-wide reproduction: After several months under naturalistic sun- and moonlight, wildtype worms spawned with a major peak highly comparable to the wildtype precision reported at its natural habitat (Fig. 2d, h vs. j, k), and also exhibited an increased population synchrony (r = 0.398 compared to r = 0.295 under standard worm room light conditions). This increased similarity to the spawning distribution at the natural habitat (“Sea”) is confirmed by statistical analyses (Fig. 2l): The phase difference (angle between the two mean vectors) is only one day (corresponding to 12°). In contrast, the spawning distribution of wild types under standard worm room light versus naturalistic light conditions is highly significantly different in linear and circular statistical tests and has a phase difference of 7.7 days (Fig. 2l).These findings show that it is the naturalistic light that is critical for a highly precise entrainment of the monthly clock of wild-type worms. Given that l-cry–/– animals reach this high precision with the artificial light (i.e. standard lab light) implies that in wildtype L-Cry blocks artificial, but not naturalistic full-moonlight from efficiently synchronizing the circalunar clock. This block is removed in l-cry–/– animals, leading to a better synchronization of the l-cry–/– population. This finding suggests that L-Cry’s major role could be that of a gatekeeper controlling which ambient light is interpreted as full-moonlight stimulus for circalunar clock entrainment.
    l-cry functions as a light signal gatekeeper for circalunar clock entrainmentA prediction of this hypothesis is that mutants should entrain better to an artificial full-moonlight stimulus provided out-of-phase than their wild type counterparts (in which L-Cry should block the “wrong” moonlight at least partially from re-entraining the circalunar oscillator).We thus compared the spawning rhythms of l-cry+/+ and l-cry–/– worms under a re-entrainment paradigm, where we provided our bright artificial culture full-moonlight at the time of the subjective new moon phase (Fig. 3a). In order to compare the spawning data distribution relative to the initial full moon (FM) stimulus, as well as to the new full moon stimulus (i.e. new FM), we used two nomenclatures for the months: months with numbers are analyzed relative to the initial nocturnal light stimulus (i.e. FM), whereas months with letters are analyzed relative to the new (phase-shifted) nocturnal light stimulus (i.e. new FM, Fig. 3a). When the nocturnal light stimulus is omitted (to test for the oscillator function) we then refer to ‘free-running FM’ (FR-FM) or ‘new free-running FM’ (new FR-FM), respectively (Fig. 3a). Using these definitions, the efficiency of circalunar clock re-entrainment will be reflected in the similarity of spawning data distributions between month 1 and month D, i.e. the more similar the distribution, the more the population has shifted to the new phase.Fig. 3: l-cry−/− mutants entrain the circalunar clock faster than wt to a high-intensity artificial moonlight stimulus.a Nocturnal moonlight exposure protocol of lunar phase shift (entrained by 8 nights, phased shifted by 6 nights of artificial culture moon, light green). b, c Number of mature animals (percent per month, rolling mean with a window of 3 days) of l-cry wild-type (b) and homozygous mutant (c) animals. p-values indicate results of Kolomogorov–Smirnov tests. Dark blue arrowheads- old FM phase: wt show a spawning minimum, indicative that the worms are not properly phase shifted. Mutants spawn in high numbers, but don’t spawn at the old NM indicated by light blue arrowhead. Also compare to initial FM and NM in months 1,2. d, e Circular plots of the data shown in (b) and (c). Each circle represents one lunar month. Each dot represents one mature worm. The arrow represents the mean vector characterized by the direction angle µ and r. r (length of µ) indicates phase coherence (measure of population synchrony). The inner circle represents the Rayleigh critical value (p = 0.05). f, g Results of two-sided multisample statistics of data in (d, e). Phase differences in days can be calculated from the angle between the two mean vectors (i.e. 12°= 1 day).Full size imageWhen using the artificial nocturnal light conditions, the re-entrainment of l-cry–/– animals was both faster and more complete than for their wildtype relatives, as predicted from our gate keeper hypothesis. This is evident from the linear data analysis and Kolmogorov–Smirnov tests when comparing the month before the entrainment (month 1) with two months that should be shifted after the entrainment (months C,D, Fig. 3b, c, f, g).Most notably, while l-cry−/− worms were fully shifted in month D (Fig. 3c: compare boxes and see complete lack of spawning at the light blue arrowhead indicating the old NM/new FR-FM phase versus massive spawning at new NM phase around dark blue arrowhead), wildtype animals were still mostly spawning according to the initial lunar phase (Fig. 3b: compare boxes and see spawning at the light blue arrowhead versus almost lack of spawning at dark blue arrowhead). The faster re-entrainment of l-cry–/–, compared to l-cry+/+ animals is also confirmed by the Mardia–Watson-Wheeler test (see Methods section for details). For l-cry+/+ animals, the comparisons of the spawning distributions before and after re-entrainment show a 1000-fold (months 1 versus C) and tenfold (months 1 versus D) higher statistical significance difference than the corresponding comparisons for l-cry−/− worms (Fig. 3f, g). Consistently, the phase differences in days calculated from the angle between the two mean vectors from the circular analysis is smaller in the mutants than in the wild types when comparing the phase of the month before the entrainment (month 1) with two months after the entrainment (months C, D) (Fig. 3d–g). The fact that there are still differences in the mutant population before and after entrainment is likely due to the fact that even the mutants are not fully re-entrained. However, they have shifted more robustly in response to an artificial nocturnal light stimulus than the wild types. This provides further evidence that in wildtype worms L-Cry indeed blocks the “wrong” light from entering into the circalunar clock and thus functions as a light gatekeeper.L-Cry functions mainly as light interpreter, while its contribution as direct moonlight entraining photoreceptor is (at best) minorWe next tested to which extent L-Cry is itself a sensor for the re-entrainment signal under naturalistic light conditions. Based on the finding that l-cry−/− worms can still re-entrain the circalunar oscillator (see above), it is clear that even if L-Cry also directly contributed to the entrainment, it cannot be the only moonlight receptor mediating entrainment. With the experiments below, we aimed to test if L-Cry has any role as an entraining photoreceptor to the monthly oscillator.Thus, we tested how the circalunar clock is shifted in response to a re-entrainment with naturalistic moonlight in Platynereis wt versus l-cry−/− worms. For this, animals initially raised and entrained under standard worm room light conditions of artificial sun- and moonlight (Supplementary Fig. 1b, e) were challenged by a deviating FM stimulus of 8 nights of naturalistic moonlight (Fig. 4a, Supplementary Fig. 1c, e). This re-entraining stimulus was repeated for three consecutive months (Fig. 4a).Fig. 4: l-cry has a minor contribution as entraining photoreceptor to circalunar clock entrainment.a Nocturnal moonlight exposure protocol of lunar phase shift with 8 nights of naturalistic moonlight (dark green). Number of mature animals (percent per month, rolling mean with a window of 3 days) of l-cry wild-type (b) and mutant (c) animals. p-values: Kolomogorov–Smirnov tests. Black arrowheads indicate spawning-free intervals of the wildtype, which shifted to the position of the new FM (under free-running conditions: FR-FM). d, e Data as in (b, c) plotted as circular data. 360° correspond to 30 days of the lunar month. The arrow represents the mean vector characterized by the direction angle µ and r. r (length of µ) indicates phase coherence (measure of population synchrony). p values are results of Rayleigh Tests: Significance indicates non-random distribution of data points. The inner circle represents the Rayleigh critical value (p = 0.05). f, g Results of two-sided multisample statistics on spawning data shown in (a–e). Phase differences in days can be calculated from the angle between the two mean vectors (i.e. 12°= 1 day).Full size imageThe resulting spawning distribution was analyzed for the efficacy of the naturalistic moonlight to phase-shift the circalunar oscillator. In order to test if the animals had shifted their spawning to the new phase, we again compared the spawning pattern before the exposure to the new full moon stimulus (months with numbers: data distribution analyzed relative to the initial/old FM, see Fig. 4a for an overview) to the spawning pattern after the exposure to the new full moon stimulus (months with letters: data distribution analyzed relative to the new FM, Fig. 4a). The more similar the data distributions of month 1 is to the months C, D, the more the population was shifted to the new phase.The first re-entraining full moon stimulus (Fig. 4b, first dark green box) is given in the middle of the main spawning period. The nocturnal light itself does not cause immediate effects on the number of spawning worms (Fig. 4b, see also Fig. 2b, c), but the repeated exposure resulted in a noticeable shift of the spawning distribution indicating a phase shift of the monthly oscillator in wildtype. Already at the third re-entraining full moon stimulus, wildtype animals exhibited a completely shifted spawning pattern (Fig. 4b, d-d″, month 1, 2 vs. month C). This is supported by statistical analyses: When comparing the months 1 and 2 (relative to the old FM before the shift) to the month C (relative to the new FM after the shift), both the Kolmogorov–Smirnov test (Fig. 4b: gray rectangles, 4f) and the Mardia–Watson–Wheeler test of the same data were non-significant (Fig. 4f), indicative of the population shifting to the new phase. Consistently, the direction angle (µ) of the mean vectors before and after the shift was highly similar, resulting in a phase difference of only 0.2 days between months 1 and C and 0.5 days between month 2 and month C (Fig. 4f, for details see methods). The month under circalunar free-running conditions (month D) supports this observation, albeit with lower statistical support (Fig. 4b, d″, f).Of note, wild-type worms would eventually reach the high spawning precision found under naturalistic moonlight only after several more months based on independent experiments (Fig. 2j, k).When we analyzed the spawning distribution of l-cry mutants in the same way as the wild types, we found that the data distribution exhibited significant differences in the linear Kolmogorov–Smirnov test when comparing months 1 and 2 before the shift to the months C and D after the shift (Fig. 4c: gray rectangles, Fig. 4g); as well as in the phase distribution in the circular analyses when comparing the months before the shift (months 1 and 2) with the last months of the shift (months C,D) (Fig. 4e, e′ versus e″, e‴, g). The populations also exhibited a noticeable phase difference of ≥3.5 days (Fig. 4g).Based on the statistical significant difference in the re-entrainment of l-cry–/–, but not wild-type populations under a naturalistic sun- and moonlight regime, we conclude that L-Cry also likely contributes to circalunar entrainment as a photoreceptor. However, as these differences are rather minor, compared to the much stronger differences seen under artificial light regime, we conclude that its major role is the light gatekeeping function.In an independent study that focused on the impact of moonlight on daily timing, we identified r-Opsin1 as a lunar light receptor that mediates moonlight effects on the worms’ ~24 h clock36. We tested if r-opsin1 is similarly important for mediating the moonlight effects on the monthly oscillator of the worm, analyzed here. This is not the case. r-opsin1–/– animals re-entrain as well as wildtype worms under naturalistic light conditions (Supplementary Fig. 4). This adds to and is also consistent with our above observation that the spawning distribution is un-altered between r-opsin1–/– and wildtype animals under artificial light conditions (Supplementary Fig. 3a, b, e, f). This finding also further enforces the notion that monthly and daily oscillators use distinct mechanisms, but both require L-Cry as light interpreter.L-Cry discriminates between naturalistic sun- and moonlight by forming differently photoreduced statesGiven that the phenotype of l-cry–/– animals suggests a role of L-Cry as light gatekeeper, i.e. only allowing the ‘right’ light to most efficiently impact on the circalunar oscillator, we next investigated how this could function on the biochemical and cell biological level.While we have previously shown that Pdu-L-Cry is degraded upon light exposure in S2 cell culture15, it has remained unclear if L-Cry has the spectral properties and sensitivity to sense moonlight and whether this would differ from sunlight sensation. To test this, we purified full length L-Cry from insect cells (Supplementary Fig. 5a–c). Multi-angle light scattering (SEC-MALS) analyses of purified dark-state L-Cry revealed a molar mass of about 130 kDa, consistent with the formation of an L-Cry homodimer (theoretical molar mass of L-Cry monomer is 65.6 kDa) (Fig. 5a). Furthermore, purified L-Cry binds Flavin Adenine Dinucleotide (FAD) as its chromophore (Supplementary Fig. 5d, e). We then used UV/Vis absorption spectroscopy to analyze the FAD photoreaction of purified L-Cry in presence of 1 mM TCEP to prevent protein oxidation. The absorption spectrum of dark-state L-Cry showed maxima at 450 nm and 475 nm, consistent with the presence of oxidized FAD (Supplementary Fig. 5f, black line). As basic starting point to analyze its photocycle, L-Cry was photoreduced using a LED (PerkinElmer ACULED Dyo) with a blue-light dominated spectrum and spectral peak at 450 nm (Supplementary Fig. 1d, d′, henceforth referred to as “blue-light”) for 110 s37. The light-activated spectrum showed that blue-light irradiation of L-Cry leads to the complete conversion of FADox into an anionic FAD radical (FADo-) with characteristic FADo- absorption maxima at 370 nm and 404 nm and reduced absorbance at 450 nm (Supplementary Fig. 5f, blue spectrum, black arrows). In darkness, L-Cry reverted back to the dark-state with time constants of 2 min (18 °C), 4 min (6 °C) and 4.7 min (ice) (Supplementary Fig. 5g–k).Fig. 5: L-Cry forms differently photoreduced sunlight- and moonlight states.a Multi-Angle Light Scattering (MALS) analyses of dark-state L-Cry fractionated by size exclusion chromatography (SEC). Black dashed line: normalized UV absorbance, solid line: normalized scattering signal. The molar mass of about 130 kDa derived from MALS (mass signal shown in red) corresponds to an L-Cry homodimer. b Absorption spectrum of L-Cry in darkness (black) and after sunlight exposure (orange). Additional timepoints: Supplementary Fig. 6a. c Dark recovery of L-Cry after 20 min sunlight on ice. Absorbance at 450 nm in Supplementary Fig. 6b. d, e Absorption spectra of L-Cry after exposure to naturalistic moonlight for different durations. f Full spectra of dark recovery after 6 h moonlight. Absorbance at 450 nm: Supplementary Fig. 6d. g Absorption spectrum of L-Cry after 6 h of moonlight followed by 20 min of sunlight. h Absorption spectrum of L-Cry after 20 min sunlight followed by moonlight first results in dark-state recovery. Absorbance at 450 nm: Supplementary Fig. 6e. i Absorption spectrum of L-Cry after 20 min sunlight followed by 4 h and 6 h moonlight builds up the moonlight state. j Model of L-Cry responses to sunlight (orange), moonlight (green) and darkness (black). Only transitions between stably accumulating states are shown. Absorbances in (b–i) were normalized when a shift in the baseline occurred between different measurements of the same measurement set, which is then indicated on the Y-axis as “normalized absorbance”.Full size imageWe then investigated the response of L-Cry to ecologically relevant light, i.e. sun- and moonlight using naturalistic sun- and moonlight devices that we designed based on light measurements at the natural habitat of P. dumerilii31 (Supplementary Fig. 1a, c, e). Upon naturalistic sunlight illumination, FAD was photoreduced to FADo-, but with slower kinetics than under the stronger blue-light source, likely due to the intensity differences between the two lights (Supplementary Fig. 1c–e).While blue-light illumination led to a complete photoreduction within 110 s (Supplementary Fig. 5f), sunlight induced photoreduction to FADo- was completed after 5–20 min (Fig. 5b) and did not further increase upon continued illumation for up to 2 h (Supplementary Fig. 6a). Dark recovery kinetics had time constants of 3.2 min (18 °C) and 5 min (ice) (Fig. 1c, Supplementary Fig. 6b, c).As the absorbance spectrum of L-Cry overlaps with that of moonlight at the Platynereis natural habitat (Supplementary Fig. 1a), L-Cry has the principle spectral prerequisite to sense moonlight. However, the most striking characteristic of moonlight is its very low intensity (5.8 × 1010 photons/cm2/s at −5m, Supplementary Fig. 1a–e). To test if Pdu-L-Cry is sensitive enough for moonlight, we illuminated purified L-Cry with our custom-built naturalistic moonlight, closely resembling full-moonlight intensity and spectrum at the Platynereis natural habitat (Supplementary Fig. 1a, c, e). Naturalistic moonlight exposure up to 2.75 h did not markedly photoreduce FAD, notably there was no difference between 1 h and 2.75 h (Fig. 5d). However, further continuous naturalistic moonlight illumination of 4 h and longer resulted in significant changes (Fig. 5d), whereby the spectrum transitioned towards the light activated state of FADo- (note peak changes at 404 nm and at 450 nm). This photoreduction progressed further until 6 h naturalistic moonlight exposure (Fig. 5d). No additional photoreduction could be observed after 9 h and 12 h of naturalistic moonlight exposure (Fig. 5e), indicating a distinct state induced by naturalistic moonlight that reaches its maximum after ~6 h, when about half of the L-Cry molecules are photoreduced. This time of ~6 h is remarkably consistent with classical work showing that a minimum of ~6 h of continuous nocturnal light is important for circalunar clock entrainment, irrespective of the preceding photoperiod5. The dark recovery of L-Cry after 6 h moonlight exposure occurred with a time constant of 6.7 min at 18 °C (Fig. 5f, Supplementary Fig. 6d). Given that both sunlight and moonlight cause FAD photoreduction, but with different kinetics and different final FADo- product/FADox educt ratios, we wondered how purified L-Cry would react to transitions between naturalistic sun- and moonlight (i.e. during “sunrise” and “sunset”).Mimicking the sunrise scenario, L-Cry was first illuminated with naturalistic moonlight for 6 h followed by 20 min of sunlight exposure. This resulted in an immediate enrichment of the FADo- state (Fig. 5g). Hence, naturalistic sunlight immediately photoreduces remaining oxidized flavin molecules, that are characteristic of moonlight activated L-Cry, to FADo-, to reach a distinct fully reduced sunlight state.In contrast, when we next mimicked the day-night transition (“sunset”) by first photoreducing with naturalistic sunlight (or strong blue-light) and subsequently exposed L-Cry to moonlight, L-Cry first returned to its full dark-state within about 30 min (naturalistic sunlight: τ = 7 min (ice), Fig. 5h, Supplementary Fig. 6e; blue-light: τ = 9 min (ice), Supplementary Fig. 6f–h), despite the continuous naturalistic moonlight illumination. Prolonged moonlight illumination then led to the conversion of dark-state L-Cry to the moonlight state (Fig. 5i, Supplementary Fig. 6f). Hence, fully photoreduced sunlight-state L-Cry first has to return to the dark-state before accumulating the moonlight state characterized by the stable presence of the partial FADo- product/FADox educt. In contrast to sunlight-state L-Cry, moonlight-state L-Cry does not return to the oxidized (dark) state under naturalistic moonlight (Fig. 5e), i.e. moonlight maintains the moonlight state, but not the sunlight state. We note, that a partially photoreduced L-Cry state may be formed transiently during dark-state recovery of the sunlight state under moonlight. However, this transiently occurring partially photoreduced L-Cry state would differ from the “true” moonlight state (e.g. by an allosteric change) preventing its accumulation (see discussion and Supplementary Fig. 6i).Given that L-Cry forms a homodimer and moonlight photoreduces about half of the FAD molecules, we propose that the moonlight state corresponds to a half-reduced FADo- FADox dimer, where FAD is only photoreduced in one L-Cry monomer, whereas in the sunlight state both monomers are photoreduced (FADo- FADo-) (Fig. 5j). This implies that the quantum yield for FADox to FADo- photoreduction differs between the two L-Cry monomers. One monomer (referred to as “A” in Fig. 5j) acts as “very low intensity light sensor” with a high quantum yield ΦA. Hence, the very low photon number provided after 6 h of moonlight illumination is sufficient to photoreduce its flavin co-factor, resulting in the partially photoreduced FADo- FADox moonlight state (Fig. 5j).For direct comparison, our naturalistic moonlight’s emission (in the main absorbance range of L-Cry: 330 nm–510 nm) is 5.4 × 1010 photons/cm2/s (Supplementary Fig. 1e), which accumulates to ~1.2 × 1015 photons/cm2 in the 6 h required to reach the half-reduced moonlight state (Fig. 5d, e). For naturalistic sunlight, emitting ~7.5 × 1014 photons/cm2/s (330–510 nm), at least 5 min of sunlight illumination (i.e. > ~1.8 × 1017 photons/cm2) are required to photoreduce the flavin in both L-Cry monomers in order to reach the fully photoreduced FADo- FADo- sunlight state (Fig. 5b, j). Thus, the second L-Cry monomer (monomer “B” in Fig. 5j) has a significantly lower quantum yield ΦB for FAD photoreduction (ΦB  More

  • in

    A harmonized dataset of sediment diatoms from hundreds of lakes in the northeastern United States

    Smol, J. P. & Stoermer, E. F. The Diatoms: Application for the Environmental and Earth Sciences (Cambridge University Press, 2010).Charles, D. F. Relationships between surface sediment diatom assemblages and lake water characteristics in Adirondack lakes. Ecology 66, 994–111 (1985).Article 

    Google Scholar 
    Whitehead, D. R., Charles, D. F., Jackson, S. T., Reed S. E. & Sheehan, M. C. In Diatoms and Lake Acidity (eds J. P. Smol et al.) 251–274 (W. Junk, 1986).Whitehead, D. R. et al. The developmental history of Adirondack (N.Y.) lakes. J. Paleolimnol. 2, 185–206 (1989).ADS 
    Article 

    Google Scholar 
    Whitehead, D. R., Charles, D. F. & Goldstein, R. A. The PIRLA project (Paleoecological Investigation of Recent Lake Acidification): an introduction to the synthesis of the project. J. Paleolimnol. 3, 187–194 (1990).ADS 
    Article 

    Google Scholar 
    Dixit, S. S. et al. Diatom assemblages from Adirondack lakes (New York, USA) and the development of inference models for retrospective environmental assessment. J. Paleolimnol. 8, 27–47 (1993).ADS 
    Article 

    Google Scholar 
    Dixit, S. S. & Smol, J. P. Diatom evidence of past water quality changes in Adirondack seepage lakes (New York, USA). Diatom Res. 1, 113–129 (1995).Article 

    Google Scholar 
    Allen, A. P. et al. Concordance of taxonomic composition patterns across multiple lake assemblages: effects of scale, body size, and land use. Can. J. Fish. Aquat. 56, 2029–2040 (1999).Article 

    Google Scholar 
    Pither, J. & Aarssen, L. W. The evolutionary species pool hypothesis and patterns of freshwater diatom diversity along a pH gradient. J. Biogeogr. 32, 503–513 (2005).Article 

    Google Scholar 
    Winegardner, A. K., Legendre, P., Beisner, B. E. & Gregory-Eaves, I. Diatom diversity patterns over the past c. 150 years across the conterminous United States of America: Identifying mechanisms behind beta diversity. Global Ecol. Biogeogr. 26, 1303–1315 (2017).Article 

    Google Scholar 
    Dixit, S. S. & Smol, J. P. Diatoms as indicators in the Environmental Monitoring and Assessment Program-Surface Waters (EMAP-SW). Environ. Monit. Assess. 31, 275–37 (1994).PubMed 

    Google Scholar 
    Dixit, S. S. et al. Assessing water quality changes in the lakes of the northeastern United States using sediment diatoms. Can. J. Fish. Aquatic Sci. 56, 131–152 (1999).Article 

    Google Scholar 
    Stevenson, R. J., Zalack, J. & Wolin, J. A multimetric index of lake diatom condition using surface sediment assemblages. Freshw. Sci. 32, 1005–1025 (2013).Article 

    Google Scholar 
    Liu, B. & Stevenson, R. J. Improving assessment accuracy for lake biological condition by classifying lakes with diatom typology, varying metrics and modeling multimetric indices. Sci. Total Environ. 609, 263–271 (2017).ADS 
    Article 

    Google Scholar 
    Herlihy, A. T. et al. Using multiple approaches to develop nutrient criteria for lakes in the conterminous USA. Freshw. Sci. 32, 367–384 (2013).Article 

    Google Scholar 
    Bachmann, R. W., Hoyer, M. V. & Canfield, D. E. The extent that natural lakes in the United States of America have been changed by cultural eutrophication. Limnol. Oceanogr. 58, 945–950 (2013).ADS 
    Article 

    Google Scholar 
    McDonald, C. P. et al. Comment on Bachmann et al. (2013): A nonrepresentative sample cannot describe the extent of cultural eutrophication of natural lakes in the United States. Limnol. Oceanogr. 59, 2226–2230 (2014).ADS 
    Article 

    Google Scholar 
    Smith, V. H. et al. Comment: Cultural eutrophication of natural lakes in the United States is real and widespread. Limnol. Oceanogr. 59, 2217–2225 (2014).ADS 
    Article 

    Google Scholar 
    Bachmann, R. W., Hoyer, M. V. & Canfield, D. E. Response to comments: Quantification of the extent of cultural eutrophication of natural lakes in the United States. Limnol. Oceanogr. 59, 2231–2239 (2014).ADS 
    Article 

    Google Scholar 
    Bachmann, R. W., Hoyer, M. V., Croteau, A. C. & Canfield, D. E. Factors related to Secchi depths and their stability over time as determined from a probability sample of US lakes. Environ. Monit. Assess. 189, 206 (2017).Article 

    Google Scholar 
    Stager, J. C., Leavitt, P. R. & Dixit, S. S. Assessing impacts of past human activity on the water quality of Upper Saranac lake, New York. Lake Reserv. Manag. 13, 175–184 (1997).Article 

    Google Scholar 
    Dixit, S. S., Dixit, A. S., Smol, J. P., Hughes, R. M. & Paulsen, S. G. Water Quality Changes from Human Activities in Three Northeastern USA Lakes. Lake Reserv. Manag. 16, 35–321 (2000).Article 

    Google Scholar 
    Köster, D. et al. Paleolimnological assessment of human-induced impacts on Walden Pond (Massachusetts, USA) using diatoms and stable isotopes. Aquat. Ecosyst. Health 8, 117–131 (2005).Article 

    Google Scholar 
    Enache, M. D., Charles, D. F., Belton, T. J. & Callinan, C. W. Total phosphorus changes in New York and New Jersey lakes (USA) inferred from sediment cores. Lake Reserv. Manag. 28, 293–310 (2012).Article 

    Google Scholar 
    Rowell, H. C. et al. Quantitative paleolimnological inference models applied to a high-resolution biostratigraphic study of lake degradation and recovery, Onondaga Lake, New York (USA). J Paleolimnol. 55, 241–258 (2016).Article 

    Google Scholar 
    Tyree, M. A., Bishop, I. W., Hawkins, C. P., Mitchell, R. & Spaulding, S. A. Reduction of taxonomic bias in diatom species data. Limnol. Oceanogr. Methods 18, 271–279 (2020).Article 

    Google Scholar 
    Stribling, J. B., Pavlik, K. L., Holdsworth, S. M. & Leppo, E. W. Data quality, performance, and uncertainty in taxonomic identification for biological assessments. J. North Am. Benthol. Soc. 27, 906–919 (2008).Article 

    Google Scholar 
    Thomson, S. A. et al. Towards a global list of accepted species II. Consequences of inadequate taxonomic list governance. Org. Divers. Evol. 21, 623–630 (2021).Article 

    Google Scholar 
    Spaulding, S. A. et al. Diatoms of North America https://diatoms.org/ (2020).Lee, S. S., Bishop, I. W., Spaulding, S. A., Mitchell, R. M. & Yuan, L. L. Taxonomic harmonization may reveal a stronger association between diatom assemblages and total phosphorus in large datasets. Ecol. Indic. 102, 166–174 (2019).Article 

    Google Scholar 
    Cumming, B. F. et al. How Much Acidification Has Occurred in Adirondack Region Lakes (New York, USA) since Preindustrial Times? Can. J. Fish. Aquat. 49, 128–141 (1992).Article 

    Google Scholar 
    Larsen, D. P., Stevens, D. L., Selle, A. R. & Paulsen, S. G. Environmental Monitoring and Assessment Program, EMAP-Surface Waters: A northeast lakes pilot. Lake Reserv. Manag. 7, 1–11 (1991).Article 

    Google Scholar 
    Hughes, R. M., Paulsen, S. G. & Stoddard, J. L. EMAP-surface waters: A multiassemblage, probability survey of ecological integrity in the USA. Hydrobiologia 422, 429–443 (2000).Article 

    Google Scholar 
    Larsen, D. P., Thornton, K. W., Urquhart, N. S. & Paulsen, S. G. The role of sample surveys for monitoring the condition of the nation’s lakes. Environ. Monit. Assess. 32, 101–34 (1994).Article 

    Google Scholar 
    U.S. Environmental Protection Agency. Environmental Monitoring & Assessment Program. Northeast Lakes 1991-94 Data Sets. https://archive.epa.gov/emap/archive-emap/web/html/nelakes.html (2016).U.S. Environmental Protection Agency. National Lakes Assessment: A Collaborative Survey of the Nation’s Lakes. Report No. EPA-841-R-09-001. (U.S. Environmental Protection Agency, 2009).U.S. Environmental Protection Agency. 2012 National Lakes Assessment. Field Operations Manual. Report No. EPA 841-B-11-003. (U.S. Environmental Protection Agency, 2011)Charles, D. F., Knowles, C. & Davis, R. S. Protocols for the Analysis of Algal Samples Collected as Part of the U.S. Geological Survey National Water-Quality Assessment Program. https://water.usgs.gov/nawqa/protocols/algprotocol/algprotocol.pdf Report (2002).Krammer, K. Diatoms of Europe V. 1. (Gantner Verlag, 2000)Lange-Bertalot, H. Diatoms of Europe V. 2. (Gantner Verlag, 2001)Krammer, K. Diatoms of Europe V. 3. (Gantner Verlag, 2002)Krammer, K. Diatoms of Europe V. 4. (Gantner Verlag, 2003)Siver, P. A. & Hamilton, P. B. Iconographia Diatomologica V. 22. (Gantner Verlag, 2011).Levkov, Z., Metzeltin, D. & Pavlov, A. Diatoms of Europe V. 7. (Gantner Verlag, 2013)Levkov, Z., Mitić-Kopanja, D. & Reichardt, E. Diatoms of Europe V. 8. (Koeltz Botanical Books, 2016).Lange-Bertalot, H., Hofmann, G., Werum, M. & Cantonati, M. Freshwater Benthic Diatoms of Central Europe (Koeltz Botanical Books, 2017).Guiry, M. D. & Guiry, G. M. AlgaeBase https://www.algaebase.org (2021).Kociolek, J. P. et al. DiatomBase http://www.diatombase.org (2021).De Cáceres, M. Package ‘indicspecies’ https://cran.r-project.org/web/packages/indicspecies/indicspecies.pdf (2020).Legendre, P. & Birks, H. J. B. In Tracking Environmental Change Using Lake Sediments. V. 5: Data Handling and Numerical Techniques (eds Birks H. J. B. et al.) 201–248 (Springer Dordrecht, 2012).Legendre, P. & Gallagher, E. D. Ecologically meaningful transformations for ordination of species data. Oecologia 129, 271–280 (2001).ADS 
    Article 

    Google Scholar 
    Oksanen, J. et al. Package ‘vegan’ https://cran.r-project.org/web/packages/vegan/vegan.pdf (2020).Spaulding, S. A. Diatom Laboratory: Research Labs & Groups: INSTAAR: CU-Boulder https://instaar.colorado.edu/research/labs-groups/diatom-laboratory//research-detail (2021).Conservation Gateway. Northeast Lake and Pond Classification System. http://www.conservationgateway.org/ConservationByGeography/NorthAmerica/UnitedStates/edc/reportsdata/freshwater/Pages/Northeast-Lakes.aspx (2021).Soranno, P. & Cheruvelil, K. LAGOS-NE-LIMNO v1.087.3: A module for LAGOS-NE, a multi-scaled geospatial and temporal database of lake ecological context and water quality for thousands of U.S. Lakes: 1925–2013. Environmental Data Initiative https://doi.org/10.6073/pasta/08c6f9311929f4874b01bcc64eb3b2d7 (2019).U.S. Geological Survey. National Hydrography Dataset (NHD). USGS Unnumbered Series. (U.S. Geological Survey, 2001).Potapova, M. G., Lee, S. S., Spaulding, S. A. & Schulte, N. O. A harmonized dataset of sediment diatoms from hundreds of lakes in the northeastern United States. U.S. EPA Office of Research and Development (ORD) https://doi.org/10.23719/1524246 (2022).U.S. Environmental Protection Agency. National Aquatic Resource Surveys. National Lakes Assessment 2007 (data and metadata files) https://www.epa.gov/national-aquatic-resource-surveys/data-national-aquatic-resource-surveys (2010).U.S. Environmental Protection Agency. National Aquatic Resource Surveys. National Lakes Assessment 2017 (data and metadata files). http://www.epa.gov/national-aquatic-resource-surveys/data-national-aquatic-resource-surveys (2021). More

  • in

    Genetic, maternal, and environmental influences on sociality in a pedigreed primate population

    Study subjectsSubjects in our study are individually recognized wild capuchins found in and around the Lomas Barbudal Biological Reserve in Guanacaste, Costa Rica. This population has been under observation since 1990 (Perry 2012; Perry et al. 2012), including near continuous observation from January 2002 through March 2020.Data collectionWe use proximity data on subjects collected during group scan sampling between January 2001 and March 2020 (Altmann 1974). Included in scans are the identity of the subject, and the identity of other individuals within approximately 4 meters of them. Scans have been collected on all individuals in study groups since 2002, and on all adults and subadults since 2001. Scans are taken opportunistically, without regard to time of day. At least 10 min separate consecutive scans of the same individual to reduce the non-independence of scans taken close in time.Data in this manuscript were collected by 124 observers, with an average of 7.1 data collectors per month. Observers typically work in teams of two to three and rotate across different groups to reduce potential observer bias. Observers also rotate across observer teams to avoid observer drift in coding, since observer teams could potentially start to code behaviors differently from each other in the absence of overlap in observer composition.Initial pedigree constructionOf the 376 individuals in our behavioral dataset, 280 (74.5%) were first seen within three months of their births, and we could confidently assign maternity to them based on demographic (pregnancies) and behavioral data (primary nursing) even prior to genotyping. Of the remaining individuals, 41 (10.9%) were males of unknown origin that immigrated into our study population, while the rest were natal to our study groups but were first seen as older infants ( >3 months), juveniles, or (sub)adults (14.6%) and required genotyping to assign/confirm maternity. Paternity was assigned based on genetic information when possible (but see Non-genotyped individuals).In total, 287 subjects (76.3%) had two assigned parents, 37 had one assigned parent (9.8%), and 52 (13.8%) had no assigned parent based on demographic, behavioral, and/or genetic parentage information. Most individuals with no assigned parents were immigrant males (78.9%).GenotypingInformation on genetic parentage assignment (at up to 18 microsatellite loci) in our study population is available from previously published work (1996–2005 (Muniz et al. 2006), 2005–2012 (Godoy et al. 2016b)). Partial genotypes (up to 14 loci) have been generated for individuals in this study which more recently entered the study population through birth or immigration (n = 91, 2012–2020) (See SI File 1). Briefly, DNA was extracted primarily from non-invasively collected fecal samples, and occasionally from tissue samples obtained from deceased individuals, then amplified at up to 18 autosomal tetranucleotide microsatellite loci (Muniz and Vigilant 2008) using either a 1-step or 2-step PCR protocol (Arandjelovic et al. 2009). There were no significant deviations from Hardy-Weinberg equilibrium, and no evidence of linkage disequilibrium between loci was found (Muniz 2008).DNA samples were run at a minimum in triplicate, but additional PCRs were performed on low quality samples (e.g., with low quantities of DNA). Genotypes at each of the loci were assigned to be heterozygous when each allele was seen at least twice in independent PCRs, and assigned as homozygous when the allele was seen in at least three independent PCRs in absence of a second allele.Amplicons were analyzed using an ABI PRISM3100 automated sequencer and GeneMapper Software (Applied Biosystems, Foster City, CA, USA). Likelihood-based parentage assignments were performed using CERVUS 2.0 or 3.0 (Marshall et al. 1998; Kalinowski et al. 2007). The average exclusionary power of the 18 microsatellites was 0.9888 for the first parent and 0.9998 for second parent (Muniz et al. 2006).Individuals with unknown parents (e.g., immigrant males, founders) were genotyped twice (i.e., using two independent DNA samples) following the procedures described above to guard against sample mix up. Known mother-offspring pairs were confirmed by ascertaining the absence of Mendelian mismatches across all loci for the pair, though one mismatch was allowed to account for null alleles, mutations, and genotyping errors. We detected one null allele in the population in 19 individuals and traced it back to a male who was either the father or grandfather of those individuals (Muniz et al. 2006; Godoy et al. 2016b).Candidate males for paternity assignment were chosen based on group membership around the time of an infant’s conception (typically 1–10 males). In cases when conceptions occurred prior to the habituation of a study group, we used the identities of all adult males present when the group was first observed. Candidate mothers were similarly chosen for individuals that were first seen as older infants, juveniles, or (sub)adults. For individuals born post-group habituation, CERVUS has always assigned paternity from the pool of potential candidate fathers. Parent-offspring pairs and trios were allowed one mismatch (excluding those at the locus with the known null allele).Pedigree updatingNon-genotyped individualsDuring stable tenures, alpha males in our population sire approximately 73% of infants born in their groups, including 90% of offspring born to unrelated females (Godoy et al. 2016a). There is strong evidence of inbreeding avoidance between alpha males and their female descendants, with relatedness to females as the primary factor constraining alpha male monopolization of paternity within groups (Muniz et al. 2006, 2010; Godoy et al. 2016a, 2016b; Wikberg et al. 2017, 2018). We used this information to update our pedigree, filling missing father information with the identity of the alpha male around the time of a non-genotyped individual’s conception, but only if their mother was not the daughter or granddaughter of the alpha male (i.e., with inbreeding avoidance). This approach allowed us to assign presumed paternity to 21 non-genotyped individuals (5.6% of subjects) who were natal to our study groups.Individuals with missing or incomplete parentageOut of the original four study groups (from which fissions led to eight additional study groups), we lacked parentage information (i.e., neither parent was sampled) for 12 individuals first seen at the time of habituation. We had incomplete parentage on an additional 11 adults (i.e., only one parent was sampled). We used the software program COLONY version 2.0.6.7 to look for evidence of whether these individuals were related to each other at the level of full sibling (Jones and Wang 2010). We also looked for potential full sibling pairs among the non-natal immigrant males in the population, since co-migrant males are typically kin (Perry 2012; Wikberg et al. 2014, 2018). We assigned five full sibling pairs among co-migrant males, and four full sibling pairs among natal founders. For any remaining co-migrant males and natal founder pairs that were not assigned as full siblings, we assumed these to be either paternal (migrants) or maternal (natal) half siblings, as is typical in this study population (Perry 2012). These assignments are likely to have some error. However, based on what we know about kinship in capuchins, it would introduce more error to assume that these pairs are unrelated.We pruned our modified population pedigree using the R package pedantics version 1.01 (Morrissey and Wilson 2010), to include only individuals that were linked to the subjects in our behavioral dataset. The reduction in missing data can improve convergence and mixing of models (Hadfield 2010). The pruned pedigree contained 419 individuals, with 353 maternities, 354 paternities, 209 full sibships, 413 maternal half sibships, and 1496 paternal half sibships. Maximum pedigree depth was six generations (mean = 3.03).Sociality measures (response variables)We generated two related proximity-based measures of sociality—(1) whether an individual was seen in proximity of another monkey (within ~4 meters) during a scan (i.e., they were not alone), and (2) the number of partners an individual has nearby (within ~4 meters) during a scan. The former is measure of the propensity of an individual to be social versus alone, while the latter is more indicative of the gregariousness of an individual. These two phenotypes are not independent, as they are generated from the same data (Fig. 1a).Fig. 1: Distribution of sociality, sampling, group size, and alpha tenure length.The scatterplot in a shows the proportion of scans per individual per month where the subjects were recorded in proximity of others on the x-axis, and the average number of social partners per scan per month for subjects on the y-axis. The sizes of the circles in a are proportional to sample size (range: 5–317 scans per data point). The figure in b shows the number of calendar years of data sampling per subject (range: 1–20), c variation in group size, and d the number of calendar years represented by different alpha tenures in the dataset. Note that d does not represent the full diversity of alpha tenure lengths in the population, only within the dataset: some tenure lengths are left-truncated as data from 1990–2000 are not included in this dataset. Figure produced in R using ggplot2 version 3.3.5 (Wickham 2016) and cowplot version 1.1.1 (Wilke 2020). The capuchin image was generated in R using sketcher version 0.1.3 (Tsuda 2020) based on an image taken by Nicholas Schleissmann.Full size imageWe compiled the scans of individuals by month (mean: 31.9, range: 5–317 scans per month) so that we had counts of (1) the total number of scans where an individual was social and (2) the total number of partners an individual had. With these counts we could look at the (1) proportion of time spent social (versus alone) and (2) the average number of partners an individual had, while still preserving information about sampling density (number of scans).To be included in any month, subjects needed to have at least five scan samples in that period. As we are interested in the repeatability of our measures of social behavior, subjects had to have at least six months of data to be included.We excluded dependent infants (less than one year of age) as potential social partners of their mothers. We also excluded these dependent infants as subjects, since an infant is expected to be in close proximity of its mother, particularly during the first half of their first year of life (Godoy 2010; Perry 2012). Including data from infants would likely introduce upward bias to heritability estimates, because mothers and their dependent offspring (whom share high relatedness) would often be in close proximity of each other, and their measures of proximity to others would thus also be highly correlated.On average, subjects spent just over half of their sampled time within approximately four meters of another monkey (mean: 0.539, standard deviation: 0.193) and had approximately one social partner per scan (mean: 1.057, standard deviation: 0.619) (Fig. 1a). Our dataset consisted of 22,138 monthly sociality scores on 376 subjects generated from 641 140 scans (mean: 56.5 months per subject, range: 6–184 months per subject). Almost all subjects (99.7%, i.e., all but one) were represented by data across more than one calendar year (25, 50, 75% quantiles: 4, 7, 10 different years of data collection) (Fig. 1b).Fixed effectsWe included age (as a cubic function) and sex in our models, as well as their interaction to account for differences in how male and female capuchins sexually mature and age. Age in our dataset was right-skewed with higher representation at younger ages (mean: 9.3 years, standard deviation: 6.9) (Fig. 2). To put the ages in developmental context, mean age at first live birth is around 6.3 years for females in this population, though females can begin reproducing in their 5th year (Perry et al. 2012). Males as young as six years old have been known to sire offspring (Godoy et al. 2016b), but males tend to not reach full adult size until their 10th year (Jack et al. 2014).Fig. 2: Sociality as a function of age and sex.Circles represent individual monthly data. The sizes of the circles are proportional to sample size (range: 5–317 shows per data point). Circles in a represent the proportion of time individuals were seen in proximity of others (not alone) per month, while in b represent the average number of partners for individuals per month. Solid lines represent estimated sociality scores based on age and sex, with all other fixed effects set to the mean. The two x-axes represent age as z-scores and in years. Figure produced in R using ggplot2 version 3.3.5 (Wickham 2016).Full size imageSeasonal environmental changes, such as in food abundance, or temperature and rain, can lead to changes in how individuals cluster near others, for example, because of how food resources become distributed in the environment. For example, in black-crested gibbons (Nomascus concolor), group averages of dyadic proximity have been documented to decrease from the dry season to wet season, with increased average group proximity during cold months and lowered proximity during warm months (Guan et al. 2013). We account for seasonal variation by modelling monthly changes as a sine wave, through inclusion of the sine and cosine functions of a transformed month variable (See SI File 1 for further details).Central America is a region of ENSO-related precipitation, where the El Niño-Southern Oscillation (ENSO) has an impact on large scale patterns of temperature and precipitation (Ropelewski and Halpert 1987). Bimonthly rainfall anomalies are linked with both the warm El Niño and cool La Niña phases in a neighboring tropical dry forest in Costa Rica, where long-term monitoring of wild white-faced capuchins has shown declines in reproductive output associated with El Niño-like conditions (Campos et al. 2015). To account for the large-scale influence of ENSO on group dynamics, we included a factor variable for three different ENSO phases (Average/Neutral, Cool/La Niña, and Warm/El Niño). We used the bi-monthly Multivariate El Niño/Southern Oscillation (ENSO) index (MEI.v2) obtained from the Physical Sciences Laboratory of the National Oceanic and Atmospheric Administration (https://psl.noaa.gov/enso/mei/, retrieved: 2021-11-06) to determine the different phases. MEI.v2 is a composite index of five different variables (sea level pressure, sea surface temperature, surface zonal winds, surface meridional winds, and Outgoing Longwave Radiation) used to create a time series of ENSO conditions from 1979 to present (Zhang et al. 2019). Warm phases correspond to MEI.v2 values of 0.5 or higher, while cool phases correspond to values of −0.5 or lower.Demographic differences between groups and within groups across time can also lead to variation in behavior. For example, group size has been found to correlate with the amount of time that individuals spend grooming in various primate species (Dunbar 1991; Lehmann et al. 2007). Group size is also associated with higher sociality measures such as both the number of strong and weak ties that individuals form in diverse clades of primates (Schülke et al. 2022). We attempt to account for variation that arises from such demographic differences by including group size (mean: 24.7, standard deviation: 7.9) (Fig. 1c) as a fixed effect.In our models, group size and cubic age were centered and scaled to a mean of zero and a standard deviation of one.Random effectsAll models include the identity of the subject (VID, n = 376) as a random factor, as well as subject identity nested within year (VID:Year, n = 3150), the identity of each subject’s mother (VM, n = 142), maternal identity nested within group of residence within year of data collection (VM:GroupAlpha:Year, n = 2085), and a special variable known as the animal term to account for additive genetic variance (VA). These components contribute to long- and/or short-term repeatability of individuals. All models also include year of data collection (VYear, n = 20), month nested within year (VMonth:Year, n = 224), and the identity of each subject’s group of residence both across years (VGroupAlpha, n = 56) and within years (VGroupAlpha:Year, n = 200).VID in the models (since the models also additionally estimate VM and VA) can be thought of as estimating the “permanent environment variance” (i.e., VPE) of an individual, which is the “individual-specific variation in environmental conditions that permanently affect the phenotype (e.g. early-life conditions)” (Dingemanse et al. 2010). VID:Year captures the variance explained by the repeated sampling of the same individuals within a particular year. We use it to estimate the proportion of the phenotypic variance due to similarity in the trait within individuals from data taken closer in time (within the same year). During such a relatively short period, individuals are more likely to be stable in important social traits such as kin availability, dominance rank for adults, and maternal dominance rank for infants and young juveniles.VM estimates the variance explained by maternal effects (m2), specifically similarity between maternal siblings. Maternal identities were not available for all subjects, namely 11 immigrant males of unknown origin who were not assigned by COLONY as having a full sibling. We created unique dummy codes for their maternal identities, so that no two of these individuals shared the same mother. We additionally nested maternal identities (VM:GroupAlpha:Year) to account for similarity between maternal siblings residing in the same group in the same year. Such a nested structure might capture potential upward biases on heritability due to maternal kin biases in spatial association among siblings residing in the same group.We estimate h2 in our models by fitting a random effects term (VA), referred to as the animal term, which in the R package MCMCglmm links to the identities of individuals in our population pedigree (Hadfield 2010; see below for details on the implementation of the models in MCMCglmm). Inclusion of the animal term provides our models with an additive genetic variance component based on the estimated coefficients of relatedness between individuals in our pedigree. In short, if animals that share more alleles are also more like each other in their behavior, then variation in the behavior may well be due to genetic variation in the population (under the assumption that phenotypic similarity is not due to a shared environment, or is adequately controlled for by fixed and random effects in the model).VYear and VMonth:Year were included in order to account for temporal variation in sociality scores not captured by the fixed effects of seasonality or ENSO phase. These could arise from, for example, observer drift in coding (i.e., measurement error) or prevailing environmental conditions (e.g., drought) that could lead to changes to how individuals cluster near others. There were 218 unique observer combinations across the 224 months represented in the dataset, so VMonth:Year should also capture variance due to any differences between observer teams, though we cannot separate out the unique influence of observers.VGroupAlpha represents variance arising from the different alpha tenures within groups in our study population. VGroupAlpha captures both variance due to group of residence effects and the additional influence of alpha tenures within those groups. In capuchins, alpha males are ‘keystone’ individuals, whose influence is disproportionate relative to that of others in the population, and thus play important roles in establishing group dynamics (Jack and Fedigan 2018). Including group of residence, as defined by alpha tenure, is also important because it helps to account for the higher relatedness within groups within alpha tenures which results from high male reproductive skew toward alpha males. At Lomas Barbudal, males can remain in their alpha position for upwards of 18 years. Alpha tenures in this dataset spanned one to 14 years (Fig. 1d), so we additionally nested the identity of alpha males per group within years (VGroupAlpha:Year) so as to separate the within-year and across-year influences of group of residence.Statistical methodsWe ran analyses in R 4.1.2 (R Core Team 2021), using a Bayesian method with the R package MCMCglmm version 2.32 (Hadfield 2010). Data and code used to run all models is provided in the Supplementary Information.For our binary response variable (social versus alone), which was pooled into monthly units, we fit models with a binomial distribution and logit link function (family = “multinomial2”), with the number of scans each individual was documented social (‘successes’) versus the number of times alone (‘failures’).For our other response variable (number of partners), which was also pooled into monthly units, we fit models with a Poisson distribution (family = “poisson”), with the total number (sum) of partners per month. We included the natural log of the number of scans per month as a fixed effect to account for sampling effort. We set a strong prior for the log of sampling effort so that the rate at which events occurred was 1 (i.e., we could look at average number of partners per scan).We used a parameter-expanded prior (V = 1, nu = 1, alpha.mu = 0, alpha.V = 1000) and two inverse Wishart priors (V = 1, nu = 0.002; V = 1, nu = 0.02) for the G structures in our models (i.e., random effects variance components). The prior on the residual variance component was set to one for both the binomial and Poisson models. Estimates for variance components were robust against the choice of prior (SI Fig. 3). We therefore only report findings from models run with parameter-expanded priors in the main text.Pilot runs (thin = 10, burnin = 3000, nitt = 13,000) indicated that autocorrelation values would remain high for some variance components in models run with parameter-expanded priors, even with large thinning intervals. We therefore increased the number of iterations to guarantee effective sample sizes of at least 1000, but ideally closer to 4000. All models were run with a long burn-in period of at least 10,000 iterations.We ran multiple chains (n = 4) of each model and assessed convergence of the chains visually (SI Files 2a-b), as well as through the Gelman-Rubin criterion implemented via the ‘gelman.diag’ function from the coda package in R (version 0.19-4) (Plummer et al. 2006). Scale reduction factors were below 1.02, signifying good convergence. We used Heidelberger and Welch’s convergence diagnostic test for stationarity to check convergence of each chain using the ‘heidel.diag’ function from the coda package. Results are presented from the first chain of each model.Reduced modelsInclusion of fixed effects can potentially have an impact on the estimates of variance components in models because total phenotypic variance (VP) is estimated (and partitioned among the different random effects) after conditioning on the fixed effects. Heritability estimates, for example, can be higher because the variance explained by the fixed effects structure (VFE) is not included in VP, thus making the relative contribution of VA to VP larger compared to the same model without fixed effects (Wilson 2008). Conversely, not adequately controlling for relevant fixed effects that contribute to phenotypic variance among and within individuals may potentially lead to an underestimation of VA and associated heritability (h2).We ran multiple reduced versions of our models to look at the impact of fixed effects on our variance components. We began with an intercept-only version (i.e., no fixed effects), then built-up complexity by adding in versions with the properties of the individuals first (age, sex), then properties of the group (group size), and subsequently environmental properties (seasonality, ENSO phases). Outputs for these reduced models are provided in the Supplementary Information (SI Table 2, SI Table 3).We provide the deviance information criterion (DIC) values for models (automatically generated by the MCMCglmm package). DIC is a generalization for multi-level models of the Akaike Information Criterion (AIC); and as in AIC, lower DIC values indicate better fit.Transformations from unobserved latent scale to observed data scaleOutputs from our MCMCglmm models were on the unobserved latent scale. We used the R package QGglmm (version 0.7.4) to additionally compute parameters of interest on the observed data scale (de Villemereuil et al. 2016; de Villemereuil 2018). We used the functions ‘QCicc’ to compute Intra-Class Correlation (ICC) coefficients and ‘QGparams’ to compute additive genetic variance and thus narrow-sense heritability (h2) on the observed data scale. We implemented the ‘QGparams’ and ‘QGicc’ functions with parameters model = ‘binomN.logit’ and n.obs = 32 (the average number of scans per subject per month in our dataset) for the binomial model and model = ‘Poisson.log’ for the Poisson model. The choice of value for n.obs is somewhat arbitrary, and we show the consequences for changes in values of this parameter (i.e., higher estimates with increasing values of n.obs) in SI Fig. 4.Closed form solutions in QGglmm are not available for integrating over posterior distributions generated from binomial models with logit link functions (de Villemereuil 2016). Consequently, using the ‘QGicc’ function is particularly slow. We therefore estimate ICCs from our binomial models using a random subset of the posterior (n = 1000 iterations).The code used for transforming the MCMCglmm outputs from the latent scale to the original data scale are available online (see DATA AVAILABILITY).Repeatability and the proportion of variance explained by variance componentsTotal phenotypic variance (VP) was the sum of estimates from all variance components and residual variance in a model (VP = VID + VID:Year + VM + VM:GroupAlpha:Year + VA + VGroupAlpha + VGroupAlpha:Year + VMonth:Year + VYear + Vresidual). The proportion of variance explained by each variance component was calculated by including its estimate in the numerator while including total phenotypic variance in the denominator. So, for example the proportion of variance explained by year of data collection was calculated as (left( {frac{{V_{Year}}}{{V_P}}} right)).Long-term repeatability was calculated with the sum of VID, VM, and VA in the numerator. Short-term repeatability was calculated similarly but with inclusion of within-series variances (VID + VM + VA + VID:Year + VM:GroupAlpha:Year) in the numerator to capture additional consistency in among-individual differences resulting from greater environmental similarity within a time series (i.e., year).We report posterior modes and 95% Highest Posterior Density intervals (i.e., 95HPDI in square brackets). Unless mentioned otherwise, we present results on the unobserved latent scale, and without the variance from the fixed effects (VFE) incorporated into VP. For completeness, estimates with VFE included in VP and transformations to the observed data scale are also provided in SI Table 3. More

  • in

    Bacterial ectosymbionts in cuticular organs chemically protect a beetle during molting stages

    Wang C, Wang S. Insect pathogenic fungi: genomics, molecular interactions, and genetic improvements. Annu Rev Entomol. 2017;62:73–90.CAS 
    PubMed 
    Article 

    Google Scholar 
    Lacey LA, Grzywacz D, Shapiro-Ilan DI, Frutos R, Brownbridge M, Goettel MS. Insect pathogens as biological control agents: Back to the future. J Invertebr Pathol. 2015;132:1–41.CAS 
    PubMed 
    Article 

    Google Scholar 
    Butt TM, Coates CJ, Dubovskiy IM, Ratcliffe NA Entomopathogenic fungi: new insights into host-pathogen interactions. Advances in Genetics. 2016. Elsevier Ltd.Lu HL, St. Leger RJ. Insect immunity to entomopathogenic fungi. Adv Genet. 2016;94:251–85.CAS 
    PubMed 
    Article 

    Google Scholar 
    Yuan S, Tao X, Huang S, Chen S, Xu A. Comparative immune systems in animals. Annu Rev Anim Biosci. 2014;2:235–58.CAS 
    PubMed 
    Article 

    Google Scholar 
    Flórez LV, Biedermann PHW, Engl T, Kaltenpoth M. Defensive symbioses of animals with prokaryotic and eukaryotic microorganisms. Nat Prod Rep. 2015;32:904–36.PubMed 
    Article 

    Google Scholar 
    Oliver KM, Smith AH, Russell JA. Defensive symbiosis in the real world – advancing ecological studies of heritable, protective bacteria in aphids and beyond. Funct Ecol. 2014;28:341–55.Article 

    Google Scholar 
    Scarborough CL, Ferrari J, Godfray HC. Aphid protected from pathogen. Science 2005;310:1781.CAS 
    PubMed 
    Article 

    Google Scholar 
    Łukasik P, van Asch M, Guo H, Ferrari J, Charles H. Unrelated facultative endosymbionts protect aphids against a fungal pathogen. Ecol Lett. 2013;16:214–8.PubMed 
    Article 

    Google Scholar 
    Flórez LV, Scherlach K, Gaube P, Ross C, Sitte E, Hermes C, et al. Antibiotic-producing symbionts dynamically transition between plant pathogenicity and insect-defensive mutualism. Nat Commun. 2017;8:15172.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Flórez LV, Scherlach K, Miller IJ, Rodrigues A, Kwan JC, Hertweck C, et al. An antifungal polyketide associated with horizontally acquired genes supports symbiont-mediated defense in Lagria villosa beetles. Nat Commun. 2018;9:2478.PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Kaltenpoth M, Göttler W, Herzner G, Strohm E. Symbiotic bacteria protect wasp larvae from fungal infestation. Curr Biol. 2005;15:475–9.CAS 
    PubMed 
    Article 

    Google Scholar 
    Kroiss J, Kaltenpoth M, Schneider B, Schwinger MG, Hertweck C, Maddula RK, et al. Symbiotic streptomycetes provide antibiotic combination prophylaxis for wasp offspring. Nat Chem Biol. 2010;6:261–3.CAS 
    PubMed 
    Article 

    Google Scholar 
    Kaltenpoth M, Goettler W, Koehler S, Strohm E. Life cycle and population dynamics of a protective insect symbiont reveal severe bottlenecks during vertical transmission. Evol Ecol. 2010;24:463–77.Article 

    Google Scholar 
    Wang X, Yang X, Zhou F, Tian ZQ, Cheng J, Michaud JP, et al. Symbiotic bacteria on the cuticle protect the oriental fruit moth Grapholita molesta from fungal infection. Biol Control. 2022;169:104895.CAS 
    Article 

    Google Scholar 
    Wang L, Feng Y, Tian J, Xiang M, Sun J, Ding J, et al. Farming of a defensive fungal mutualist by an attelabid weevil. ISME J. 2015;9:1793–801.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Currie CR, Stuart AE. Weeding and grooming of pathogens in agriculture by ants. Proc R Soc B Biol Sci. 2001;268:1033–9.CAS 
    Article 

    Google Scholar 
    Currie CR, Scottt JA, Summerbell RC, Malloch D. Fungus-growing ants use antibiotic-producing bacteria to control garden parasites. Nature 1999;398:701–4.CAS 
    Article 

    Google Scholar 
    Currie CR, Bot ANM, Boomsma JJ. Experimental evidence of a tripartite mutualism: Bacteria protect ant fungus gardens from specialized parasites. Oikos 2003;101:91–102.Article 

    Google Scholar 
    Um S, Fraimout A, Sapountzis P, Oh D-CC, Poulsen M. The fungus-growing termite Macrotermes natalensis harbors bacillaene-producing Bacillus sp. that inhibit potentially antagonistic fungi. Sci Rep. 2013;3:3250.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Grubbs KJ, Surup F, Biedermann PHW, McDonald BR, Klassen JL, Carlson CM, et al. Cycloheximide-producing streptomyces associated with xyleborinus saxesenii and xyleborus affinis fungus-farming ambrosia beetles. Front Microbiol. 2020;11:1–12.Article 

    Google Scholar 
    Piel J. Metabolites from symbiotic bacteria. Nat Prod Rep. 2009;26:338–62.CAS 
    PubMed 
    Article 

    Google Scholar 
    Van Arnam EB, Currie CR, Clardy J. Defense contracts: Molecular protection in insect-microbe symbioses. Chem Soc Rev. 2018;47:1638–51.PubMed 
    Article 

    Google Scholar 
    Beemelmanns C, Guo H, Rischer M, Poulsen M. Natural products from microbes associated with insects. Beilstein J Org Chem. 2016;12:314–27.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Lackner G, Peters EE, Helfrich EJN, Piel J. Insights into the lifestyle of uncultured bacterial natural product factories associated with marine sponges. Proc Natl Acad Sci USA. 2017;114:E347–E356.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Schoenian I, Spiteller M, Ghaste M, Wirth R, Herz H, Spiteller D. Chemical basis of the synergism and antagonism in microbial communities in the nests of leaf-cutting ants. Proc Natl Acad Sci USA. 2011;108:1955–60.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Kaltenpoth M, Strupat K, Svatoš A. Linking metabolite production to taxonomic identity in environmental samples by (MA)LDI-FISH. ISME J. 2016;10:527–31.PubMed 
    Article 

    Google Scholar 
    Geier B, Sogin EM, Michellod D, Janda M, Kompauer M, Spengler B, et al. Spatial metabolomics of in situ host–microbe interactions at the micrometre scale. Nat Microbiol. 2020;5:498–510.CAS 
    PubMed 
    Article 

    Google Scholar 
    De Roode JC, Lefèvre T. Behavioral immunity in insects. Insects 2012;3:789–820.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Kerwin AH, Gromek SM, Suria AM, Samples RM, Deoss DJ, O’Donnell K, et al. Shielding the next generation: Symbiotic bacteria from a reproductive organ protect bobtail squid eggs from fungal fouling. mBio. 2019;10:e02376-19.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Soler JJ, Martín-Vivaldi M, Ruiz-Rodríguez M, Valdivia E, Martín-Platero AM, Martínez-Bueno M, et al. Symbiotic association between hoopoes and antibiotic-producing bacteria that live in their uropygial gland. Funct Ecol. 2008;22:864–71.Article 

    Google Scholar 
    Bunker ME, Elliott G, Martin MO, Arnold AE, Weiss SL. Vertically transmitted microbiome protects eggs from fungal infection and egg failure. Anim Microbiome. 2021;3:43.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Nyholm SV. In the beginning: Egg-microbe interactions and consequences for animal hosts: Egg microbiomes in animals. Philos Trans R Soc B Biol Sci. 2020;375:20190593.CAS 
    Article 

    Google Scholar 
    Smith DFQ, Dragotakes Q, Kulkarni M, Hardwick M, Casadevall A, Microbiology M, et al. Melanization is an important antifungal defense mechanism in Galleria mellonella hosts. bioRxiv 2022.04.02.486843.Yokoi K, Hayakawa Y, Kato D, Minakuchi C, Tanaka T, Ochiai M, et al. Prophenoloxidase genes and antimicrobial host defense of the model beetle, Tribolium castaneum. J Invertebr Pathol. 2015;132:190–200.CAS 
    PubMed 
    Article 

    Google Scholar 
    Zhang J, Huang W, Yuan C, Lu Y, Yang B, Wang CY, et al. Prophenoloxidase-mediated ex vivo immunity to delay fungal infection after insect ecdysis. Front Immunol. 2017;8:1–14.
    Google Scholar 
    Zhang J, Lu A, Kong L, Zhang Q, Ling E. Functional analysis of insect molting fluid proteins on the protection and regulation of ecdysis. J Biol Chem. 2014;289:35891–906.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Soluk DA. Postmolt susceptibility of ephemerella larvae to predatory stoneflies: constraints on defensive armour. Oikos 1990;58:336.Article 

    Google Scholar 
    Kanyile SN, Engl T, Kaltenpoth M. Nutritional symbionts enhance structural defence against predation and fungal infection in a grain pest beetle. J Exp Biol. 2022;225:1–9.Article 

    Google Scholar 
    Flórez LV, Kaltenpoth M. Symbiont dynamics and strain diversity in the defensive mutualism between Lagria beetles and Burkholderia. Environ Microbiol. 2017;19:3674–88.PubMed 
    Article 
    CAS 

    Google Scholar 
    Uberti A, Smaniotto MA, Giacobbo CL, Lovatto M, Lugaresi A, Girardi GC. Novo inseto praga na cultura do pessegueiro: biologia de Lagria villosa Fabricius, 1783 (Coleoptera: Tenebrionidae) alimentados com pêssego. Sci Electron Arch. 2017;10:72–76.
    Google Scholar 
    Stammer HJ. Die Symbiose der Lagriiden (Coleoptera). Z für Morphol und Ökologie der Tiere. 1929;15:1–34.Article 

    Google Scholar 
    Boucias DG, Pendland JC Principles of Insect Pathology. 1998. Springer Science + Business Media, LLC, New York.Garcia MA, Pierozzi IJ. Aspectos da biologia e ecologia de Lagria villosa Fabricius, 1781 (Coleoptera, Lagriidae). Rev Bras Biol. 1982;42:415–20.
    Google Scholar 
    Vega FE, Posada F, Catherine Aime M, Pava-Ripoll M, Infante F, Rehner SA. Entomopathogenic fungal endophytes. Biol Control. 2008;46:72–82.Article 

    Google Scholar 
    Kabaluk JT, Ericsson JD. Metarhizium anisopliae seed treatment increases yield of field corn when applied for wireworm control. Agron J. 2007;99:1377–81.Article 

    Google Scholar 
    Hallouti A, Ait Hamza M, Zahidi A, Ait Hammou R, Bouharroud R, Ait Ben Aoumar A, et al. Diversity of entomopathogenic fungi associated with Mediterranean fruit fly (Ceratitis capitata (Diptera: Tephritidae)) in Moroccan Argan forests and nearby area: impact of soil factors on their distribution. BMC Ecol. 2020;20:1–13.Article 
    CAS 

    Google Scholar 
    Iwanicki NS, Pereira AA, Botelho ABRZ, Rezende JM, Moral RDA, Zucchi MI, et al. Monitoring of the field application of Metarhizium anisopliae in Brazil revealed high molecular diversity of Metarhizium spp in insects, soil and sugarcane roots. Sci Rep. 2019;9:1–12.CAS 
    Article 

    Google Scholar 
    Roberts DW, St. Leger RJ. Metarhizium spp., cosmopolitan insect-pathogenic fungi: Mycological aspects. Adv Appl Microbiol. 2004;54:1–70.CAS 
    PubMed 
    Article 

    Google Scholar 
    Wierz JC, Gaube P, Klebsch D, Kaltenpoth M, Flórez LV. Transmission of bacterial symbionts with and without genome erosion between a beetle host and the plant environment. Front Microbiol. 2021;12:715601.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Gillespie JP, Bailey AM, Cobb B, Vilcinskas A. Fungi as elicitors of insect immune responses. Arch Insect Biochem Physiol. 2000;44:49–68.CAS 
    PubMed 
    Article 

    Google Scholar 
    Ortiz-Urquiza A, Keyhani NO. Action on the surface: Entomopathogenic fungi versus the insect cuticle. Insects 2013;4:357–74.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Grizanova EV, Coates CJ, Dubovskiy IM, Butt TM. Metarhizium brunneum infection dynamics differ at the cuticle interface of susceptible and tolerant morphs of Galleria mellonella. Virulence 2019;10:999–1012.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Eaton WD, Love DC, Botelho C, Meyers TR, Imamura K, Koeneman T. Preliminary results on the seasonality and life cycle of the parasitic dinoflagellate causing bitter crab disease in Alaskan Tanner crabs (Chionoecetes bairdi). J Invertebr Pathol. 1991;57:426–34.CAS 
    PubMed 
    Article 

    Google Scholar 
    Field RH, Chapman CJ, Taylor AC, Neil DM, Vickerman K. Infection of the Norway lobster Nephrops norvegicus by a Hematodinium-like species of dinoflagellate on the west coast of Scotland. Dis Aquat Organ. 1992;13:1–15.Article 

    Google Scholar 
    Threlkeld ST, Chiavelli DA, Willey RL. The organization of zooplankton epibiont communities. Trends Ecol Evol. 1993;8:317–21.CAS 
    PubMed 
    Article 

    Google Scholar 
    Duneau D, Ebert D. The role of moulting in parasite defence. Proc R Soc B Biol Sci. 2012;279:3049–54.Article 

    Google Scholar 
    Vandenberg JD, Ramos M, Altre JA. Dose-Response and Age- and Temperature-Related Susceptibility of the Diamondback Moth (Lepidoptera: Plutellidae) to Two Isolates of Beauveria bassiana (Hyphomycetes: Moniliaceae). Environ Entomol. 1998;27:1017–21.Article 

    Google Scholar 
    Vey A, Fargues J. Histological and ultrastructural studies of Beauveria bassiana infection in Leptinotarsa decemlineta larvae during ecdysis. J Invertebr Pathol. 1977;30:207–15.Article 

    Google Scholar 
    Reynolds SE, Samuels RI. Physiology and biochemistry of insect moulting fluid. Adv Insect Phys. 1996;26:157–232.CAS 
    Article 

    Google Scholar 
    Lopanik NB. Chemical defensive symbioses in the marine environment. Funct Ecol. 2014;28:328–40.Article 

    Google Scholar 
    Sen R, Ishak HD, Estrada D, Dowd SE, Hong E, Mueller UG. Generalized antifungal activity and 454-screening of Pseudonocardia and Amycolatopsis bacteria in nests of fungus-growing ants. Proc Natl Acad Sci USA. 2009;106:17805–10.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Currie CR, Poulsen M, Mendenhall J, Boomsma JJ, Billen J. Coevolved crypts and exocrine glands support mutualistic bacteria in fungus-growing ants. Science 2006;311:81–3.CAS 
    PubMed 
    Article 

    Google Scholar 
    Li H, Sosa-Calvo J, Horn HA, Pupo MT, Clardy J, Rabeling C, et al. Convergent evolution of complex structures for ant-bacterial defensive symbiosis in fungus-farming ants. Proc Natl Acad Sci USA. 2018;115:10720–5.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Kaltenpoth M, Roeser-Mueller K, Koehler S, Peterson A, Nechitaylo TY, Stubblefield JW, et al. Partner choice and fidelity stabilize coevolution in a Cretaceous-age defensive symbiosis. Proc Natl Acad Sci. 2014;111:6359–64.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Engl T, Kroiss J, Kai M, Nechitaylo TY, Svatoš A, Kaltenpoth M. Evolutionary stability of antibiotic protection in a defensive symbiosis. Proc Natl Acad Sci USA. 2018;115:E2020–E2029.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Gil-Turnes MS, Hay ME, Fenical W. Symbiotic marine bacteria chemically defend crustacean embryos from a pathogenic fungus. Science 1989;246:116–8.CAS 
    PubMed 
    Article 

    Google Scholar 
    Gil-Turnes MS, Fenical W. Embryos of Homarus americanus are protected by epibiotic bacteria. Biol Bull. 1992;182:105–8.CAS 
    PubMed 
    Article 

    Google Scholar 
    Hoffmann KH Insect Molecular Biology and Ecology. 2015. CRC Press.Eisner T, Morgan RC, Attygalle AB, Smedley SR, Herath KB, Meinwald J. Defensive production of quinoline by a phasmid insect (Oreophoetes peruana). J Exp Biol. 1997;200:2493–2500.CAS 
    PubMed 
    Article 

    Google Scholar 
    Waterworth SC, Flórez LV, Rees ER, Hertweck C, Kaltenpoth M, Kwan JC. Horizontal gene transfer to a defensive symbiont with a reduced genome in a multipartite beetle microbiome. mBio. 2020;11:e02430-19.PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Niehs SP, Kumpfmüller J, Dose B, Little RF, Ishida K, Flórez LV, et al. Insect‐associated bacteria assemble the antifungal butenolide gladiofungin by non‐canonical polyketide chain termination. Angew Chem. 2020;132:23322–6.Article 

    Google Scholar 
    Dose B, Niehs SP, Scherlach K, Flórez LV, Kaltenpoth M, Hertweck C. Unexpected bacterial origin of the antibiotic icosalide: two-tailed depsipeptide assembly in multifarious Burkholderia symbionts. ACS Chem Biol. 2018;13:2414–20.CAS 
    PubMed 
    Article 

    Google Scholar 
    Parada AE, Needham DM, Fuhrman JA. Every base matters: Assessing small subunit rRNA primers for marine microbiomes with mock communities, time series and global field samples. Environ Microbiol. 2016;18:1403–14.CAS 
    PubMed 
    Article 

    Google Scholar 
    Caporaso JG, Lauber CL, Walters WA, Berg-Lyons D, Huntley J, Fierer N, et al. Ultra-high-throughput microbial community analysis on the Illumina HiSeq and MiSeq platforms. ISME J. 2012;6:1621–4.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Caporaso JG, Lauber CL, Walters WA, Berg-Lyons D, Lozupone CA, Turnbaugh PJ, et al. Global patterns of 16S rRNA diversity at a depth of millions of sequences per sample. Proc Natl Acad Sci USA. 2011;108:4516–22.CAS 
    PubMed 
    Article 

    Google Scholar 
    Callahan BJ, McMurdie PJ, Rosen MJ, Han AW, Johnson AJA, Holmes SP. DADA2: High-resolution sample inference from Illumina amplicon data. Nat Methods. 2016;13:581–3.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Quast C, Pruesse E, Yilmaz P, Gerken J, Schweer T, Yarza P, et al. The SILVA ribosomal RNA gene database project: Improved data processing and web-based tools. Nucleic Acids Res. 2013;41:590–6.Article 
    CAS 

    Google Scholar 
    Yilmaz P, Parfrey LW, Yarza P, Gerken J, Pruesse E, Quast C, et al. The SILVA and ‘all-species Living Tree Project (LTP)’ taxonomic frameworks. Nucleic Acids Res. 2014;42:643–8.Article 
    CAS 

    Google Scholar 
    Weiss B, Kaltenpoth M. Bacteriome-localized intracellular symbionts in pollen-feeding beetles of the genus Dasytes (Coleoptera, Dasytidae). Front Microbiol. 2016;7:1–10.Article 

    Google Scholar 
    Amann RI, Binder BJ, Olson RJ, Chisholm SW, Devereux R, Stahl DA. Combination of 16S rRNA-targeted oligonucleotide probes with flow cytometry for analyzing mixed microbial populations. Appl Environ Microbiol. 1990;56:1919–25.CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Paschke C, Leisner A, Hester A, Maass K, Guenther S, Bouschen W, et al. Mirion – A software package for automatic processing of mass spectrometric images. J Am Soc Mass Spectrom. 2013;24:1296–306.CAS 
    PubMed 
    Article 

    Google Scholar  More

  • in

    The Holocene temperature conundrum answered by mollusk records from East Asia

    Jiang, D. B., Lang, X. M., Tian, Z. P. & Wang, T. Considerable model-data mismatch in temperature over China during the mid-Holocene: results of PMIP simulations. J. Clim. 25, 4135–4153 (2012).ADS 
    Article 

    Google Scholar 
    Marcott, S. A., Shakun, J. D., Clark, P. U. & Mix, A. C. A reconstruction of regional and global temperature for the past 11300 years. Science 339, 1198–1201 (2013).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Liu, Z. et al. The Holocene temperature conundrum. Proc. Natl Acad. Sci. USA 111, E3501–E3505 (2014).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Marsicek, J., Shuman, B., Bartlein, P., Shafer, S. L. & Brewer, S. Reconciling divergent trends and millennial variations in Holocene temperatures. Nature 554, 92–96 (2018).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Affolter, S., Huselmann, A., Fleitmann, D., Edwards, R. L. & Leuenberger, M. Central Europe temperature constrained by speleothem fluid inclusion water isotopes over the past 14,000 years. Sci. Adv. 5, eaav3809 (2019).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Bova, S., Rosenthal, Y., Liu, Z. & Yan, M. Seasonal origin of the thermal maxima at the Holocene and the last interglacial. Nature 589, 548–553 (2021).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Meyer, H. et al. Long-term winter warming trend in the Siberian Arctic during the mid- to late Holocene. Nat. Geosci. 8, 122–125 (2015).ADS 
    CAS 
    Article 

    Google Scholar 
    Baker, J., Lachniet, M., Chervyatsova, O., Asmerom, Y. & Polyak, V. J. Holocene warming in western continental Eurasia driven by glacial retreat and greenhouse forcing. Nat. Geosci. 10, 430–435 (2017).ADS 
    CAS 
    Article 

    Google Scholar 
    Mann, M., Schmidt, G., Miller, S. & LeGrande, A. Potential biases in inferring Holocene temperature trends from long-term borehole information. Geophys. Res. Lett. 36, L05708 (2009).ADS 
    Article 

    Google Scholar 
    Liu, T. S. Loess and the Environment (In Chinese). (China Ocean Press, Beijing, 1985).Rousseau, D. D. & Wu, N. Q. A new molluscan record of the monsoon variability over the past 130 000 yr in the Luochuan loess sequence, China. Geology 25, 275–278 (1997).ADS 
    Article 

    Google Scholar 
    Wu, N. Q., Li, F. J. & Rousseau, D. D. Terrestrial mollusk records from Chinese loess sequences and changes in the East Asian monsoonal environment. J. Asian Earth Sci. 155, 35–48 (2018).ADS 
    Article 

    Google Scholar 
    Qian, L. Q. Climate of Loess Plateau (in Chinese). (China Meteorological Press, Beijing, 1991).Chen, D. & Gao, J. Economic Fauna Sinica of China: Terrestrial Mollusca (in Chinese). (Science Press, Beijing, 1987).Proćków, M., Drvotová, M., Juřičková, L. & Kuźnik-Kowalska, E. Field and laboratory studies on the life-cycle, growth and feeding preference in the hairy snail Trochulus hispidus (L., 1758) (Gastropoda: Pulmonata: Hygromiidae). Biologia 68, 131–141 (2013).Article 

    Google Scholar 
    Rousseau, D. Climatic transfer function from Quaternary molluscs in European loess deposits. Quat. Res. 36, 195–209 (1991).Article 

    Google Scholar 
    Rousseau, D., Preece, R. & Limondin-Lozouet, N. British late glacial and Holocene climatic history reconstructed from land snail assemblages. Geology 26, 651–654 (1998).ADS 
    Article 

    Google Scholar 
    Wang, Z. H., Fang, J. Y., Tang, Z. Y. & Lin, X. Patterns, determinants and models of woody plant diversity in China. Proc. R. Soc. B 278, 2122–2132 (2011).PubMed 
    Article 

    Google Scholar 
    Gu, Z. Y., Liu, Z. X., Xu, B. & Wu, N. Q. Stable carbon and oxygen isotopes in land snail carbonate shells from a last glacial loess sequence and their implications of environmental changes (in Chinese). Quat. Sci. 29, 13–22 (2009).CAS 

    Google Scholar 
    Sun, X. H., Gu, Z. Y. & Xu, B. Oxygen isotopic variations in the shells collected monthly from a live species of land snails at local in Zhenjiang, Jiangsu Province, China (in Chinese). Quat. Sci. 29, 976–980 (2009).CAS 

    Google Scholar 
    Huang, L., Wu, N., Gu, Z. & Chen, X. Variability of snail growing season at the Chinese Loess Plateau during the last 75 ka. Chin. Sci. Bull. 57, 1036–1045 (2012).CAS 
    Article 

    Google Scholar 
    Dong, Y. J. et al. Paleorecords reveal the increased temporal instability of species diversity under biodiversity loss. Quat. Sci. Rev. 269, 107147 (2021).Article 

    Google Scholar 
    Horsák, M. Mollusc assemblages in palaeoecological reconstructions: an investigation of their predictive power using transfer function models. Boreas 40, 459–467 (2011).Article 

    Google Scholar 
    Sümegi, P. & Gulyás, S. Some notes on the interpretation and reliability of malacological proxies in paleotemperature reconstructions from loess- comments to Obreht et al.‘s “A critical reevaluation of paleoclimate proxy records from loess in the Carpathian Basin”. Earth-Sci. Rev. 221, 103675 (2021).Samartin, S. et al. Warm Mediterranean mid-Holocene summers inferred from fossil midge assemblages. Nat. Geosci. 10, 207–212 (2017).ADS 
    CAS 
    Article 

    Google Scholar 
    Seppä, H., Birks, H., Odland, A., Poska, A. & Veski, S. A modern pollen-climate calibration set from northern Europe: developing and testing a tool for palaeoclimatological reconstructions. J. Biogeogr. 31, 251–267 (2004).Article 

    Google Scholar 
    Allen, J. et al. Rapid environmental changes in southern Europe during the last glacial period. Nature 400, 740–743 (1999).ADS 
    CAS 
    Article 

    Google Scholar 
    Rioual, P. et al. High-resolution record of climate stability in France during the last interglacial period. Nature 413, 293–296 (2001).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Salonen, J. S. et al. Abrupt high-latitude climate events and decoupled seasonal trends during the Eemian. Nat. Commun. 9, 2851 (2018).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Lu, H. et al. Seasonal climatic variation recorded by phytolith assemblages from Baoji loess sequence in central China over the last 150000 a. Sci. China, Ser. D. 26, 629–639 (1996).
    Google Scholar 
    Telford, R. J. & Birks, H. J. B. A novel method for assessing the statistical significance of quantitative reconstructions inferred from biotic assemblages. Quat. Sci. Rev. 30, 1272–1278 (2011).ADS 
    Article 

    Google Scholar 
    Lu, H. Y., Wu, N. Q., Liu, K. B., Jiang, H. & Liu, T. S. Phytoliths as quantitative indicators for the reconstruction of past environmental conditions in China II: palaeoenvironmental reconstruction in the Loess Plateau. Quat. Sci. Rev. 26, 759–772 (2007).ADS 
    Article 

    Google Scholar 
    Sun, J. M., Diao, G. Y., Wen, Q. Z. & Zhou, H. Y. A preliminary study on quantitative estimate of Palaeoclimate by using geochemical transfer function in the Loess Plateau (In Chinese). Geochimica 28, 265–272 (1999).CAS 

    Google Scholar 
    Wen, R. et al. Pollen–climate transfer functions intended for temperate eastern Asia. Quat. Int. 311, 3–11 (2013).Article 

    Google Scholar 
    Xu, Q., Xiao, J., Li, Y., Tian, F. & Nakagawa, T. Pollen-based quantitative reconstruction of Holocene climate changes in the Daihai lake area, inner Mongolia, China. J. Clim. 23, 2856–2868 (2010).ADS 
    Article 

    Google Scholar 
    Li, J. et al. Quantitative Holocene climatic reconstructions for the lower Yangtze region of China. Clim. Dyn. 50, 1101–1113 (2018).Article 

    Google Scholar 
    Nakagawa, T., Tarasov, P. E., Nishida, K., Gotanda, K. & Yasuda, Y. Quantitative pollen-based climate reconstruction in central Japan: application to surface and Late Quaternary spectra. Quat. Sci. Rev. 21, 2099–2113 (2002).ADS 
    Article 

    Google Scholar 
    Chen, M.-T. et al. Dynamic millennial-scale climate changes in the northwestern Pacific over the past 40,000 years. Geophys. Res. Lett. 37, L23603 (2010).ADS 

    Google Scholar 
    Sun, Y., Oppo, D. W., Xiang, R., Liu, W. & Gao, S. Last deglaciation in the Okinawa Trough: Subtropical northwest Pacific link to Northern Hemisphere and tropical climate. Paleoceanography 20, PA4005 (2005).ADS 
    Article 

    Google Scholar 
    de Garidel-Thoron, T. et al. A multiproxy assessment of the western equatorial Pacific hydrography during the last 30 kyr. Paleoceanography 22, PA3204 (2007).ADS 

    Google Scholar 
    Chen, F., Duan, Y. & Hou, J. An 88 ka temperature record from a subtropical lake on the southeastern margin of the Tibetan Plateau (third pole): new insights and future perspectives. Sci. Bull. 66, 1056–1057 (2021).Article 

    Google Scholar 
    James, R. P. & Arguez, A. On the estimation of daily climatological temperature variance. J. Atmos. Ocean. Tech. 32, 2297–2304 (2015).Article 

    Google Scholar 
    Mearns, L. O., Rosenzweig, C. & Goldberg, R. Mean and variance change in climate scenarios: methods, agricultural applications, and measures of uncertainty. Climatic Change 35, 367–396 (1997).Article 

    Google Scholar 
    Laskar, J., Fienga, A., Gastineau, M. & Manche, H. La2010: a new orbital solution for the long-term motion of the Earth. Astron. Astrophys. 532, A89 (2011).ADS 
    MATH 
    Article 

    Google Scholar 
    Bader, J. et al. Global temperature modes shed light on the Holocene temperature conundrum. Nat. Commun. 11, 4726 (2020).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Liu, Y. et al. A possible role of dust in resolving the Holocene temperature conundrum. Sci. Rep. 8, 4434 (2018).ADS 
    PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Thompson, A. J., Zhu, J., Poulsen, C. J., Tierney, J. E. & Skinner, C. B. Northern Hemisphere vegetation change drives a Holocene thermal maximum. Sci. Adv. 8, eabj6535 (2022).PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Lambert, F. et al. The role of mineral-dust aerosols in polar temperature amplification. Nat. Clim. Chang. 3, 487–491 (2013).ADS 
    Article 

    Google Scholar 
    Xu, Y., Wang, H., Liao, H. & Jiang, D. Simulation of the direct radiative effect of mineral dust aerosol on the climate at the last glacial maximum. J. Clim. 24, 843–858 (2011).ADS 
    Article 

    Google Scholar 
    Wohlfahrt, J., Harrison, S. P. & Braconnot, P. Synergistic feedbacks between ocean and vegetation on mid- and high-latitude climates during the mid-Holocene. Clim. Dynam. 22, 223–238 (2004).ADS 
    Article 

    Google Scholar 
    Jahn, A., Claussen, M., Ganopolski, A. & Brovkin, V. Quantifying the effect of vegetation dynamics on the climate of the Last Glacial Maximum. Clim. Past 1, 1–7 (2005).Article 

    Google Scholar 
    Braconnot, P., Joussaume, S., Marti, O. & de Noblet, N. Synergistic feedbacks from ocean and vegetation on the African monsoon response to mid-Holocene insolation. Geophys. Res. Lett. 26, 2481–2484 (1999).ADS 
    Article 

    Google Scholar 
    Braconnot, P. et al. Evaluation of climate models using palaeoclimatic data. Nat. Clim. Change 2, 417–424 (2012).ADS 
    Article 

    Google Scholar 
    Zhang, X. & Chen, F. Non-trivial role of internal climate feedback on interglacial temperature evolution. Nature 600, E1–E3 (2021).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Li, F. J. et al. Quantitative distribution and calculation of ecological amplitude of land snail Metodontia in the Chinese Loess Plateau and adjacent regions (In Chinese with English abstract). Quat. Sci. 36, 564–574 (2016).
    Google Scholar 
    Dong, Y. J. et al. Influence of monsoonal water-energy dynamics on terrestrial mollusk species-diversity gradients in northern China. Sci. Total Environ. 676, 206–214 (2019).ADS 
    CAS 
    PubMed 
    Article 

    Google Scholar 
    Dong, Y. J. et al. Anthropogenic modification of soil communities in northern China for at least two millennia: Evidence from a quantitative mollusk approach. Quat. Sci. Rev. 248, 106579 (2020).Article 

    Google Scholar 
    Cameron, R. A. D. & Pokryszko, B. M. Estimating the species richness and composition of land mollusc communities: Problems, consequences and practical advice. J. Conchol. 38, 529–547 (2005).
    Google Scholar 
    Dong, Y., Wu, N., Li, F., Huang, L. & Wen, W. Time-transgressive nature of the magnetic susceptibility record across the Chinese Loess Plateau at the Pleistocene/Holocene transition. PLoS One 10, e0133541 (2015).PubMed 
    PubMed Central 
    Article 
    CAS 

    Google Scholar 
    Ter Braak, C. J. F. & Smilauer, P. CANOCO Reference Manual and CanoDraw for Windows User’s Guide: Software for Canonical Community Ordination (Version 4.5). (Microcomputer Power, New York, 2002).Ter Braak, C. J. F. & Smilauer, P. CANOCO Reference Manual and User’s Guide: Software for Ordination (Version 5.0). (Microcomputer Power, New York, 2012).Ter Braak, C. J. F. & Juggins, S. Weighted averaging partial least squares regression (WA-PLS): an improved method for reconstructing environmental variables from species assemblages. Hydrobiologia 269, 485–502 (1993).Article 

    Google Scholar 
    Birks, H. J. B., Lotter, A. F., Juggins, S. & Smol, J. P. Tracking Environmental Change Using Lake Sediments Volume 5: Data Handling and Numerical Techniques. p. 123–141. (Springer, London, 2012).Ter Braak, C. J. F. Canonical community ordination. Part I: Basic theory and linear methods. Ecoscience 1, 127–140 (1994).Article 

    Google Scholar 
    Juggins, S. C2 data analysis (version 1.7.4). (Newcastle University, Newcastle, 2011).Juggins, S. Rioja: Analysis of Quaternary Science Data. R package version 0.9-21 http://cran.r-project.org/package=rioja (2017).Simpson, G. L. & Oksanen, J. Analogue: Analogue matching and Modern Analogue. Technique Transfer Function Models. R package version 0.17-4 https://cran.r-project.org/package=analogue (2020).Telford, R. J. palaeoSig: Significance Tests of Quantitative Palaeoenvironmental Reconstructions. R package version 2.0-3 http://cran.r-project.org/package=palaeoSig (2019).Hansen, J., Sato, M. & Ruedy, R. Perception of climate change. Proc. Natl Acad. Sci. USA 109, E2415–E2423 (2012).ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Olonscheck, D., Schurer, A. P., Lücke, L. & Hegerl, G. C. Large-scale emergence of regional changes in year-to-year temperature variability by the end of the 21st century. Nat. Commun. 12, 7237 (2021).ADS 
    CAS 
    PubMed 
    PubMed Central 
    Article 

    Google Scholar 
    Liu, Y., Ren, G., Kang, H. & Sun, X. A significant bias of Tmax and Tmin average temperature and its trend. J. Appl. Meteorol. Clim. 58, 2235–2246 (2019).ADS 
    Article 

    Google Scholar 
    Parey, S., Dacunha-Castelle, D. & Hoang, T. T. H. Mean and variance evolutions of the hot and cold temperatures in Europe. Clim. Dyn. 34, 345–359 (2010).Article 

    Google Scholar 
    Dong, Y. SeaTemCon_R code for “The Holocene temperature conundrum answered by mollusk records from East Asia”. Zenodo https://doi.org/10.5281/zenodo.6426798 (2022).Article 

    Google Scholar 
    Dong, Y. Data repository for “The Holocene temperature conundrum answered by mollusk records from East Asia”. Zenodo https://doi.org/10.5281/zenodo.6426911 (2022).Article 

    Google Scholar  More

  • in

    Multiproxy study of 7500-year-old wooden sickles from the Lakeshore Village of La Marmotta, Italy

    Snir, A. et al. The origin of cultivation and proto-weeds, long before Neolithic farming. PLoS ONE 10(7), e0131422. https://doi.org/10.1371/journal.pone.0131422 (2015).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Groman-Yaroslavski, I., Weiss, E. & Nadel, D. Composite sickles and cereal harvesting methods at 23,000-years-old Ohalo II Israel. PLoS ONE 11(11), e0167151. https://doi.org/10.1371/journal.pone.0167151 (2016).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Edwards, P. C. A 14000 year-old hunter-gatherer’s toolkit. Antiquity 81(314), 865–876. https://doi.org/10.1017/S0003598X0009596X (2007).Article 

    Google Scholar 
    Le Dosseur, G. Bone Objects in the Southern Levant from the Thirteenth to the Fourth Millennia. Bulletin du Centre de recherche français à Jérusalem 12, 111–125 (2003).
    Google Scholar 
    Garrard, A., & Yazbeck, C. The Natufian of Moghr el-Ahwal in the Qadisha valley, northern Lebanon. in Natufian Foragers in the Levant. International Monographs in Prehistory (eds. Bar-Yosef, O. & Valla, F. R.). 17–47. (Michigan, Ann Arbor, 2013).Belfer-Cohen, A. The Natufian in the Levant. Annu. Rev. Anthropol. 20, 167–186. https://doi.org/10.1146/annurev.an.20.100191.001123 (1991).Article 

    Google Scholar 
    Stordeur, D. Le Natoufien et son évolution à travers les artefacts en os in Natufian Foragers in the Levant. International Monographs in Prehistory (eds. Bar-Yosef, O. & Valla, F. R.). 457–482. (Michigan, Ann Arbor, 2013).Rosen, S. A. Lithics after the Stone Age: a handbook of stone tools from the Levant. (Rowman Altamira, 1997).Anderson, P. C. Prehistory of agriculture: new experimental and ethnographic approaches. (Cotsen Institute of Archaeology Press, 1999).Ibáñez, J. J., González Urquijo, J. E., & Rodríguez, A. The evolution of technology during the PPN in the Middle euphrates. A view from use wear analysis of lithic tools. in Systèmes techniques et communautés du Néolithique Préceramique au Proche Orient. Technical Systems and Near Eastern PPN Communities (eds. Astruc, L., Binder, D. & Briois, F.) 153–165 (Editions APDCA, 2007).Maeda, O., Lucas, L., Silva, F., Tanno, K. I. & Fuller, D. Q. Narrowing the harvest: Increasing sickle investment and the rise of domesticated cereal agriculture in the Fertile Crescent. Quatern. Sci. Rev. 145, 226–237. https://doi.org/10.1016/j.quascirev.2016.05.032 (2016).ADS 
    Article 

    Google Scholar 
    Pichon, F. Exploitation of the cereals during the Pre-pottery Neolithic of Dja’de-el-Mughara: Preliminary results of the functional study of the glossy blades. Quatern. Int. 427, 138–151. https://doi.org/10.1016/j.quaint.2016.01.064 (2017).Article 

    Google Scholar 
    Borrell, F., & Molist, M. Projectile Points, Sickle Blades and Glossed Points. Tools and Hafting Systems at Tell Halula (Syria) during the 8th millennium cal. BC Paléorient, 33(2), 59–77 (2007). https://doi.org/10.2307/41496812.Douka, K., Efstratiou, N., Hald, M., Henriksen, P. & Karetsou, A. Dating Knossos and the arrival of the earliest Neolithic in the southern Aegean. Antiquity 91(356), 304–321. https://doi.org/10.15184/aqy.2017.29 (2017).Article 

    Google Scholar 
    Perlès, C. From the Near East to Greece: Let’s reverse the focus. Cultural elements that didn’t transfer. in How did farming reach Europe? (ed. Lichter, C.) 275–290 (Istanbul, Ege Yayınları, 2005).Gijn A.L. van & Wentink K. The role of flint in mediating identities: The microscopic evidence. in Mobilty, meaning & transformations of things, shifting contexts of material culture through time and space. (eds. Hahn, H.P. & Weiss, H.) 120–132 (Oxford, Oxbow Books, 2013).Guilaine, J. The neolithic transition: From the Eastern to the Western Mediterranean. in Times of Neolithic Transition along the Western Mediterranenn. (eds. O., García-Puchol & D. C., Salazar-García) 15–31 (New York, Springer, 2017). https://doi.org/10.1007/978-3-319-52939-4_2.Forenbaher, S. & Miracle, P. T. The spread of farming in the Eastern Adriatic. Antiquity 79(305), 514–528 (2005).Article 

    Google Scholar 
    Gabriele, M. et al. Long-distance mobility in the North-Western Mediterranean during the Neolithic transition using high resolution pottery sourcing. J. Archaeol. Sci. Rep. 28, 102050. https://doi.org/10.1016/j.jasrep.2019.102050 (2019).Article 

    Google Scholar 
    Manen, C., Perrin, T., Guilaine, J., Bouby, L., Bréhard, S., Briois, F., Durand, F., Marinval, P. & Vigne, J. D. The Neolithic transition in the western Mediterranean: A complex and non-linear diffusion process—the radiocarbon record revisited. Radiocarbon 61(2), 531–571 (2019). https://doi.org/10.1017/RDC.2018.98Ibáñez, J. J., Clemente Conte, I., Gassin, B., Gibaja, J. F., Gonzáles Urquijo, J. E., Márquez, B., Philibert, S., Rodríguez Rodríguez, A. Harvesting technology during the Neolithic in south-west Europe. in Prehistoric technology 40 years later: functional studies and the Russian legacy (eds. Longo L. & Skakun, N.) 183–95 (Oxford, Archaeopress, 2008).Gibaja, J. F., Ibáñez, J. J., González Urquijo, J. E. Neolithic Sickles in the Iberian Peninsula. in Exploring and Explaining Diversity in Agricultural Technology, EARTH 2 (eds. van Gijn, A., Whittaker, P. & Anderson, P.) 112–118 (Oxford, Oxbow Books, 2014).Mazzucco, N., Capuzzo, G., Petrinelli-Pannocchia, C., Ibáñez, J. J., Gibaja, J. F. Harvesting tools and the spread of the Neolithic into the Central-Western Mediterranean area. Quat. Int. 470(Part B), 511–528 (2018). https://doi.org/10.1016/j.quaint.2017.04.018.Mazzucco, N., Guilbeau, D., Kačar, S., Podrug, E., Forenbaher, S., Radić, D., Moore, A. T. M. The time is ripe for a change. The evolution of harvesting technologies in Central Dalmatia during the Neolithic period (6th millennium cal BC). J. Anthropol. Archaeol. 51, 88–103 (2018). https://doi.org/10.1016/j.jaa.2018.06.003Mazzucco, N. et al. Migration, adaptation, innovation: The spread of Neolithic harvesting technologies in the Mediterranean. PLoS ONE 15(4), e0232455. https://doi.org/10.1371/journal.pone.0232455 (2020).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Fugazzola Delpino, M. A., D’Eugenio, G. & Pessina, A. “La Marmotta” (Anguillara Sabazia, RM): Scavi 1989—un abitato perilacustre di età Neolitica. Bull. Paletnol. Ital. 84, 181–315 (1993).
    Google Scholar 
    Fugazzola Delpino, M. A., Pessina, A. Le village néolithique submergé de La Marmotta (lac de Bracciano, Rome). in Le Néolithique du Nord-Ouest méditerranéen (ed. Vaquer, J.) 35–38 (Société préhistorique française, Paris, 1999)Fugazzola Delpino, M. A. La Marmotta. in Le ceramiche impresse nel Neolitico antico. Italia e Mediterraneo (eds. Fugazzola, M.A., Pessina, A. & Tiné, V) 373–395 (Istituto Poligrafico e Zecca dello Stato, Roma, 2002).Grantham, G. L. faucille et la faux. Études rurales 151–152, 103–131 (1999).Article 

    Google Scholar 
    Sigaut, F. Identification des techniques de récolte des graines alimentaires. J. Agric. Trad. Bot. Appl. 25(3), 145–161 (1978).
    Google Scholar 
    Anderson, P. C., Sigaut, F. Introduction: reasons for variability in harvesting techniques and tools. in Exploring and Explaining Diversity in Agricultural Technology, EARTH 2 (eds. van Gijn, A., Whittaker, P. & Anderson, P.) 85–93 (Oxford, Oxbow Books, 2014).Halstead, P. Two oxen ahead: Pre-mechanized farming in the Mediterranean (John Wiley & Sons, 2014).Book 

    Google Scholar 
    Fugazzola Delpino, M. A. & Mineo, M. La piroga neolitica di Bracciano (La Marmotta 1). Bull. Paletnol. Ital. 86, 197–266 (1995).
    Google Scholar 
    Fugazzola Delpino, M. A., Tinazzi, O. Dati di cronologia da un villaggio del Neolitico Antico. Le indagini dendrocronologiche condotte sui legni de La Marmotta (lago di Bracciano-Roma). in Miscellanea in ricordo di Francesco Nicosia, 1–10 (Studia Erudita, Fabrizio Serra Editore, 2010).Salavert, A. et al. Direct dating reveals the early history of opium poppy in western Europe. Sci. Rep. 10, 20263. https://doi.org/10.1038/s41598-020-76924-3 (2020).ADS 
    CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Ghiselli, L. et al. Nutritional characteristics of ancient Tuscan varieties of Triticum aestivum L. Ital. J. Agron. 11(4), 237–245 (2016).Article 

    Google Scholar 
    Pichon, F. Une moisson expérimentale de céréales, Séranon (août 2016), ArchéOrient – Le Blog, 14 octobre2016, (2016). https://archeorient.hypotheses.org/6667.Banks, W. E. & Kay, M. High-resolution casts for lithic use-wear analysis. Lithic Technol. 28(1), 27–34. https://doi.org/10.1080/01977261.2003.11721000 (2003).Article 

    Google Scholar 
    Ibáñez, J. J., Anderson, P. C., Gonzalez-Urquijo, J. & Gibaja, J. Cereal cultivation and domestication as shown by microtexture analysis of sickle gloss through confocal microscopy. J. Archaeol. Sci. 73, 62–81. https://doi.org/10.1016/j.jas.2016.07.011 (2016).Article 

    Google Scholar 
    Caruso Fermé, L. Modalidades de adquisición y uso del material leñoso entre grupos cazadores-recolectores patagónicos (Argentina). Métodos y técnicas de estudios del material leñoso arqueológico. PhD Dissertation (Universidad Autónoma de Barcelona, Barcelona, 2012).Caruso Fermé, L., Clemente, I., Civalero, M.T. A use-wear analysis of wood technology of patagonian hunter-gatherers. The case of Cerro Casa de Piedra 7, Argentina. J. Archaeol. Sci. 15, 315–321 (2015). https://doi.org/10.1016/j.jas.2015.03.015.Caruso Fermé, L., Aschero, C. Manufacturing and use of the wooden artifacts. A use-wear analysis of wood technology in hunter-gatherer groups (Cerro Casa de Piedra 7 site, Argentina). J. Archaeol. Sci. 31, 102291 (2020). https://doi.org/10.1016/j.quaint.2020.10.067.Schweingruber, F. H. Anatomy of European wood: An atlas for the identification of European trees, shrubs and dwarf shrubs (Paul Haupt, 1990).
    Google Scholar 
    Rageot, M. et al. Birch bark tar production: Experimental and biomolecular approaches to the study of a common and widely used prehistoric adhesive. J. Archaeol. Method Theory 26, 276–312. https://doi.org/10.1007/s10816-018-9372-4 (2019).Article 

    Google Scholar 
    Rageot, M. et al. New insights into Early Celtic consumption practices: Organic residue analyses of local and imported pottery from Vix-Mont Lassois. PLoS ONE 14(6), e0218001. https://doi.org/10.1371/journal.pone.0218001 (2019).CAS 
    Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Arobba, D., Caramiello, R., Martino, G. P. Analisi paleobotaniche di resine dal relitto navale romano del Golfo Dianese. Rivista di Studi Liguri, LXIII-LXIV: 339–355 (1999).Marshall, D. M. Archaeological pollen: extraction from ancient resins. The American Association of Stratigraphic Palynologists. Prog. and Abstr., 38th Ann. Mtg., 34 (2005).Berglund, B. E., Ralska-Jasiewiczowa, M. Pollen analysis and pollen diagrams. in Handbook of Holocene Palaeoecology and Palaeohydrology. (eds. Berglund, B. E.) 455–484 (Chichester, Wiley, 1986).Traverse, A. Paleopalynology. Second Edition, 813 p. (Dordrecht, Springer, 2007).Punt W. (ed.) The Northwest European pollen flora (NEPF), vol. 2 (1980), vol. 3 (1981), vol. 4 (1984) vol. 5 (1988), vol. 6 (1991), vol. 7 (1996), vol. 8 (2003) (Elsevier, Wim Punt, Amsterdam, 1980–2003)Fægri, K. & Iversen, J. Textbook of pollen analysis (John Wiley and Sons, 1989).
    Google Scholar 
    Moore, P. D., Webb, J. A. & Collinson, M. E. Pollen analysis 2nd edn. (Blackwell, 1991).
    Google Scholar 
    Beug, H.-J. Leitfaden der Pollenbestimmung für Mitteleuropa und angrenzende Gebiete (Pfeil, 2004).
    Google Scholar 
    Reille, M. Pollen et spores d’Europe et d’Afrique du Nord. (Marseille, Laboratoire de Botanique Historique et Palynologie, 1992).Katz, O. et al. Rapid phytolith extraction for analysis of phytolith concentrations and assemblages during an excavation: An application at Tell es-Safi/Gath Israel. J. Archaeol. Sci. 37(7), 1557–1563. https://doi.org/10.1016/j.jas.2010.01.016 (2010).Article 

    Google Scholar 
    Brown, D. A. Prospects and limits of a phytolith key for grasses in the central United States. J. Archaeol. Sci. 11, 345–368. https://doi.org/10.1016/0305-4403(84)90016-5 (1984).Article 

    Google Scholar 
    Rosen, A. M. Preliminary identification of silica skeletons from Near Eastern archaeological sites: an anatomical approach. in Phytolith Systematics: Emerging Issues, Advances in Archaeological and Museum Science (eds. Rapp, G. Jr. & Mulholland, S. C.) 129–148 (New York, Plenum Press, 1992)Mulholland, S. C., Rapp Jr. G. A morphological classification of grass silica-bodies. in Phytolith Systematics: Emerging Issues, Advances in Archaeological and Museum Science (eds. Rapp, G. Jr. & Mulholland, S. C.) 65–89 (New York, Plenum Press, 1992)Piperno, D. R. Phytoliths: A comprehensive Guide for Archaeologists and Paleoecologists (Altamira Press, 2006).
    Google Scholar 
    Albert, R. M., & Weiner, S. Study of phytoliths in prehistoric ash layers from Kebara and Tabun caves using a quantitative approach. in Phytoliths: applications in earth sciences and human history, (eds. Meunier, J.D. & Colin, F.) 251–266 (Tokyo, Balkema Publisher, 2001)Albert, R. M. et al. Phytolith-rich layers from the Late Bronze and Iron Ages at Tel Dor (Israel): Mode of formation and archaeological significance. J. Archaeol. Sci. 35(1), 57–75. https://doi.org/10.1016/j.jas.2007.02.015 (2008).Article 

    Google Scholar 
    Albert, R. M., Ruíz, J. A. & Sans, A. PhytCore ODB: A new tool to improve efficiency in the management and exchange of information on phytoliths. J. Archaeol. Sci. 68, 98–105 (2016).Article 

    Google Scholar 
    Portillo, M., Kadowaki, S., Nishiaki, Y. & Albert, R. M. Early Neolithic household behavior at Tell Seker al-Aheimar (Upper Khabur, Syria): A comparison to ethnoarchaeological study of phytoliths and dung spherulites. J. Archaeol. Sci. 42, 107–118 (2014).Article 

    Google Scholar 
    Tsartsidou, G. et al. The phytolith archaeological record: strengths and weaknesses evaluated based on a quantitative modern reference collection from Greece. J. Archaeol. Sci. 34, 1262–1275. https://doi.org/10.1016/j.jas.2006.10.017 (2007).Article 

    Google Scholar 
    Neumann, K., Strömberg , A. E. C., Ball, T., Albert, R. M., Vrydaghs, L. Scott-Cummings, L. (International Committee for Phytolith Taxonomy ICPT). International Code for Phytolith Nomenclature (ICPN) 2.0. Annals of Botany, 124(2): 189–199 (2019).Anderson, P. C. Insight into plant harvesting and other activities at Hatoula, as revealed by microscopic functional analysis of selected chipped stone tools. Le site de Hatoula en Judée occidental. (eds. Lechevallier, M. & Ronen, A.) 277–293 (Paris, Association Paléorient, 1994)Fugazzola Delpino, M.A. La vita quotidiana del Neolitico. Il sito della Marmotta sul Lago di Bracciano. in Settemila anni fa il primo pane. Ambienti e culture delle società neolitiche (eds. Pessina, A. & Muscio G.) 185–192 (Udine, Museo Friulano di Storia Naturale, 1998–1999)Mineo, M. Monossili d’Europa: costruite anche per le rotte marine? in Ubi minor: le isole minori del Mediterraneo centrale: dal Neolitico ai primi contatti coloniali (eds. Guidi, A., Cazzella, A. & Nomi, F.). Scienze dell’Antichità 22, 453–475 (2016)Helwig, K., Monahan, V. & Poulin, J. The identification of hafting adhesive on a slotted antler point from a southwest Yukon ice patch. Am. Antiq. 73, 279–288. https://doi.org/10.1017/S000273160004227X (2008).Article 

    Google Scholar 
    Steigenberger, G. & Herm, C. Natural resins and balsams from an eighteenth-century pharmaceutical collection analysed by gas chromatography/mass spectrometry. Anal. Bioanal. Chem. 401, 1771–1784. https://doi.org/10.1007/s00216-011-5169-y (2011).CAS 
    Article 
    PubMed 

    Google Scholar 
    van den Berg, K. J., Boon, J. J., Pastorova, I. & Spetter, L. F. M. Mass spectrometric methodology for the analysis of highly oxidized diterpenoid acids in Old Master paintings. J. Mass Spectrom. 35, 512–533. https://doi.org/10.1002/(SICI)1096-9888(200004)35:4%3c512::AID-JMS963%3e3.0.CO;2-3 (2000).ADS 
    CAS 
    Article 
    PubMed 

    Google Scholar 
    Behre K. E. Anthropogenic Indicators in Pollen Diagrams, A.A. (Rotterdam, Balkema, 1986).Mercuri, A. M. et al. Anthropogenic Pollen Indicators (API) from archaeological sites as local evidence of human-induced environments in the Italian peninsula. Ann. Bot. 3, 143–153 (2013).
    Google Scholar 
    Andersen, S.-T., Identification of wild grass and cereal pollen. in Danmarks Geologiske Undersøgelse (ed. Aaby, B.) 69–92 (Geological Survey of Denmark, 1978).Bottema, S. Cereal-type pollen in the Near East as indicators of wild or domestic crops. in Préhistoire de l’agriculture: nouvelles approches expérimentales et ethnographiques (ed. Anderson P. C.) 95–106 (Paris, CRA, 1992). https://doi.org/10.1007/BF00217499.Lagerås, P. Long-term history of land-use and vegetation at Femtingagölen—a small lake in the Småland Uplands, southern Sweden. Veg. Hist. Archaeobot. 5, 215–228 (1996).Article 

    Google Scholar 
    Joly, C., Barillé, L., Barreau, M., Mancheron, A. & Visset, L. Grain and annulus diameter as criteria for distinguishing pollen grains of cereals from wild grasses. Rev. Palaeobot. Palynol. 146, 221–233. https://doi.org/10.1016/j.revpalbo.2007.04.003 (2007).Article 

    Google Scholar 
    Punt, W. Umbelliferae. Rev. Palaeobot. Palynol. 42, 155–364 (1984).Article 

    Google Scholar 
    Ellis, M. B. & Ellis, J. P. Microfungi of Land Plants. An Identification Handbook (London, Croom Helm, 1985) (Figure 1270).Ellis, M. B. & Ellis, J. P. Microfungi of Land Plants. An Identification Handbook (London, Croom Helm, 1985) (Figures 174; 176).Rottoli, M., Pessina, A. Neolithic agriculture in Italy: an update of archaeobotanical data with particular emphasis on northern settlements. in The Origins and Spread of Domestic Plants in Southwest Asia and Europe. (eds. Colledge, S. & Conolly, J.) 141–154 (Routledge, New York, 2016)Gurova, M. Prehistoric sickles in the collection of the National Museum of Archaeology in Sofia. in Southeast Europe and Anatolia in Prehistory: Essays in Honor of Vassil Nikolov on his 65th Anniversary (eds. Bacvarov, K. & Gleser, E.) 159–165 (Bonn, Verlag Dr. Rudolf Habelt GmbH, 2016)Sidéra, I. Nouveaux éléments d’origine proche-orientale dans le Néolithique ancien balkanique. in Analyse de l’industrie osseuse. in Préhistoire d’Anatolie. Genèse de deux mondes (ed. Otte, M.), 215–239 (Liège, ERAUL, 1997)Mellaart, J. Excavations at Hacılar: Fourth preliminary report, 1960. Anat. Stud. Anat. Stud. 11, 39–75 (1961).Article 

    Google Scholar 
    Nag, P. K., Goswami, A., Ashtekar, S. P. & Pradhan, C. K. Ergonomics in sickle operation. Appl. Ergon. 19(3), 233–239 (1988).CAS 
    Article 

    Google Scholar 
    Astruc, L., Tkaya, M. B. & Torchy, L. D. l’efficacité des faucilles néolithiques au Proche-Orient: approche expérimentale. Bulletin de la Société préhistorique française 109(4), 671–687 (2012).Article 

    Google Scholar 
    Sigaut, F. Les techniques de récolte des grains : identification, localisation, problèmes d’interprétation. in Rites et rythmes agraires (ed. Cauvin, M.-C.) 31–43 (Lyon, Maison de l’Orient et de la Méditerranée Jean Pouilloux, 1991)Magri, D. Late Quaternary vegetation history at Lagaccione near Lago di Bolsena (central Italy). Rev. Palaeobot. Palynol. 106(3–4), 171–208 (1999).Article 

    Google Scholar 
    Gale, R., & Cutler, D. F. Plants in archaeology: identification manual of vegetative plant materials used in Europe and the Southern Mediterranean to c. 1500 (Westbury and Royal Botanic Gardens, Kew, 2000).Chabal, L. & Feugère, M. L. Le mobilier organique des puits antiques et autres contextes humides de Lattara. Lattara 18, 137–188 (2005).
    Google Scholar 
    Chabal, L. (ed.) Quatre puits de l’agglomération routière gallo-romaine d’Ambrussum (Villetelle, Hérault). Supplément. Revue Archéologique de Narbonnaise, 42: 65–71 (2013).Caruso Fermé, L. & Piqué Huerta, R. Landscape and forest exploitation at the ancient Neolithic site of La Draga (Banyoles, Spain). The Holocene, 24(3): 266 (2014).Boschian, G. Il Riparo “Ermanno de Pompeis” presso l’Eremo di San Bartolomeo di Legio. Scavi 1990–1999. in Atti della XXXVI Riunione Scientifica IIPP, Preistoria e Protostoria dell’Abruzzo, Chieti-Celano, 27–30 settembre 2001, 105–116 (IIPP; Firenze, 2003).Radi, G. & Danese, E. L’abitato di Colle Santo Stefano di Ortucchio (L’Aquila). in Atti della XXXVI Riunione Scientifica IIPP, Preistoria e Protostoria dell’Abruzzo, Chieti-Celano, 27–30 settembre 2001, 145–161 (IIPP; Firenze, 2003).De Francesco, A. M., Bocci, M., Crisci, G. M., & Francaviglia, V. Obsidian provenance at several Italian and Corsican archaeological sites using the non-destructive X-ray fluorescence method. in Obsidian and ancient manufactured glass (eds. Liritzis, I., & Stevenson, C. M.) 115–129 (Albuquerque, UNM Press, 2012).Degano, I. et al. Hafting of Middle Paleolithic tools in Latium (central Italy): New data from Fossellone and Sant’Agostino caves. PLoS ONE 14, e0213473 (2019).CAS 
    Article 

    Google Scholar 
    Nardella, F. et al. Chemical investigations of bitumen from Neolithic archaeological excavations in Italy by GC/MS combined with principal component analysis. Anal. Methods 11, 1449–1459. https://doi.org/10.1039/c8ay02429d (2019).CAS 
    Article 

    Google Scholar 
    Rageot, M. et al. Management systems of adhesive materials throughout the Neolithic in the North-West Mediterranean. J. Archaeol. Sci. 126, 105309 (2021).Article 

    Google Scholar 
    Binder, D., Bourgeois, G., Benoist, F. & Vitry, C. Identification de brai de bouleau (betula) dans le néolithique de Giribaldi (Nice, France) par la spectrométrie de masse. Revue d’Archéométrie 14, 37–42 (1990).Article 

    Google Scholar 
    Vuorela, I. Relative pollen rain around cultivated fields. Acta Bot. Fenn. 102, 1–27 (1973).
    Google Scholar 
    Robinson, M. & Hubbard, R. N. L. B. The transport of pollen in the bracts of hulled cereals. J. Archaeol. Sci. 4(2), 197–199. https://doi.org/10.1016/0305-4403(77)90067-X (1977).Article 

    Google Scholar 
    Hall, V.A., The role of harvesting techniques in the dispersal of pollen grains of Cerealia. Pollen et Spores, XXX, 2, pp. 265–270.Portillo, M., Llergo, Y., Ferrer, A. & Albert, R. M. Tracing microfossil residues of cereal processing in the archaeobotanical record: an experimental approach. Veg. Hist. Archaeobot. 26(1), 59–74. https://doi.org/10.1007/s00334-016-0571-1 (2017).Article 

    Google Scholar 
    Negri, G. Nuovo erbario figurato (Hoepli ed., Milano, 1981).Paris R. R. & Moyse H. Matière Médicale. Vol 2°, (Masson, Paris. 1976).Bulgarelli, G. & Flamigni, S. Le piante tossiche e velenose (Hoepli ed., Milano, 2010).Les, D. H. Aquatic Dicotyledons of North America: Ecology, Life History, and Systematics (CRC Press, 2017).Book 

    Google Scholar 
    Curti, L. Herbarium, un’inedita collezione di piante del XVIII secolo conservata presso l’orto Botanico dell’Università di Padova (Offset Invicta S.p.A., Padova, 1992).Rottoli, M. Zafferanone selvatico (Carthamus lanatus) e cardo della Madonna (Silybum marianum), piante raccolte o coltivate nel Neolitico antico a “La Marmotta”? Bollettino di Paletnologia Italiana, 91–92, 47–61 (2000–2001).Rottoli, M. “La Marmotta”, Anguillara Sabazia (RM), scavi 1989. Analisi paletnobotaniche: prime risultanze. Bullettino di Paletnologia Italiana 84, 305–315 (1993).Van Geel, B. Non-pollen palynomorphs. in Tracking Environmental Change Using Lake Sediments: Terrestrial, vol. 3. (ed. Smol, J. P., Birks, H. J. B., Last W. M.) 99–119 (Algal and Siliceous Indicators, New York, 2001)Hawksworth, David L., van Geel, Bas, Wiltshire, Patricia E. J. The enigma of the Diporotheca palynomorph. Rev. Palaeobot. Palynol. 235, 94–98 (2016). https://doi.org/10.1016/j.revpalbo.2016.09.010.Krug, J. C., Benny, G. L., Keller, H. W. Coprophilous fungi. In Biodiversity of Fungi. Inventory and Monitoring Methods (ed. Foster M., Bill, G.) 467–499 (Elsevier Science, Amsterdam, 2004). More