More stories

  • in

    Effects of precipitation and temperature on precipitation use efficiency of alpine grassland in Northern Tibet, China

    1.
    Xu, X. et al. Interannual variability in responses of belowground net primary productivity (NPP) and NPP partitioning to long-term warming and clipping in a tallgrass prairie. Global Change Biol. 18(5), 1648–1656 (2012).
    ADS  Article  Google Scholar 
    2.
    Hui, D. & Jackson, R. B. Geographical and interannual variability in biomass partitioning in grassland ecosystems: a synthesis of field data. New Phytol. 169(1), 85–93 (2006).
    CAS  Article  Google Scholar 

    3.
    Zhang, X. et al. Impact of prolonged drought on rainfall use efficiency using MODIS data across China in the early 21st century. Remote Sens. Environ. 150, 188–197 (2014).
    ADS  Article  Google Scholar 

    4.
    Jiang, Y. et al. Effects of community structure on precipitation-use efficiency of grasslands in Northern Tibet. J. Veg Sci. 28, 281–290 (2017).
    Article  Google Scholar 

    5.
    Roupsard, O. et al. Scaling-up productivity (NPP) using light or water use efficiencies (LUE, WUE) from a two-layer tropical plantation. Agrofor. Syst. 76(2), 409–422 (2009).
    Article  Google Scholar 

    6.
    Prince, S. D., De Colstoun, E. B. & Kravitz, L. L. Evidence from rain-use efficiencies does not indicate extensive Sahelian desertification. Global Change Biol. 4(4), 359–374 (1998).
    ADS  Article  Google Scholar 

    7.
    Fensholt, R. & Rasmussen, K. Analysis of trends in the Sahelian “rain-use efficiency” using GIMMS NDVI, RFE and GPCP rainfall data. Remote Sens. Environ. 115(2), 438–451 (2011).
    ADS  Article  Google Scholar 

    8.
    Ye, H., Wang, J., Huang, M. & Qi, S. Spatial pattern of vegetation precipitation use efficiency and its response to precipitation and temperature on the Qinghai-Xizang Plateau of China. Chin. J. Plant. Ecol. 36(12), 1237–1247 (2012).
    Article  Google Scholar 

    9.
    Bai, Y. et al. Primary production and rain use efficiency across a precipitation gradient on the Mongolia plateau. Ecology 89(8), 2140–2153 (2008).
    Article  Google Scholar 

    10.
    Paruelo, J. M., Lauenroth, W. K., Burke, I. C. & Sala, O. E. Grassland precipitation-use efficiency varies across a resource gradient. Ecosystems 2(1), 64–68 (1999).
    Article  Google Scholar 

    11.
    Yang, Y., Fang, J., Fay, P. A., Bell, J. E. & Ji, C. Rain use efficiency across a precipitation gradient on the Tibetan Plateau. Geophys. Res. Lett. 37, L15702 (2010).
    ADS  Google Scholar 

    12.
    Li, H. X., Liu, G. H. & Fu, B. J. Spatial variations of rain-use efficiency along a climate gradient on the Tibetan Plateau: a satellite-based analysis. Int. J. Remote Sens. 34(21), 7487–7503 (2013).
    ADS  Article  Google Scholar 

    13.
    Huxman, T. E. et al. Convergence across biomes to a common rain-use efficiency. Nature 429(6992), 651–654 (2004).
    ADS  CAS  Article  Google Scholar 

    14.
    Hu, Z. et al. Precipitation-use efficiency along a 4500-km grassland transect. Glob. Ecol. Biogeogr. 19(6), 842–851 (2010).
    Article  Google Scholar 

    15.
    Lauenroth, W. K., Burke, I. C. & Paruelo, J. M. Patterns of production and precipitation-use efficiency of winter wheat and native grasslands in the central Great Plains of the United States. Ecosystems 3(4), 344–351 (2000).
    Article  Google Scholar 

    16.
    Hooper, D. U. & Johnson, L. Nitrogen limitation in dryland ecosystems: Responses to geographical and temporal variation in precipitation. Biogeochemistry 46(1–3), 247–293 (1999).
    CAS  Google Scholar 

    17.
    Xu, X., Sherry, R. A., Niu, S., Li, D. & Luo, Y. Net primary productivity and rain-use efficiency as affected by warming, altered precipitation, and clipping in a mixed-grass prairie. Global Change Biol. 19(9), 2753–2764 (2013).
    ADS  Article  Google Scholar 

    18.
    De Boeck, H. J. et al. How do climate warming and plant species richness affect water use in experimental grasslands?. Plant Soil 288(1–2), 249–261 (2006).
    ADS  Article  Google Scholar 

    19.
    Campos, G. E. P. et al. Ecosystem resilience despite large-scale altered hydroclimatic conditions. Nature 494(7437), 349–352 (2013).
    ADS  Article  Google Scholar 

    20.
    Qiu, J., Zhang, H. & Shen, W. Spatial characteristics of precipitation use efficiency on the Qinghai-Tibet Plateau From 1982 to 2007. J. Fudan. Univ. Nat. Sci. 53(1), 126–133 (2014).
    Google Scholar 

    21.
    Wang, Q. W., Yu, D. P., Dai, L. M., Zhou, L. & Zhou, W. M. Research progress in water use efficiency of plants under global climate change. Chin. J. Appl. Ecol. 21(12), 3255–3265 (2000).
    Google Scholar 

    22.
    Chen, S. P., Bai, Y. F., Zhang, L. X. & Han, X. G. Comparing physiological responses of two dominant grass species to nitrogen addition in Xilin River Basin of China. Environ. Exp. Bot. 53(1), 65–75 (2005).
    Article  Google Scholar 

    23.
    Qiu, J. The third pole. Nature 454(7203), 393–396 (2008).
    CAS  Article  Google Scholar 

    24.
    Chen, B. X. et al. The impact of climate change and anthropogenic activities on alpine grassland over the Qinghai-Tibet Plateau. Agric. For. Meteorol. 189, 11–18 (2014).
    ADS  Article  Google Scholar 

    25.
    Jiang, Y. B. et al. Effects of community structure on precipitation-use efficiency of grasslands in northern Tibet. J. Veg. Sci. 28, 281–290 (2017).
    Article  Google Scholar 

    26.
    Gao, Q. Z. et al. Effects of topography and human activity on the net primary productivity (NPP) of alpine grassland in northern Tibet from 1981 to 2004. Int. J. Remote Sens. 34(6), 2057–2069 (2013).
    ADS  Article  Google Scholar 

    27.
    Zhang, J. H., Yao, F. M., Zheng, L. G. & Yang, L. M. Evaluation of grassland dynamics in the Northern-Tibet Plateau of China using remote sensing and climate data. Sensors 7(12), 3312–3328 (2007).
    Article  Google Scholar 

    28.
    Li, Z., Huffman, T., McConkey, B. & Townley-Smith, L. Monitoring and modeling spatial and temporal patterns of grassland dynamics using time-series MODIS NDVI with climate and stocking data. Remote Sens. Environ. 138, 232–244 (2013).
    ADS  Article  Google Scholar 

    29.
    Zhang, X. K., Lu, X. Y. & Wang, X. D. Spatial-temporal NDVI variation of different alpine grassland classes and groups in Northern Tibet from 2000 to 2013. Mt. Res. Dev. 35(3), 254–263 (2015).
    Article  Google Scholar 

    30.
    Yu, D. Y., Shi, P. J., Shao, H. B., Zhu, W. Q. & Pan, Y. H. Modelling net primary productivity of terrestrial ecosystems in East Asia based on an improved CASA ecosystem model. Int. J. Remote Sens. 30(18), 4851–4866 (2009).
    ADS  Article  Google Scholar 

    31.
    Gao, Q. Z. et al. Dynamics of alpine grassland NPP and its response to climate change in Northern Tibet. Clim. Change 97(3–4), 515–528 (2009).
    ADS  CAS  Article  Google Scholar 

    32.
    Zhu, W. Q., Pan, Y. Z., He, H., Yu, D. Y. & Hu, H. B. Simulation of maximum light use efficiency for some typical vegetation types in China. Chin. Sci. Bull. 51(4), 457–463 (2006).
    CAS  Article  Google Scholar 

    33.
    Zhao, G. S. et al. Spatial-temporal variation of ANPP and rain-use efficiency along a precipitation gradient on Changtang Plateau, Tibet. Remote Sens. 11, 325 (2019).
    ADS  Article  Google Scholar 

    34.
    Sun, J. & Du, W. Effects of precipitation and temperature on net primary productivity and precipitation use efficiency across China’s grasslands. GISci. Remote Sens. 54(6), 881–897 (2017).
    Article  Google Scholar 

    35.
    Chen, Z. Q., Shao, Q. Q., Liu, J. Y. & Wang, J. B. Analysis of net primary productivity of terrestrial vegetation on the Qinghai-Tibet Plateau, based on MODIS remote sensing data. Sci. China Earth Sci. 55(8), 1306–1312 (2012).
    ADS  Article  Google Scholar 

    36.
    Piao, S. & Fang, J. Terrestrial net primary production and its spatio-temporal patterns in Qinghai-Xizang Plateau, China during 1982–1999. J. Nat. Resour. 03, 373–380 (2002).
    Google Scholar 

    37.
    Gao, Q. Z., Wan, Y. F., Li, Y. E., Lin, E. D. & Yang, K. Grassland net primary production and its spatiotemporal distribution in Northern Tibet: a study with CASA model. Chin. J. Appl. Ecol. 11, 2526–2532 (2007).
    Google Scholar 

    38.
    Zhou, C. P., Ouyang, H., Wang, Q. X., Watanabe, M. & Sun, Q. Q. Estimation net primary productivity in Tibetan Plateau. Acta Geogr. Sin. 01, 74–79 (2004).
    Google Scholar 

    39.
    Yang, Y. H. et al. Storage, patterns and controls of soil organic carbon in the Tibetan grasslands. Global Change Biol. 14, 1592–1599 (2008).
    ADS  Article  Google Scholar  More

  • in

    Arbuscular mycorrhizal fungi favor invasive Echinops sphaerocephalus when grown in competition with native Inula conyzae

    1.
    Spatafora, J. W. et al. A phylum-level phylogenetic classification of zygomycete fungi based on genome-scale data. Mycologia 108, 1028–1046 (2016).
    CAS  PubMed  PubMed Central  Article  Google Scholar 
    2.
    Smith, S. E. & Read, D. J. Mycorrhizal Symbiosis (Academic Press, Amsterdam, 2008).
    Google Scholar 

    3.
    van der Heijden, M. G. A., Martin, F. M., Selosse, M. A. & Sanders, I. R. Mycorrhizal ecology and evolution: the past, the present, and the future. New Phytol. 205, 1406–1423 (2015).
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    4.
    Lekberg, Y., Hammer, E. C. & Olsson, P. A. Plants as resource islands and storage units—adopting the mycocentric view of arbuscular mycorrhizal networks. FEMS Microbiol. Ecol. 74, 336–345 (2010).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    5.
    Allen, M. F. Mycorrhizal fungi: highways for water and nutrients in arid soils. Vadose Zone J. 6, 291–297 (2007).
    Article  Google Scholar 

    6.
    Newsham, K. K., Fitter, A. H. & Watkinson, A. R. Arbuscular mycorrhiza protect an annual grass from root pathogenic fungi in the field. J. Ecol. 83, 991–1000 (1995).
    Article  Google Scholar 

    7.
    Vigo, C., Norman, J. R. & Hooker, J. E. Biocontrol of the pathogen Phytophthora parasitica by arbuscular mycorrhizal fungi is a consequence of effects on infection loci. Plant Pathol. 49, 509–514 (2000).
    Article  Google Scholar 

    8.
    Aroca, R., Porcel, R. & Ruiz-Lozano, J. M. How does arbuscular mycorrhizal symbiosis regulate root hydraulic properties and plasma membrane aquaporins in Phaseolus vulgaris under drought, cold or salinity stresses?. New Phytol. 173(4), 808–816 (2007).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    9.
    Augé, R. M., Toler, H. D. & Saxton, A. M. Arbuscular mycorrhizal symbiosis and osmotic adjustment in response to NaCl stress: a meta-analysis. Front Plant Sci. 5, ARTN 562. https://doi.org/10.3389/fpls.2014.00562 (2014).

    10.
    Augé, R. M., Toler, H. D. & Saxton, A. M. Arbuscular mycorrhizal symbiosis alters stomatal conductance of host plants more under drought than under amply watered conditions: a meta-analysis. Mycorrhiza 25(1), 13–24 (2015).
    PubMed  Article  PubMed Central  Google Scholar 

    11.
    Pfeffer, P. E., Douds, D. D., Becard, G. & Shachar-Hill, Y. Carbon uptake and the metabolism and transport of lipids in an arbuscular mycorrhiza. Plant Physiol. 120(2), 587–598 (1999).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    12.
    Bago, B., Pfeffer, P. E. & Shachar-Hill, Y. Carbon metabolism and transport in arbuscular mycorrhizas. Plant Physiol. 124(3), 949–958 (2000).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    13.
    Horton, T. R. Mycorrhizal networks (Springer, Dordrecht, 2015).
    Google Scholar 

    14.
    Walder, F. & van der Heijden, M. G. A. Regulation of resource exchange in the arbuscular mycorrhizal symbiosis. Nat. Plants 1(11), 7 (2015).
    Article  CAS  Google Scholar 

    15.
    van der Heijden, M. G. A. et al. Mycorrhizal fungal diversity determines plant biodiversity, ecosystem variability and productivity. Nature 396(6706), 69–72 (1998).
    ADS  Article  CAS  Google Scholar 

    16.
    Wilson, G. W. T., Hartnett, D. C. & Rice, C. W. Mycorrhizal-mediated phosphorus transfer between the tallgrass prairie plants Sorghastrum nutans and Artemisia ludoviciana. Funct. Ecol. 20, 427–435 (2006).
    Article  Google Scholar 

    17.
    Bever, J. D. et al. Rooting theories of plant community ecology in microbial interactions. Trends Ecol. Evol. 25(8), 468–478 (2010).
    PubMed  PubMed Central  Article  Google Scholar 

    18.
    Walder, F. et al. Mycorrhizal networks: common goods of plants shared under unequal terms of trade. Plant Physiol. 159, 789–797 (2012).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    19.
    Weremijewicz, J., Sternberg, L. & Janos, D. P. Common mycorrhizal networks amplify competition by preferential mineral nutrient allocation to large host plants. New Phytol. 212(2), 461–471 (2016).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    20.
    Řezáčová, V. et al. Little cross-feeding of the mycorrhizal networks shared between C3-Panicum bisulcatum and C4-Panicum maximum under different temperature regimes. Front. Plant Sci. 9, 16. https://doi.org/10.3389/fpls.2018.00449 (2018).
    Article  Google Scholar 

    21.
    Deslippe, J. R. & Simard, S. W. Below-ground carbon transfer among Betula nana may increase with warming in Arctic tundra. New Phytol. 192, 689–698 (2011).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    22.
    Bever, J. D., Richardson, S. C., Lawrence, B. M., Holmes, J. & Watson, M. Preferential allocation to beneficial symbiont with spatial structure maintains mycorrhizal mutualism. Ecol. Lett. 12(1), 13–21 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    23.
    Lendenmann, M. et al. Symbiont identity matters: carbon and phosphorus fluxes between Medicago truncatula and different arbuscular mycorrhizal fungi. Mycorrhiza 21(8), 689–702 (2011).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    24.
    Kiers, E. T. et al. Reciprocal rewards stabilize cooperation in the mycorrhizal symbiosis. Science 333(6044), 880–882 (2011).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    25.
    Rillig, M. C. Arbuscular mycorrhizae and terrestrial ecosystem processes. Ecol. Lett. 7, 740–754 (2004).
    Article  Google Scholar 

    26.
    Verbruggen, E. & Kiers, E. T. Evolutionary ecology of mycorrhizal functional diversity in agricultural systems. Evol Appl. 3(5–6), 547–560 (2010).
    PubMed  PubMed Central  Article  Google Scholar 

    27.
    van Kleunen, M. et al. Global exchange and accumulation of non-native plants. Nature 525(7567), 100–103 (2015).
    ADS  PubMed  Article  CAS  PubMed Central  Google Scholar 

    28.
    Pejchar, L. & Mooney, H. A. Invasive species, ecosystem services and human well-being. Trends Ecol. Evol. 24(9), 497–504 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    29.
    Pyšek, P. et al. A global assessment of invasive plant impacts on resident species, communities and ecosystems: the interaction of impact measures, invading species’ traits and environment. Glob. Change Biol. 18(5), 1725–1737 (2012).
    ADS  Article  Google Scholar 

    30.
    Blackburn, T. M. et al. A unified classification of alien species based on the magnitude of their environmental impacts. PLoS Biol. 12(5), ARTN e1001850. https://doi.org/10.1371/journal.pbio.1001850 (2014).

    31.
    Mitchell, C. E. et al. Biotic interactions and plant invasions. Ecol. Lett. 9(6), 726–740 (2006).
    PubMed  Article  PubMed Central  Google Scholar 

    32.
    Catford, J. A., Jansson, R. & Nilsson, C. Reducing redundancy in invasion ecology by integrating hypotheses into a single theoretical framework. Divers. Distrib. 15(1), 22–40 (2009).
    Article  Google Scholar 

    33.
    van der Putten, W. H. Impacts of soil microbial communities on exotic plant invasions. Trends Ecol. Evol. 25(9), 512–519 (2010).
    PubMed  Article  Google Scholar 

    34.
    Keane, R. M. & Crawley, M. J. Exotic plant invasions and the enemy release hypothesis. Trends Ecol. Evol. 17(4), 164–170 (2002).
    Article  Google Scholar 

    35.
    Pyšek, P. et al. Naturalization of central European plants in North America: species traits, habitats, propagule pressure, residence time. Ecology 96(3), 762–774 (2015).
    PubMed  Article  Google Scholar 

    36.
    Davis, M. A., Grime, J. P. & Thompson, K. Fluctuating resources in plant communities: a generaltheory of invasibility. J. Ecol. 88, 528–534 (2000).
    Article  Google Scholar 

    37.
    Callaway, R. M., Thelen, G. C., Rodriguez, A. & Holben, W. E. Soil biota and exotic plant invasion. Nature 427(6976), 731–733 (2004).
    ADS  CAS  PubMed  Article  Google Scholar 

    38.
    Rudgers, J. A. & Orr, S. Non-native grass alters growth of native tree species via leaf and soil microbes. J. Ecol 97(2), 247–255 (2009).
    Article  Google Scholar 

    39.
    Sun, Z. K. & He, W. M. Evidence for enhanced mutualism hypothesis: Solidago canadensis plants from regular soils perform better. PLoS ONE 5(11), 5. https://doi.org/10.1371/journal.pone.0015418 (2010).
    CAS  Article  Google Scholar 

    40.
    Dickie, I. A. et al. The emerging science of linked plant-fungal invasions. New Phytol. 215(4), 1314–1332 (2017).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    41.
    Cronk, Q. C. B. & Fuller, J. R. Plant Invaders: The Threat to Natural Ecosystems (Earthscan Publications, London, 2001).
    Google Scholar 

    42.
    Richardson, D. M., Allsopp, N., D’Antonio, C. M., Milton, S. J. & Rejmanek, M. Plant invasions—the role of mutualisms. Biol. Rev. 75(1), 65–93 (2000).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    43.
    Pringle, A. et al. Mycorrhizal symbioses and plant invasions. Ann Rev. Ecol. Evol. Syst. 40, 699–715 (2009).
    Article  Google Scholar 

    44.
    Wilson, G. W. T., Hickman, K. R. & Williamson, M. M. Invasive warm-season grasses reduce mycorrhizal root colonization and biomass production of native prairie grasses. Mycorrhiza 22, 327–336 (2012).
    PubMed  Article  PubMed Central  Google Scholar 

    45.
    Nunez, M. A. & Dickie, I. A. Invasive belowground mutualists of woody plants. Biol. Invasions 16, 645–661 (2014).
    Article  Google Scholar 

    46.
    Bunn, R. A., Ramsey, P. W. & Lekberg, Y. Do native and invasive plants differ in their interactions with arbuscular mycorrhizal fungi? A meta-analysis. J. Ecol. 103, 1547–1556 (2015).
    CAS  Article  Google Scholar 

    47.
    Gucwa-Przepiora, E., Chmura, D. & Sokolowska, K. AM and DSE colonization of invasive plants in urban habitat: a study of Upper Silesia (southern Poland). J. Plant Res. 129, 603–614 (2016).
    PubMed  PubMed Central  Article  Google Scholar 

    48.
    Waller, L. P., Callaway, R. M., Klironomos, J. N., Ortega, Y. K. & Maron, J. L. Reduced mycorrhizal responsiveness leads to increased competitive tolerance in an invasive exotic plant. J. Ecol. 104, 1599–1607 (2016).
    Article  Google Scholar 

    49.
    Menzel, A. et al. Mycorrhizal status helps explain invasion success of alien plant species. Ecology 98, 92–102 (2017).
    PubMed  Article  PubMed Central  Google Scholar 

    50.
    Broadbent, A. A. D., Stevens, C. J., Ostle, N. J. & Orwin, K. H. Biogeographic differences in soil biota promote invasive grass response to nutrient addition relative to co-occurring species despite lack of belowground enemy release. Oecologia 186, 611–620 (2018).
    ADS  PubMed  Article  PubMed Central  Google Scholar 

    51.
    Vogelsang, K. M. & Bever, J. D. Mycorrhizal densities decline in association with nonnative plants and contribute to plant invasion. Ecology 90, 399–407 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    52.
    Reinhart, K. O. & Callaway, R. M. Soil biota and invasive plants. New Phytol. 170, 445–457 (2006).
    PubMed  Article  PubMed Central  Google Scholar 

    53.
    Pakpour, S. & Klironomos, J. The invasive plant, Brassica nigra, degrades local mycorrhizas across a wide geographical landscape. R. Soc. Open Sci. 2, 4 (2015).
    Article  Google Scholar 

    54.
    Shah, M. A., Reshi, Z. A. & Khasa, D. P. Arbuscular mycorrhizas: Drivers or passengers of alien plant invasion. Bot. Rev. 75, 397–417 (2009).
    Article  Google Scholar 

    55.
    De Souza, T. A. F., Rodriguez-Echeverria, S., de Andrade, L. A. & Freitas, H. Could biological invasion by Cryptostegia madagascariensis alter the composition of the arbuscular mycorrhizal fungal community in semi-arid Brazil?. Acta Bot. Bras. 30, 93–101 (2016).
    Article  Google Scholar 

    56.
    Awaydul, A. et al. Common mycorrhizal networks influence the distribution of mineral nutrients between an invasive plant, Solidago canadensis, and a native plant, Kummerowa striata. Mycorrhiza 29, 29–38 (2019).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    57.
    Štajerová, K., Šmilauerová, M. & Šmilauer, P. Arbuscular mycorrhizal symbiosis of herbaceous invasive neophytes in the Czech Republic. Preslia 81, 341–355 (2009).
    Google Scholar 

    58.
    Hempel, S. et al. Mycorrhizas in the Central European flora: relationships with plant life history traits and ecology. Ecology 94, 1389–1399 (2013).
    PubMed  Article  PubMed Central  Google Scholar 

    59.
    Callaway, R. M., Newingham, B., Zabinski, C. A. & Mahall, B. E. Compensatory growth and competitive ability of an invasive weed are enhanced by soil fungi and native neighbours. Ecol. Lett. 4, 429–433 (2001).
    Article  Google Scholar 

    60.
    Workman, R. E. & Cruzan, M. B. Common mycelial networks impact competition in an invasive grass. Am. J. Bot. 103, 1041–1049 (2016).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    61.
    Zhang, Q. et al. Potential allelopathic effects of an invasive species Solidago canadensis on the mycorrhizae of native plant species. Allelopathy J. 20, 71–77 (2007).
    ADS  CAS  Google Scholar 

    62.
    Callaway, R. M. et al. Novel weapons: Invasive plant suppresses fungal mutualists in America but not in its native Europe. Ecology 89, 1043–1055 (2008).
    PubMed  Article  PubMed Central  Google Scholar 

    63.
    Sarma, K. K. V. Allelopathic potential of Echinops echinatus and Solanum surratense on seed germination of Argemone mexicana. Trop. Ecol. 15, 156–157 (1974).
    Google Scholar 

    64.
    Smith, M. D., Hartnett, D. C. & Wilson, G. W. T. Interacting influence of mycorrhizal symbiosis and competition on plant diversity in tallgrass prairie. Oecologia 121, 574–582 (1999).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    65.
    Bennett, J. A. et al. Plant-soil feedbacks and mycorrhizal type influence temperate forest population dynamics. Science 355, 181–184 (2017).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    66.
    Liao, H. X. et al. Soil microbes regulate forest succession in a subtropical ecosystem in China: evidence from a mesocosm experiment. Plant Soil 430, 277–289 (2018).
    CAS  Article  Google Scholar 

    67.
    Řezáčová, V. et al. Mycorrhizal symbiosis induces plant carbon reallocation differently in C3 and C4Panicum grasses. Plant Soil 425, 441–456 (2018).
    Article  CAS  Google Scholar 

    68.
    Newman, E. I. A method of estimating total length of root in a sample. J. Appl. Ecol. 3, 139–145 (1966).
    Article  Google Scholar 

    69.
    Bukovská, P., Gryndler, M., Gryndlerová, H., Püschel, D. & Jansa, J. Organic nitrogen-driven stimulation of arbuscular mycorrhizal fungal hyphae correlates with abundance of ammonia oxidizers. Front. Microbiol. 7, 711 (2016).
    PubMed  PubMed Central  Article  Google Scholar 

    70.
    Hewitt, E. J. Sand and water culture methods used in the study of plant nutrition. CAB Tech. Commun. 22, 431–432 (1966).
    Google Scholar 

    71.
    Řezáčová, V. et al. Imbalanced carbon-for-phosphorus exchange between European arbuscular mycorrhizal fungi and non-native Panicum grasses—a case of dysfunctional symbiosis. Pedobiologia 62, 48–55 (2017).
    Article  Google Scholar 

    72.
    Ohno, T. & Zibilske, L. M. Determination of low concentrations of phosphorus in soil extracts using malachite green. Soil Sci. Soc. Am. J. 55, 892–895 (1991).
    ADS  CAS  Article  Google Scholar 

    73.
    McGonigle, T. P., Miller, M. H., Evans, D. G., Fairchild, G. L. & Swan, J. A. A new method which gives an objective-measure of colonization of roots by vesicular arbuscular mycorrhizal fungi. New Phytol. 115, 495–501 (1990).
    Article  Google Scholar 

    74.
    Koske, R. E. & Gemma, J. N. A modified procedure for staining roots to detect VA-mycorrhizas. Mycol. Res. 92, 486–505 (1989).
    Article  Google Scholar 

    75.
    Gryndler, M. et al. Tuber aestivum Vittad. mycelium quantified: advantages and limitations of a qPCR approach. Mycorrhiza 23, 341–348 (2013).
    PubMed  Article  PubMed Central  Google Scholar 

    76.
    Thonar, C., Erb, A. & Jansa, J. Real-time PCR to quantify composition of arbuscular mycorrhizal fungal communities-marker design, verification, calibration and field validation. Mol. Ecol. Res. 12, 219–232 (2012).
    CAS  Article  Google Scholar 

    77.
    von Felten, A., Défago, G. & Maurhofer, M. Quantification of Pseudomonas fluorescens strains F113, CHA0 and Pf153 in the rhizosphere of maize by strain-specific real-time PCR unaffected by the variability of DNA extraction efficiency. J. Microbiol. Methods 81, 108–115 (2010).
    Article  CAS  Google Scholar 

    78.
    Janoušková, M., Püschel, D., Hujslová, M., Slavíková, R. & Jansa, J. Quantification of arbuscular mycorrhizal fungal DNA in roots: how important is material preservation?. Mycorrhiza 25, 205–214 (2015).
    PubMed  Article  CAS  Google Scholar  More

  • in

    Effectiveness of the European Natura 2000 network to sustain a specialist wintering waterbird population in the face of climate change

    International Waterbird Census (IWC) data suggest 309,000 Scaup were wintering in North-West Europe in 1988–1991, compared with 192,300 in 2015–2018, indicating that the number of Scaup in this flyway has declined by 38.1% over 31 years (equivalent to a 30.3% decline over three generations, given a Scaup generation length of 8.2 years). Such a rate of decrease qualifies this population as vulnerable (VU) according to criterion A2(c) of the International Union for Conservation of Nature24. Thus, our results confirm the recent attribution of Scaup as a VU on the European Red List18. We suggest that the 1% threshold for the North-West Europe population of the Scaup should be revised to 1900.
    In addition to the overall decline in abundance, we also show that changes in winter temperature on the eastern and northern edges of the wintering range potentially explain the observed dramatic shift in winter distribution closer to the breeding grounds. Climate change appears to have opened up more wintering sites to Scaup, especially in the more northern and eastern areas where reductions in winter ice cover have made previous staging sites increasingly accessible in winter. This might be expected to have a positive effect on the population, given that Scaup have more potential wintering sites to choose between and that they face a diminished risk from mass starvation because of the reduced probability of unexpected ice cover of potential feeding areas25. However, the ultimate causes of shifts in wintering distribution remain unknown and could equally relate to deterioration of food quality in southern and western wintering grounds. At Lake IJsselmeer, the annual changes in the large numbers of wintering Scaup there in the 1980s and 1990s were explained by fluctuations in the abundance of their main prey, Zebra Mussel Dreissena polymorpha26. The decline in Zebra Mussels in the IJsselmeer lake and its replacement by Quagga Mussels Dreissena rostriformis bugensis27 resulted in a deterioration in the quality of food resources at the site. These are likely contributory reasons to explain the shift in the centre of gravity of the Scaup wintering grounds to Poland and eastern Germany, although we lack data to determine the magnitude of this effect (Fig. 4, Unit#3). This area now constitutes the most important wintering area for this population, although the detection of Quagga Mussels in this region in 201428 represents a potential threat to the quality of this important wintering ground.
    Assuming that some of the birds remain to winter along the migration route on sites formerly only used as stopovers, we can retrospectively infer the migration route of the Scaup population breeding in northern Russia and Fennoscandia (Fig. 1). It would appear that after birds reach the Baltic, they stop in Estonia before traversing the Baltic south-west to Gotland, migrating along the southern coast of Sweden and onwards to the main wintering area in Danish, German and Polish Baltic waters (Unit#3). Some Scaup continue west to reach Unit#2 in the Netherlands, and small numbers continue to reach France and the UK. The small population breeding in Iceland likely winter exclusively in the UK and Ireland, where fewer of the Russian/Fennoscandia population reach in recent winters. The Iceland breeding birds likely constitute a separate biogeographic population, with little contact with the main one discussed here (Fig. 1). Assuming the continuing effects of global warming, we can predict further separation of the two sub-populations and that Unit#4 (Fig. 4), the coast of Gotland and the islands and bays in Estonia will most likely play an increasingly important future role as winter quarters for this species. This is likely to be the case at other sites within eastern Baltic where this species can find suitable habitats.
    Our historical analysis has shown that after a period of most rapid decline during 1988–2003, this population could be interpreted as remaining stable during 2003–2018 (Fig. 2). We suspect that this may be partly the result of the significant decrease in the Scaup bycatch in the Netherlands29,30,31. The added mortality from fisheries bycatch represents one of the most important threats to the relatively long-lived Scaup32. Evidence showed that drowning mortality was extremely high between 1985 and 1994, when an estimated average of 17,672 birds died annually in fishing gear (6% of the total population of the time), but this has declined since the 2000s32. Of all Scaup from this flyway population that drowned in fishing nets in years 1978–1990, up to 65% died at the most important wintering site at the time—the Dutch IJsselmeer32. However, our highly uneven knowledge of the extent of the Scaup bycatch throughout its winter range should be taken into account here. Exceptionally detailed estimates from IJsselmeer during the earlier period14 contrasts our lack of data or poor estimates from elsewhere, which may result in a bias that implies a greater importance for Scaup bycatch at the IJsselmeer for the population than was actually the case. Current estimates of bycatch levels throughout the flyway suggest that Scaup death in fishing nets has decreased, amounting to c.4000 individuals yearly, partly explained by the substantial decrease in the Dutch bycatch32.
    The second highly important threat to Scaup, perhaps as important as the bycatch, is the deterioration of their food resources. Detailed energy budget studies on Lake IJsselmeer14 suggested that foraging Scaup there were operating on the margins of energetic profitability and the limited number of important wintering sites elsewhere suggest that alternative sites are really scarce, implying that food availability at core wintering sites could potentially affect winter survival.
    The specialist habitat selection of the Scaup restricts it to a narrow range of habitats during the wintering period where it aggregates in large concentrations, a factor which causes the entire wintering population to concentrate in relatively few locations. Potentially, this makes them more vulnerable at the population level than most other, more dispersed diving duck species. During the January 2015 count, 91% of counted birds were present at 31 locations in five countries (Denmark, Germany, the Netherlands, Poland and Sweden). The four most important locations supported over two-thirds of the total wintering numbers: namely IJsselmeer in the Netherlands, Barther Bodden and Greifswalder Bodden in Germany and Odra river estuary in Poland (Fig. 4). Taken together, these areas have consistently been the most important wintering areas for Scaup over the last 30 years3,14,20, with two thirds of the flyway population during winter concentrated within 5300 km2 (2000 km2 in the Netherlands and 3300 km2 in Poland/Germany).
    Wintering areas in Germany and Poland also act as stopover sites, so much larger numbers are counted there in autumn and spring migration, with up to 100,000 individuals on the Szczecin Lagoon (c.470 km29). Similarly, in Estonia, where a few hundred birds winter (Fig. 1), numbers may exceed 100,000 individuals in spring33. Therefore, cohesive planning for the effective conservation of the species, requires adequate protection at both the most important wintering sites (analysed in this article) and stopover sites along the entire migration route. During spring migration, extremely large Scaup concentrations can occur in these important sites, which provide for other biological functions such a communal courtship, displaying, pair-bonding etc.32. Given that Scaup are among the most vulnerable of diving ducks to bycatch34 (constituting more than 50% of diving birds drowned in fishing nets in the Polish Odra Estuary35) potentially high mortality during the prelude to the breeding season is likely to have severe adverse effects on the entire population. It is important to remember that this site can simultaneously support up to 75% of the total population9 and intensive fishing takes place here with gillnets35—the method of fishing recognised as the most dangerous for drowning diving birds in the Baltic Sea6.
    Other environmental pressures on Scaup are no less serious, but currently less well quantified. Many important wintering areas are situated in estuaries of large rivers that invariably host major sea ports, where large vessels cause disturbance and pollution. Maintenance of shipping channels requires dredging (as in the case of the channel leading to the port of Amsterdam on IJsselmeer in the Netherlands and that serving the port of Szczecin on the Szczecin Lagoon in Poland). Dredging of shallow marine and brackish substrates can disrupt sediment horizons, mobilising suspended material, creating turbidity and disrupting the food resource and the ability of Scaup to forage for their prey. The proximity to human settlements also makes these shallow marine waters attractive to the increasing practice of water sports, kite- and wind-surfing, boating and recreational fishing from boats, which although not a source of direct mortality, contributes to disturbance and displacement of Scaup from favoured areas36,37.
    SPAs and effectiveness of protection
    The long-term conservation aim for a decreasing qualifying species, in accordance with European Union (EU) law (Birds Directive—Council Directive 2009/147/EC), should be to recover them to former level of abundance. To achieve this aim, SPAs should be designated in sites where 1% or more of the biogeographic population regularly occurs. In the case of Scaup, all of such areas are protected in the form of SPA (Table 2). Subsequently, such a SPAs should have a Management Plan (MP) defining the conservation objectives within each site, updated every 6 years. Of the three most important Scaup SPAs in Europe, only the IJsselmeer (NL9803028, Unit#2, Core wintering area, Fig. 4) has a MP for 2013–201738, which described the long term decline (since 1994) in wintering numbers of Scaup in the IJsselmeer and identified the greatest threats for Scaup as declining food resources and disturbance by developing water sports. Although bycatch was conspicuously not listed as a threat, the MP documents previous measures, taken to reduce fishing effort, had resulted from the implementation of another EU Directive—the Water Framework Directive (WFD, Directive 2000/60/EC). The WFD committed EU Member States to achieve good qualitative and quantitative status of all water bodies by 201538. Conservation measures carried out on Lake IJsselmeer over the last 75 years aimed to maintain sustainable fishing did not bring about the intended results on fish stock39. However, they may have had a positive effect on reducing bycatch of Scaup from 11,500 killed annually during 1978–199032 to insignificant numbers in the years 2011–201231, which may have contributed to the slowing in the rate of population decline at this time. In the most important wintering area for this flyway populations—the lagoons and bays either side of the German-Polish border, out of ten SPAs forming one coherent area (Fig. 4) only two have MPs. Moreover, the key SPAs within this area that regularly hold the highest Scaup numbers do not have MPs, they are: Greifswalder Bodden und südlicher Strelasund (DE1747402) in Germany and Szczecin Lagoon (PLB320009) in Poland. The Greifswalder Bodden, Teile des Strelasundes und Nordspitze Usedom (DE1747301) Special Area of Conservation (SAC), which overlaps with the Greifswalder Bodden und südlicher Strelasund SPA, was created under the Habitats Directive (Council Directive 92/43/EEC) and has a MP that identifies the threats to Scaup (e.g. from bycatch). However, because MPs for SACs (as against SPAs) are not primarily directed towards bird conservation, there are no specific regulations to limit the current stressors upon Scaup at this site40. The existing MPs for two other SPAs (“Vorpommersche Boddenlandschaft und nördlicher Strelasund” and “Dolina Dolnej Odry”) either do not identify main threats to Scaup or fail to impose sufficient conservation measures41,42.
    Other SPAs that are less important for Scaup within Unit#3 west of the core wintering area include Östliche Kieler Bucht (DE1530491) and Ostsee östlich Wagrien (DE1633491), which have MPs identifying the threat from bycatch. This includes a voluntary agreement between the Schleswig Holstein Ministry of the Environment and local fishery associations, under which areas are closed to fishing if “concentrations of ducks” ( > 100) are present in the areas. Fishermen have two days to remove their gear after closure. There are rigid legal provisions at these two sites that prohibit fishing with gillnets within 200 m of the shore43,44. To date, there is no evidence of a positive effect and reduction of bycatch of diving birds, so we recommend a study of the effectiveness of these provisions.
    The shift in the centre of gravity of the wintering population to Germany and Poland highlights the ineffectiveness of conservation measures directed towards Scaup (and other diving birds) there. Despite the existence of SPAs in which the Scaup is specifically protected and evidence of the cost of gillnet bycatch to local diving ducks, the most serious pressure remains unchecked. In 2011–2012, results from research work in the Szczecin Lagoon37 recommended the MP proposed reducing the Scaup (and other diving birds) bycatch by spatiotemporal regulation of gillnet fisheries to avoid key areas used by the birds. Unfortunately, the effective solutions to deliver results for bird conservation were considered too far-reaching by fishing interests. The fishing lobby blocked official approval of the MP by government and so these measures were never implemented. Given the high rates of Scaup bycatch, the designation of the area as a SPA offers no effective protection to the species at this site32. The effectiveness of SPA designation for a particular species remains ineffective, as long as effective management is not implemented. Given the increasing relative importance of the German/Polish resorts to the species in recent years, the lack of effective measures within these SPAs is becoming more critical to safeguard the conservation of the North-West Europe population of Scaup. Suitably prepared MPs, containing a bycatch monitoring order, would solve this problem, setting bycatch thresholds, according to the recommendations of BirdLife International45—1% of natural mortality calculated on the basis of local species abundance. If this value is exceeded, spatiotemporal restrictions on gillnet fishery should be introduced.
    Looking to the future, areas that were formerly stopovers are already becoming wintering sites in Sweden and Estonia. Although currently not numerically significant in winter, these sites already hold significant numbers during migration. In the future, satellite areas (Unit#4) have the potential to develop into important wintering grounds and therefore require adequate protection from factors known to affect Scaup survival.
    Previous studies show that bycatch in fishing nets is one of the most serious anthropogenic pressures during the non-breeding period for many diving birds6, although we cannot exclude the influence of other factors such as food availability and quality26 and disturbance from hunting46 and water sports37. Because the majority of North-West Europe’s Scaup winter in relatively few places, conservation interventions at these key sites are particularly important. The shift in wintering distribution poses new challenges for countries increasingly responsible for the conservation of this species in winter. Lack of adequate protection in this region means that these areas may act as sink habitats (in the sense of the source-sink model 47). The shift of wintering areas to sink habitats exposes an increasing part of the population to the pressures present there. This is not only the case for Scaup but also for a range of other diving bird species. These birds concentrate in the most attractive areas rich in food, often biologically productive transitional waters, where marine and freshwater birds meet in high densities. For this reason, effective conservation measures directed at Scaup will positively impact upon a whole range of other species with similar ecology. This suggests that protection measures taken for the Scaup could also benefit associated marine species in the same areas such as Long-tailed Duck, Velvet Scoter, Common Scoter Melanitta nigra, as well as for coastal zone species such as: Tufted Duck, Smew Mergellus albellus, and Goosander Mergus merganser. More

  • in

    Male and female genotype and a genotype-by-genotype interaction mediate the effects of mating on cellular but not humoral immunity in female decorated crickets

    Adamo SA (2004) Estimating disease resistance in insects: phenoloxidase and lysozyme-like activity and disease resistance in the cricket Gryllus texensis. J Insect Physiol 50:209–216
    CAS  PubMed  Article  Google Scholar 

    Alcock J (1994) Postinsemination associations between males and females in insects: the mate-guarding hypothesis. Annu Rev Entomol 39:1–21
    Article  Google Scholar 

    Archer CR, Zajitschek F, Sakaluk SK, Royle NJ, Hunt J (2012) Sexual selection affects the evolution of lifespan and ageing in the decorated cricket Gryllodes sigillatus. Evolution 66:3088–3100
    CAS  PubMed  Article  Google Scholar 

    Arnqvist G (1988) Mate guarding and sperm displacement in the water strider Gerris lateralis Schumm. (Heteroptera: Gerridae). Freshw Biol 19:269–274
    Article  Google Scholar 

    Arnqvist G, Nilsson T (2000) The evolution of polyandry: multiple mating and female fitness in insects. Anim Behav 60:145–164
    CAS  PubMed  Article  Google Scholar 

    Arnqvist G, Rowe L (2002) Antagonistic coevolution between the sexes in a group of insects. Nature 415:787–789
    CAS  PubMed  Article  Google Scholar 

    Avila FW, Sirot LK, Laflamme BA, Rubinstein CD, Wolfner MF (2010) Insect seminal fluid proteins: Identification and function. Annu Rev Entomol 56:21–40
    Article  CAS  Google Scholar 

    Azad P, Ryu J, Haddad GG (2011) Distinct role of Hsp70 in Drosophila hemocytes during severe hypoxia. Free Radic Biol Med 51:530–538
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Baer B, Morgan ED, Schmid-Hempel P (2001) A nonspecific fatty acid within the bumblebee mating plug prevents females from remating. Proc Natl Acad Sci U S A 98:3926–3928
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Barribeau SM, Schmid-Hempel P (2017) Sexual healing: mating induces a protective immune response in bumblebees. J Evol Biol 30:202–209
    CAS  PubMed  Article  Google Scholar 

    Bascuñán-García AP, Lara C, Córdoba-Aguilar A (2010) Immune investment impairs growth, female reproduction and survival in the house cricket, Acheta domesticus. J Insect Physiol 56:204–211
    PubMed  Article  CAS  Google Scholar 

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

    Burnham KP, Anderson DR (2002) Model selection and multimodel inference: a practical information-theoretic approach, 2nd edn. Springer-Verlag, New York
    Google Scholar 

    Burpee DM, Sakaluk SK (1993) Repeated matings offset costs of reproduction in female crickets. Evol Ecol 7:240–250
    Article  Google Scholar 

    Castella G, Christe P, Chapuisat (2009) Mating triggers dynamic immune regulations in wood ant queen. J Evol Biol 22:564–570
    CAS  PubMed  Article  Google Scholar 

    Chapman T, Arnqvist G, Bangham J, Rowe L (2003) Sexual conflict. Trends Ecol Evol 18:41–47
    Article  Google Scholar 

    Cordero A (1990) The adaptive significance of the prolonged copulations of the damselfly, Ischnura graellsii (Odonata: Coenagrionidae). Anim Behav 40:43–48
    Article  Google Scholar 

    Cordero A (1999) Forced copulations and female contact guarding at a high male density in a calopterygid damselfly. J Insect Behav 12:27–37
    Article  Google Scholar 

    Delbare SYN, Chow CY, Wolfner MF, Clark AG (2017) Roles of female and male genotype in post-mating responses in Drosophila melanogaster. J Hered 4:740–753
    Article  CAS  Google Scholar 

    Dickinson JL, Rutowski RL (1989) The function of the mating plug in the chalcedon checkerspot butterfly. Anim Behav 38:154–162
    Article  Google Scholar 

    Dougherty LR, van Lieshout E, McNamara KB, Moschilla JA, Arnqvist G, Simmons LW (2017) Sexual conflict and correlated evolution between male persistence and female resistance traits in the seed beetle Callosobruchus maculatus. Proc R Soc B Biol Sci 284:20170132
    Article  Google Scholar 

    Duffield KR, Hampton KJ, Houslay TM, Hunt J, Rapkin J, Sakaluk SK, Sadd BM (2018) Age-dependent variation in the terminal investment threshold in male crickets. Evolution 72:578–589
    PubMed  Article  Google Scholar 

    Duffield KR, Hampton KJ, Houslay TM, Hunt J, Sadd BM, Sakaluk SK (2019) Inbreeding alters context‐dependent reproductive effort and immunity in male crickets. J Evol Biol 32:731–741
    PubMed  Google Scholar 

    Edward DA, Poissant J, Wilson AJ, Chapman T (2014) Sexual conflict and interacting phenotypes: a quantitative genetic analysis of fecundity and copula duration in Drosophila melanogaster. Evolution 68:1651–1660
    PubMed  Article  Google Scholar 

    Eggert AK, Reinhardt K, Sakaluk SK (2003) Linear models for assessing mechanisms of sperm competition: the trouble with transformations. Evolution 57:173–176
    PubMed  Article  Google Scholar 

    Fedorka KM, Zuk M (2005) Sexual conflict and female immune suppression in the cricket, Allonemobious socius. J Evol Biol 18:1515–1522
    PubMed  Article  Google Scholar 

    Fedorka KM, Zuk M, Mousseau TA (2004) Immune suppression and the cost of reproduction in the ground cricket, Allonemobius socius. Evolution 58:2478–2485
    PubMed  Article  Google Scholar 

    Fedorka KM, Linder JE, Winterhalter W, Promislow D (2007) Post-mating disparity between potential and realized immune response in Drosophila melanogaster. Proc R Soc B Biol Sci 274:1211–1217
    Article  Google Scholar 

    Fricke C, Ávila‐Calero S, Armitage SA (2020) Genotypes and their interaction effects on reproduction and mating‐induced immune activation in Drosophila melanogaster. J Evol Biol 33:930–941
    CAS  PubMed  Article  Google Scholar 

    Gershman SN, Barnett CA, Pettinger AM, Weddle CB, Hunt J, Sakaluk SK (2010a) Inbred decorated crickets exhibit higher measures of macroparasitic immunity than outbred individuals. Heredity 105:282–289
    CAS  PubMed  Article  Google Scholar 

    Gershman SN, Barnett CA, Pettinger AM, Weddle CB, Hunt J, Sakaluk SK (2010b) Give ‘til it hurts: trade-offs between immunity and male reproductive effort in the decorated cricket, Gryllodes sigillatus. J Evol Biol 23:829–839
    CAS  PubMed  Article  Google Scholar 

    Gershman SN, Hunt J, Sakaluk SK (2013) Food fight: sexual conflict over free amino acids in the nuptial gifts of male decorated crickets. J Evol Biol 26:693–704
    CAS  PubMed  Article  Google Scholar 

    Gershman SN, Mitchell C, Sakaluk SK, Hunt J (2012) Biting off more than you can chew: sexual selection on the free amino acid composition of the spermatophylax in decorated crickets. Proc R Soc B Biol Sci 279:2531–2538
    CAS  Article  Google Scholar 

    Gillespie JP, Kanost MR, Trenczek T (1997) Biological mediators of insect immunity. Annu Rev Entomol 42:611–654
    CAS  PubMed  Article  Google Scholar 

    Gillott C (2003) Male accessory gland secretions: modulators of female reproductive physiology and behavior. Annu Rev Entomol 48:163–184
    CAS  PubMed  Article  Google Scholar 

    Goenaga J, Yamane T, Rönn J, Arnqvist G (2015) Within-species divergence in the seminal fluid proteome and its effect on male and female reproduction in a beetle. BMC Evol Biol 15:266
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    González-Santoyo I, Córdoba-Aguilar A (2012) Phenoloxidase: a key component of the insect immune system. Entomol Exp Appl 142:1–16
    Article  CAS  Google Scholar 

    Haerty W, Jagadeeshan S, Kulathinal RJ, Wong A, Ravi Ram K, Sirot LK et al (2007) Evolution in the fast lane: Rapidly evolving sex-related genes in Drosophila. Genetics 177:1321–1335
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Hall MD, Lailvaux SP, Brooks RC (2013) Sex‐specific evolutionary potential of pre‐and postcopulatory reproductive interactions in the field cricket, Teleogryllus commodus. Evolution 67:1831–1837
    PubMed  Article  PubMed Central  Google Scholar 

    Hardin JW, Hilbe, JM (2007) Generalized linear models and extensions. 2nd edn, Stata Press, College Station, Texas

    Ivy TM, Sakaluk SK (2005) Polyandry promotes enhanced offspring survival in decorated crickets. Evolution 59:152–159
    PubMed  Article  Google Scholar 

    Ivy TM, Weddle CB, Sakaluk SK (2005) Females use self-referent cues to avoid mating with previous mates. Proc R Soc B Biol Sci 272:2475–2478
    Article  Google Scholar 

    Kacsoh BZ, Schlenke TA (2012) High hemocyte load is associated with increased resistance against parasitoids in Drosophila suzukii, a relative of D. melanogaster. PLoS ONE 7:e34721
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Kerr AM, Gershman SN, Sakaluk SK (2010) Experimentally induced spermatophore production and immune responses reveal a trade-off in crickets. Behav Ecol 21:647–654
    Article  Google Scholar 

    Klowden MJ (1999) The check is in the male: Male mosquitoes affect female physiology and behavior. J Am Mosq Contr 15:213–220
    CAS  Google Scholar 

    Kwon H, Smith RC (2019) Chemical depletion of phagocytic immune cells in Anopheles gambiae reveals dual roles of mosquito hemocytes in anti-Plasmodium immunity. Proc Natl Acad Sci U S A 116:14119–14128
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Lawniczak MKN, Barnes AI, Linklater JR, Boone JM, Wigby S, Chapman T (2007) Mating and immunity in invertebrates. Trends Ecol Evol 22:48–55
    PubMed  Article  Google Scholar 

    Lavine MD, Strand MR (2002) Insect hemocytes and their role in immunity. Insect Biochem Mol Biol 32:1295–1309
    CAS  PubMed  Article  Google Scholar 

    Lenth, SingmannH, Love J, Buerkner P, Herve M (2020) emmeans: estimated marginal means, aka least-squares means. Release 1.4.5. https://CRAN.R-project.org/package=emmeans

    Lung O, Wolfner MF (2001) Identification and characterization of the major Drosophila melanogaster mating plug protein. Insect Biochem Mol Biol 31:543–551
    CAS  PubMed  Article  Google Scholar 

    Miller JS, Nguyen T, Stanley-Samuelson DW (1994) Eicosanoids mediate insect nodulation responses to bacterial infections. Proc Natl Acad Sci U S A 91:12418–12422
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Morrow EH, Innocenti P (2012) Female postmating immune responses, immune system evolution and immunogenic males. Biol Rev 87:631–638
    PubMed  Article  Google Scholar 

    Nappi AJ, Vass E (1993) Melanogenesis and the generation of cytotoxic molecules during insect cellular immune reactions. Pigment Cell Res 6:117–126
    CAS  PubMed  Article  Google Scholar 

    Oku K, Price TAR, Wedell N (2019) Does mating negatively affect female immune defences in insects? Anim Biol 69:117–136
    Article  Google Scholar 

    Otti O (2015) Genitalia‐associated microbes in insects. Insect Sci 22:325–339
    PubMed  Article  Google Scholar 

    Parker GA, Birkhead TR (2013) Polyandry: the history of a revolution. Philos Trans R Soc Lond B Biol Sci 368:1–13
    Article  Google Scholar 

    Pauchet Y, Wielsch N, Wilkinson PA, Sakaluk SK, Svatoš A, ffrench-Constant RH et al (2015) What’s in the gift? Towards a molecular dissection of nuptial feeding in a cricket. PLoS ONE 10:e0140191
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    Peng J, Zipperlen P, Kubli E (2005) Drosophila sex-peptide stimulates female innate immune system after mating via the Toll and Imd pathways. Curr Biol 15:1690–1694
    CAS  PubMed  Article  Google Scholar 

    Perry JC, Siro L, Wigby S (2013) The seminal symphony: how to compose an ejaculate. Trends Ecol Evol 28:414–422
    PubMed  PubMed Central  Article  Google Scholar 

    Ramirez JL, Garver LS, Brayner FA, Alves LC, Rodrigues J, Molina-Cruz A et al (2014) The role of hemocytes in Anopheles gambiae antiplasmodial immunity. J Innate Immun 6:119–128
    CAS  PubMed  Article  Google Scholar 

    Ravi Ram K, Wolfner MF (2007) Seminal influences: Drosophila acps and the molecular interplay between males and females during reproduction. Integr Comp Biol 47:427–445
    CAS  PubMed  Article  Google Scholar 

    R Core Team (2019) R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria

    Rice WR, Holland B (1997) The enemies within: intergenomic conflict, interlocus contest evolution (ICE), and the intraspecific red queen. Behav Ecol Sociobiol 41:1–10
    Article  Google Scholar 

    Rolff J, Siva-Jothy MT (2002) Copulation corrupts immunity: a mechanism for a cost of mating in insects. Proc Natl Acad Sci U S A 99:9916–9918
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Rowe L, Arnqvist G (2002) Sexually antagonistic coevolution in a mating system: cominbing experimental and comparative approaches to address evolutionary processes. Evolution 56:754–767
    PubMed  Article  Google Scholar 

    Sakaluk SK (1984) Male crickets feed females to ensure complete sperm transfer. Science 223:609–610
    CAS  PubMed  Article  Google Scholar 

    Sakaluk SK (1986) Sperm competition and the evolution of nuptial feeding behavior in the cricket, Gryllodes supplicans (Walker). Evolution 40:584–593
    PubMed  Article  Google Scholar 

    Sakaluk SK (1987) Reproductive behaviour of the decorated cricket, Gryllodes supplicans (Orthoptera: Gryllidae): calling schedules, spatial distribution, and mating. Behaviour 100:202–225
    Article  Google Scholar 

    Sakaluk SK (1991) Post-copulatory mate guarding in decorated crickets. Anim Behav 41:207–216
    Article  Google Scholar 

    Sakaluk SK (2000) Sensory exploitation as an evolutionary origin to nuptial food gifts in insects. Proc R Soc Lond B Biol Sci 267:339–343

    Sakaluk SK, Avery RL, Weddle CB (2006) Cryptic sexual conflict in gift-giving insects: chasing the chase-away. Am Nat 167:94–104
    PubMed  Article  Google Scholar 

    Sakaluk SK, Duffield KR, Rapkin J, Sadd BM, Hunt J (2019) The troublesome gift: the spermatophylax as a purveyor of sexual conflict and coercion in crickets. Adv Stud Behav 51:1–30
    Article  Google Scholar 

    Sakaluk SK, Eggert AK (1996) Female control of sperm transfer and intraspecific variation in sperm precedence: antecedents to the evolution of a courtship food gift. Evolution 50:694–703
    PubMed  Article  Google Scholar 

    Sakaluk SK, Schaus JM, Eggert AK, Snedden WA, Brady PL (2002) Polyandry and fitness of offspring reared under varying nutritional stress in decorated crickets. Evolution 56:1999–2007
    PubMed  Article  Google Scholar 

    Schneider CA, Rasband WS, Eliceiri KW (2012) NIH Image to ImageJ: 25 years of image analysis. Nat Methods 9:671–675
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Schneider PM (1985) Purification and properties of three lysozymes from hemolymph of the cricket, Gryllus bimaculatus (De Geer). Insect Biochem 15:463–470
    CAS  Article  Google Scholar 

    Schwenke RA, Lazzaro BP, Wolfner MF (2016) Reproduction-immunity trade-offs in insects. Annu Rev Entomol 61:239–256
    CAS  PubMed  Article  Google Scholar 

    Sherman KJ (1983) The adaptive significance of postcopulatory mate guarding in a dragonfly, Pachydiplax longipennis. Anim Behav 31:1107–1115
    Article  Google Scholar 

    Shoemaker KL, Parsons NM, Adamo SA (2006) Mating enhances parasite resistance in the cricket Gryllus texensis. Anim Behav 71:371–380
    Article  Google Scholar 

    Short SM, Lazzaro BP (2010) Female and male genetic contributions to post-mating immune defence in female Drosophila melanogaster. Proc R Soc B Biol Sci 277:3649–3657
    Article  Google Scholar 

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

    Stanley D, Kim Y (2014) Eicosanoid signaling in insects: from discovery to plant protection. Crit Rev Plant Sci 33:20–63
    CAS  Article  Google Scholar 

    Sugumaran M, Nellaiappan K, Valivittan K (2000) A new mechanism for the control of phenoloxidase activity: inhibition and complex formation with quinone isomerase. Arch Biochem Biophys 379:252–260
    CAS  PubMed  Article  Google Scholar 

    Theopold U, Schmidt O, Söderhäll K, Dushay MS (2004) Coagulation in arthropods: defence, wound closure and healing. Trends Immunol 25:289–294
    CAS  PubMed  Article  Google Scholar 

    Therneau T (2020) A package for survival analysis in R. Release 3.1-11. https://CRAN.R-project.org/package=survival

    Warwick S, Vahed K, Raubenheimer D, Simpson SJ (2009) Free amino acids as phagostimulants in cricket nuptial gifts: support for the “candymaker” hypothesis. Biol Lett 5:194–196
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Wigby S, Chapman T (2005) Sex peptide causes mating costs in female Drosophila melanogaster. Curr Biol 15:316–321
    CAS  PubMed  Article  Google Scholar 

    Wolfner MF (1997) Tokens of love: functions and regulation of Drosophila male accessory gland products. Insect Biochem Mol Biol 27:179–192
    CAS  PubMed  Article  Google Scholar 

    Worthington AM, Jurenka RA, Kelly CD (2015) Mating for male-derived prostaglandin: a functional explanation for the increased fecundity of mated female crickets? J Exp Biol 218:2720–2727
    PubMed  Article  Google Scholar 

    Worthington AM, Kelly CD (2016) Females gain survival benefits from immune-boosting ejaculates. Evolution 70:928–933
    PubMed  Article  Google Scholar 

    Yi HY, Chowdhury M, Huang YD, Yu XQ (2014) Insect antimicrobial peptides and their applications. Appl Microbiol Biotechnol 98:5807–5822
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Zhong W, Priest NK, McClure CD, Evans CR, Mlynski DT, Immonen E et al (2013) Immune anticipation of mating in Drosophila: turandot m promotes immunity against sexually transmitted fungal infections. Proc R Soc B Biol Sci 280:1–9
    Google Scholar  More

  • in

    Fairy circles in Namibia are assembled from genetically distinct grasses

    Fairy circles (FCs) are remarkably circular and regular vegetation patterns covering an area of thousands of square kilometres in hyper-arid grasslands in Southern Africa (e.g. Namibia)1,2,3. They are typically made up of perennial tufts of Stipagrostis grass growing around barren circular patches roughly 2–10 m in diameter, with annual or perennial grasses in the area between circles, the matrix1,2. The edges of FCs are about 5–10 m apart and when dense they have a regular hexagonal arrangement3. Similar vegetation rings also occur in arid lands in Australia, Israel and elsewhere1,2,3,4,5,6,7. The processes leading to the formation of FCs have been the subject of debate for several decades8,9,10,11,12,13.
    Aspects of FCs that require explanation include how and why individual circles form, how they persist and how the similar size and regular pattern is established at the landscape scale. The bare centres of individual FCs persist in situ almost permanently, although the peripheral plants delimiting the circles are much shorter-lived. Van Rooyen et al.8 noted no change in the position of five marked FCs over 22 years, despite intervening droughts8. Tschinkel14 analysed pairs of aerial photographs taken 4 years apart, and on the basis of how many FC positions were unchanged, died or emerged, estimated that average sized FCs persist in situ for an average of 75 years, implying some circles remain in situ for a century14. Similarly, using aerial photographs, Juergens2 noted a high survival (97%) of FCs in situ over a 50-year period (=0.06% pa mortality), implying an age of millennia for some FCs2.
    Many alternative hypotheses have been proposed for FC spatial and temporal patterns, but without agreement. The first hypothesis is based on ecosystem engineering by termites that remove plants from the centres of circles2, facilitating localized underground water accumulation in circle centres. This moisture maintains the termites and the band of perennial grass on which termites feed year-round. The spatial patterning of the FCs is considered to result from competition between termite colonies2. However, the poor correlation of FCs with the presence of specific termites is an important concern with this hypothesis13.
    The second hypothesis is based on the clonal mode of growth of individuals of many arid-land species that create vegetation rings4,6,7,15. For example, rings are formed by one of the Namibian FC species, Stipagrostis ciliata in the Negev desert. Here individual plants of this species send out underground rhizomes, which, with increasing age, results in a ring of ramets (i.e. sprouts from the same clonal colony or genet) around a barren central patch, which forms as the plant centre dies4. Such vegetation rings form and enlarge centrifugally due to competition between ramets, and as this process continues over time new ramets establish successfully only towards the periphery4,7. Globally, all of the many plant species that form circles do so by this type of clonal growth15. As the pattern of FCs is virtually fixed in situ for centuries, this suggests that the plants that indicate the pattern may also be long-lived. Individual clonal plants can be extremely long-lived15,16, which could then match longevities between the plants and the FCs they delimit. The clonality hypothesis could thus explain how circular shapes form (by self-thinning of ramets spreading from source plant), why bare centres occur (resource depletion and source plant death) and how they can persist over long periods of time (by continuous production of short-lived ramets). However, clonality has been disputed8 as an explanation for FCs for two reasons. Firstly, van Rooyen et al. suggested that the FCs referred to by Danin and Orshan are considerably smaller (about 2 m) than most FCs in Namibia4,8. Although this is correct that FCs tend to be much larger in Namibia, the mean size of FCs can be as small as 2.5 m in some areas2. Secondly, van Rooyen and colleagues suggested that clonality cannot explain why FC centres are bare8. Bare centres are now well known in rings and are commonly explained as being due to inter-ramet competition and resource depletion7.
    The third hypothesis for FC spatial patterns is that it emerges through vegetation self-organization (the VSO hypothesis)1,5,12,17,18. Grasses in the peripheral band outcompete grasses in the FC centre and keep it bare and moist at depth1. The plants in the peripheral bands also compete with the matrix grasses and with the peripheral bands on adjacent FCs to maintain the regular pattern1. Mathematical vegetation models based on partial differential equations represent the theoretical basis of the VSO hypothesis17,18. These partial differential equations do not operate at the scale of individual-plant interactions, but at the level of local processes (largely rates of lateral water movement due to plant evapo-transpiration) in comparison to rates of lateral biomass spread. For these current VSO models, lateral water flow needs to be about 100 times faster than lateral biomass spread18. Thus, the mode and rate of how biomass spreads, whether via clonal growth or seed dispersal and population growth18, has an impact on the viability of vegetation-patterning models. In the absence of such information for FCs, the most recent VSO modelling study18 assumed a rate of biomass dispersal (1.2 m2 yr−1) derived from the Canadian woodlands19, for clonal spread or seed dispersal. These literature values may be unrealistic for FCs. For example, 1.2 m2 yr−1 is likely to be too high for clonal growth in the arid circumstances in which FCs occur. Stipagrostis individual plants across the landscape are typically only 0.005–0.13 m2 in canopy area with a mean area of 0.05 m2 (ref. 20). Similarly, clonal spread in other systems can also be lower than 1.2 m2 yr−1 by orders of magnitude (ref. 19). Alternatively, if Stipagrostis plants are not clones, their highly awned seeds will disperse much further than 1.2 m2 yr−1. Finally, if the assumed 1.2 m2 yr−1 biomass spread18 is due to the total growth of new recruits per parent individual, then this implies unrealistically high population growth rates ( >20 new recruits, each with 0.05 m2 in canopy volume20 required per parent individual). Getzin et al.3 acknowledge that in the context of their VSO model it needs to be further investigated whether grass tufts experience a central dieback due to self‐thinning, i.e. the role of clonality33. Thus, clonality may be relevant to some VSO models.
    Juergens has argued that there is a spatial mismatch in the root length and inter-FC distances10 and therefore that FCs cannot directly interact with each other to produce the regular spatial pattern. An extreme example of this mismatch is a study by Ravi et al. reporting that the roots of peripheral plants in FCs had a mean length of 5.9 cm, which is more than two orders of magnitude shorter than inter-circle distances (10–20 m)12. These short root lengths would make competition for water and other resources over long distances crucial for explaining FC patterning1. Most recently, Ravi et al.12 rejected the termite hypothesis due to an absence of termites and partially invoked clonal dynamics as well as the VSO to explain their FC patterning12.
    In summary, there are critical issues with all the hypotheses explaining the formation of FCs and the role of clonality appears to be relevant to two of the hypotheses. Here, we use ddRAD-seq as a genetic test of clonality of peripheral grasses. Our analysis indicates that most individual grasses surrounding FCs are genetically distinct and does not support the clonality hypothesis. More

  • in

    Characteristics of temperature evolution from 1960 to 2015 in the Three Rivers’ Headstream Region, Qinghai, China

    The spatiotemporal characteristics analysis
    The temporal characteristics
    As seen in Fig. 2a, the climate tendency rate of the annual mean temperature series is 0.337 °C per decade, and its correlation coefficient is 0.7674 ( > r0.01 = 0.3357). In conjunction with the result of the M–K trend test, this finding demonstrates that the annual mean temperature has significantly increased in the past 56 years. In addition, the climate tendency rates of the annual mean temperature in the LARHR, YERHR and YARHR are 0.352 °C per decade, 0.34 °C per decade and 0.319 °C per decade, respectively (except for the climate tendency rate of the annual temperature in the YARHR, the correlation coefficients of the annual temperature series exceeded the significance level of 0.01). These rates exceed the mean rising rate of annual mean temperature in China (0.21–0.25 °C per decade) and that of the global annual mean temperature (0.07℃ per decade) in the past 56 years. The climate tendency rates of the annual mean temperature in the LARHR and YERHR were higher than those in Northwest China (0.32 °C/10a)1,27, and the climate tendency rate of the annual mean temperature in the YARHR was similar to that in Northwest China. The higher annual climate tendency rate of the THRHR is related to the increase in the lowest night temperature in the study area28 and the large amount of solar radiation in the lower altitude area of the THRHR13; furthermore, it may also be related to the decrease in total cloud cover and the increase in snow cover in the study area29. In addition, as seen in Fig. 2a,b, the annual mean temperature of the THRHR has increased significantly since the late1990s.
    Figure 2

    The change in annual mean temperature in the THRHR from 1960 to 2015.

    Full size image

    As seen in Fig. 3, the annual and seasonal mean temperatures in the THRHR and its three subregions have been increasing in fluctuations since the 1960s, and the fluctuations in winter and spring are greater than those of summer and autumn. Due to the higher climate tendency rate of the summer, autumn and winter in LARHR than in the other two subregions, the orders of the climate tendency rate of seasonal mean temperature in the THRHR are completely consistent with that of LARHR (autumn, summer, winter, and spring mean temperatures), which are different from that in Northwest China (winter, autumn, spring and summer mean temperatures). In the three subregions, the spring climate tendency rate is lowest, and the orders of the climate tendency rate of autumn and winter are higher than that of spring, while that of summer varies greatly, it is highest in the YARHR, and that of autumn in the LARHR and the YERHR are highest. The high temperature rise rate in autumn and winter in the THRHR is related to the increase in the lowest temperature at night, the climate high tendency rate of winter in three subregions corresponds with the positive phase of the Arctic Oscillation (AO), and namely, the high climate tendency rate of winter corresponds with the high value of the AO. In addition, the annual and seasonal mean temperatures in the LARHR are ≥ 0 °C, respectively; meanwhile, the increasing range of the seasonal mean temperature climate tendency rate is significantly higher than that in the YERHR and YARHR, respectively. The decadal mean temperatures, the decadal minimum and maximum temperatures and their extreme values in the YERHR are lower than those in the LARHR, while they are higher than those in the YARHR (the value of decadal mean temperature is  More

  • in

    Bowhead whales use two foraging strategies in response to fine-scale differences in zooplankton vertical distribution

    1.
    Laidre, K. L., Heide-Jørgensen, M. P., Nielsen, T. G. & Gissel Nielsen, T. Role of the bowhead whale as a predator in West Greenland. Mar. Ecol. Prog. Ser. 346, 285–297 (2007).
    ADS  Article  Google Scholar 
    2.
    Pomerleau, C., Ferguson, S. H. & Walkusz, W. Stomach contents of bowhead whales (Balaena mysticetus) from four locations in the Canadian Arctic. Polar Biol. 34, 615–620 (2011).
    Article  Google Scholar 

    3.
    Pomerleau, C. et al. Prey assemblage isotopic variability as a tool for assessing diet and the spatial distribution of bowhead whale Balaena mysticetus foraging in the Canadian eastern Arctic. Mar. Ecol. Prog. Ser. 469, 161–174 (2012).
    ADS  Article  Google Scholar 

    4.
    Kenney, R. D., Hyman, M. A. M., Owen, R. E., Scott, G. P. & Winn, H. E. Estimation of prey densities required by western North Atlantic right whales. Mar. Mamm. Sci. 2, 1–13 (1986).
    Article  Google Scholar 

    5.
    Baumgartner, M. F. & Tarrant, A. M. The physiology and ecology of diapause in marine copepods. Ann. Rev. Mar. Sci. 9, 387–411 (2017).
    PubMed  Article  Google Scholar 

    6.
    Fortune, S. M., Trites, A. W., Mayo, C. A., Rosen, D. A. S. & Hamilton, P. K. Energetic requirements of North Atlantic right whales and the implications for species recovery. Mar. Ecol. Prog. Ser. 478, 253–272 (2013).
    ADS  Article  Google Scholar 

    7.
    Hays, G. C., Richardson, A. J. & Robinson, C. Climate change and marine plankton. Trends Ecol. Evol. 20, 337–344 (2005).
    PubMed  Article  PubMed Central  Google Scholar 

    8.
    Beaugrand, G., Mackas, D. & Goberville, E. Applying the concept of the ecological niche and a macroecological approach to understand how climate influences zooplankton: advantages, assumptions, limitations and requirements. Prog. Oceanogr. 111, 75–90 (2013).
    ADS  Article  Google Scholar 

    9.
    Beaugrand, G., Reid, P. C., Ibañez, F., Lindley, J. A. & Edwards, M. Reorganization of North Atlantic marine copepod biodiversity and climate. Science 296, 1692–1694 (2002).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    10.
    Beaugrand, G. Decadal changes in climate and ecosystems in the North Atlantic Ocean and adjacent seas. Deep Res. Part II Top. Stud. Oceanogr. 56, 656–673 (2009).
    ADS  Article  Google Scholar 

    11.
    Chust, G. et al. Are Calanus spp. shifting poleward in the North Atlantic? A habitat modelling approach. ICES J. Mar. Sci. 71, 241–253 (2014).
    Article  Google Scholar 

    12.
    Grieve, B. D., Hare, J. A. & Saba, V. S. Projecting the effects of climate change on Calanus finmarchicus distribution within the U.S. Northeast Continental Shelf. Sci. Rep. 7, 6264 (2017).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    13.
    Feng, Z., Ji, R., Campbell, R. G., Ashjian, C. J. & Zhang, J. Early ice retreat and ocean warming may induce copepod biogeographic boundary shifts in the Arctic Ocean. J. Geophys. Res. Ocean. 121, 6137–6158 (2016).
    ADS  Article  Google Scholar 

    14.
    Feng, Z., Ji, R., Ashjian, C., Campbell, R. & Zhang, J. Biogeographic responses of the copepod Calanus glacialis to a changing Arctic marine environment. Glob. Chang. Biol. 24, e159–e170 (2018).
    ADS  PubMed  Article  PubMed Central  Google Scholar 

    15.
    Kwok, R. et al. Thinning and volume loss of the Arctic Ocean sea ice cover: 2003–2008. J. Geophys. Res. Ocean. 114, 1–16 (2009).
    Article  Google Scholar 

    16.
    Stroeve, J., Holland, M. M., Meier, W., Scambos, T. & Serreze, M. Arctic sea ice decline: Faster than forecast. Geophys. Res. Lett. 34, 1–5 (2007).
    Article  Google Scholar 

    17.
    Notz, D. & Stroeve, J. Observed Arctic sea-ice loss directly follows anthropogenic CO2 emission. Science 354, 747–750 (2016).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    18.
    Pomerleau, C. et al. Spatial patterns in zooplankton communities across the eastern Canadian sub-Arctic and Arctic waters: insights from stable carbon (delta C-13) and nitrogen (delta N-15) isotope ratios. J. Plankton Res. 33, 1779–1792 (2011).
    CAS  Article  Google Scholar 

    19.
    Pomerleau, C., Lesage, V., Winkler, G., Rosenberg, B. & Ferguson, S. H. Contemporary diet of bowhead whales (Balaena mysticetus) from the eastern Canadian Arctic inferred from fatty acid biomarkers. Arctic 67, 84–92 (2014).
    Article  Google Scholar 

    20.
    Heide-Jørgensen, M. P. et al. Large scale sexual segregation of bowhead whales. Endang. Species Res. 13, 73–78 (2010).
    Article  Google Scholar 

    21.
    Heide-Jørgensen, M. P. et al. Winter and spring diving behavior of bowhead whales relative to prey. Anim. Biotelemetry 1, 1–15 (2013).
    Article  Google Scholar 

    22.
    Curry, B., Lee, C. M., Petrie, B., Moritz, R. E. & Kwok, R. Multiyear volume, liquid freshwater, and sea ice transports through Davis Strait, 2004–10. J. Phys. Oceanogr. 44, 1244–1266 (2014).
    ADS  Article  Google Scholar 

    23.
    Pomerleau, C. et al. Mercury and stable isotope cycles in baleen plates are consistent with year-round feeding in two bowhead whale (Balaena mysticetus) populations. Polar Biol. 41, 1881–1893 (2018).
    Article  Google Scholar 

    24.
    Doniol-Valcroze, T. et al. Abundance estimate of the Eastern Canada-West Greenland bowhead whale population based on the 2013 High Arctic Cetacean Survey. (2015).

    25.
    Frasier, T. et al. Abundance estimates of the Eastern Canada-West Greenland bowhead whale (Balaena mysticetus) population based on genetic capture-mark-recapture analyses. (2015).

    26.
    Frasier, T. R. et al. Abundance estimation from genetic mark-recapture data when not all sites are sampled: an example with the bowhead whale. Glob. Ecol. Conserv. 22, e00903 (2020).
    Article  Google Scholar 

    27.
    Dunbar, M. J. Physical oceanographic results of the ‘Calanus’ expeditions in Ungava Bay, Frobisher Bay, Cumberland Sound, Hudson Strait and Northern Hudson Bay, 1949–1955. J. Fish. Res. Board Canada 15, 155–201 (1958).
    Article  Google Scholar 

    28.
    Aitken, A. & Gilbert, R. Holocene nearshore environments and sea-level history in Pangnirtung fjord, Baffin Island, NWT, Canada. Arct. Alp. Res. 21, 34–44 (1989).
    Article  Google Scholar 

    29.
    McMeans, B. C. et al. Seasonal patterns in fatty acids of Calanus hyperboreus (Copepoda, Calanoida) from Cumberland Sound, Baffin Island, Nunavut. Mar. Biol. 159, 1095–1105 (2012).
    CAS  Article  Google Scholar 

    30.
    Bedard, J. M. et al. Outside influences on the water column of Cumberland Sound, Baffin Island. J. Geophys. Res. C Ocean. 120, 5000–5018 (2015).
    ADS  Article  Google Scholar 

    31.
    Tang, C. C. L. et al. The circulation, water masses and sea-ice of Baffin Bay. Prog. Oceanogr. 63, 183–228 (2004).
    ADS  Article  Google Scholar 

    32.
    Falk-Petersen, S., Mayzaud, P., Kattner, G. & Sargent, J. R. Lipids and life strategy of Arctic Calanus. Mar. Biol. Res. 5, 18–39 (2009).
    Article  Google Scholar 

    33.
    Davies, K. T. A., Ryan, A. & Taggart, C. T. Measured and inferred gross energy content in diapausing Calanus spp. in a Scotian shelf basin. J. Plankton Res. 34, 614–625 (2012).
    Article  Google Scholar 

    34.
    Koski, W. R., Davis, R. A., Miller, G. W. & Withrow, D. E. Reproduction. in The bowhead whale (eds. Burns, J. J., Montague, J. J. & Cowles, C. J.) 239–274 (Special Publication Number 2. The Society of Marine Mammalogy, Lawrence, KS, 1993).

    35.
    George, J. C. et al. Inferences from bowhead whale ovarian and pregnancy data: age estimates, length at sexual maturity and ovulation rates. International Whaling Commission Scientific Paper 56 (2004).

    36.
    Higdon, J. W. & Ferguson, S. H. Past, present, and future for bowhead whales (Balaena mysticetus) in northwest Hudson Bay. In A Little Less Arctic: Top Predators in the World’s Largest Northern Inland Sea, Hudson Bay (eds Ferguson, S. H. et al.) 159–177 (Springer, New York, 2010).
    Google Scholar 

    37.
    Liu, H. & Hopcroft, R. R. Growth and development of Pseudocalanus spp. in the northern Gulf of Alaska. J. Plankton Res. 30, 923–935 (2008).
    Article  Google Scholar 

    38.
    DeLorenzo Costa, A., Durbin, E. G. & Mayo, C. A. Variability in the nutritional value of the major copepods in Cape Cod Bay (Massachusetts, USA) with implications for right whales. Mar. Ecol. 27, 109–123 (2006).
    ADS  Article  CAS  Google Scholar 

    39.
    Madsen, S. D., Nielsen, T. G. & Hansen, B. W. Annual population development and production by Calanus finmarchicus, C. glacialisand C. hyperboreus in Disko Bay, western Greenland. Mar. Biol. 139, 75–93 (2001).
    Article  Google Scholar 

    40.
    Reeves, R., Mitchell, E., Mansfield, A. & McLaughlin, M. Distribution and migration of the bowhead whale, Balaena mysticetus, in the Eastern North American. Arctic 36, 60 (1983).
    Article  Google Scholar 

    41.
    Holland, C. A. William penny, 1809–92: Arctic whaling master. Polar Rec. 15, 25–43 (1970).
    Article  Google Scholar 

    42.
    Higdon, J. W. Commercial and subsistence harvests of bowhead whales (Balaena mysticetus) in eastern Canada and West Greenland. J. Cetacean Res. Manag. 11, 185–216 (2010).
    Google Scholar 

    43.
    Diemer, K. M. et al. Marine mammal and seabird summer distribution and abundance in the fjords of northeast Cumberland Sound of Baffin Island, Nunavut, Canada. Polar Biol. 34, 41–48 (2011).
    Article  Google Scholar 

    44.
    Matthews, C. et al. Boat-based surveys for marine mammals and seabirds in Cumberland Sound. Field report. (2012).

    45.
    Baumgartner, M. F., Wenzel, F. W., Lysiak, N. S. J. & Patrician, M. R. North Atlantic right whale foraging ecology and its role in human-caused mortality. Mar. Ecol. Prog. Ser. 581, 165–181 (2017).
    ADS  Article  Google Scholar 

    46.
    Fortune, S. et al. Seasonal diving and foraging behaviour of Eastern Canada-West Greenland bowhead whales. Mar. Ecol. Prog. Ser. 643, 197–217 (2020).
    ADS  Article  Google Scholar 

    47.
    Block, B. A. Physiological ecology in the 21st century: Advancements in biologging science. Integr. Comp. Biol. 45, 305–320 (2005).
    PubMed  Article  PubMed Central  Google Scholar 

    48.
    Hays, G. C. New insights: animal-borne cameras and accelerometers reveal the secret lives of cryptic species. J. Anim. Ecol. 84, 587–589 (2015).
    PubMed  Article  PubMed Central  Google Scholar 

    49.
    Bograd, S. J., Block, B. A., Costa, D. P. & Godley, B. J. Biologging technologies: new tools for conservation. Introduction. Endanger. Species Res. 10, 1–7 (2010).
    Article  Google Scholar 

    50.
    Unstad, K. H. & Tande, K. S. Depth distribution of Calanus finmarchicus and C. glacialis in relation to environmental conditions in the Barents Sea. Polar Res. 10, 409–420 (1991).
    Article  Google Scholar 

    51.
    Hirche, H. J. & Niehoff, B. Reproduction of the Arctic copepod Calanus hyperboreus in the Greenland Sea-field and laboratory observations. Polar Biol. 16, 209–219 (1996).
    Article  Google Scholar 

    52.
    Madsen, S. J., Nielsen, T. G., Tervo, O. M. & Söderkvist, J. Importance of feeding for egg production in Calanus finmarchicus and C. glacialis during the Arctic spring. Mar. Ecol. Prog. Ser. 353, 177–190 (2008).
    ADS  CAS  Article  Google Scholar 

    53.
    Darnis, G. & Fortier, L. Temperature, food and the seasonal vertical migration of key arctic copepods in the thermally stratified Amundsen Gulf (Beaufort Sea, Arctic Ocean) GE. J. Plankton Res. 36, 1092–1108 (2014).
    CAS  Article  Google Scholar 

    54.
    Parent, G. J., Plourde, S. & Turgeon, J. Overlapping size ranges of Calanus spp. off the Canadian Arctic and Atlantic Coasts: impact on species abundances. J. Plankton Res. 33, 1654–1665 (2011).
    CAS  Article  Google Scholar 

    55.
    Hyslop, E. J. Stomach contents analysis—a review of methods and their application. J. Fish Biol. 17, 411–429 (1980).
    Article  Google Scholar 

    56.
    Dunweber, M. et al. Succession and fate of the spring diatom bloom in Disko Bay, western Greenland. Mar. Ecol. Prog. Ser. 419, 11–29 (2010).
    ADS  Article  CAS  Google Scholar 

    57.
    Swalethorp, R. et al. Grazing, egg production, and biochemical evidence of differences in the life strategies of Calanus finmarchicus, C. glacialis and C. hyperboreus in Disko Bay, Western Greenland. Mar. Ecol. Prog. Ser. 429, 125–144 (2011).
    ADS  Article  Google Scholar 

    58.
    Baumgartner, M. F. & Mate, B. R. Summertime foraging ecology of North Atlantic right whales. Mar. Ecol. Prog. Ser. 264, 123–135 (2003).
    ADS  Article  Google Scholar 

    59.
    Hirche, H. J. Long-term experiments on lifespan, reproductive activity and timing of reproduction in the Arctic copepod Calanus hyperboreus. Mar. Biol. 160, 2469–2481 (2013).
    Article  Google Scholar 

    60.
    Visser, A. W. & Jónasdóttir, S. H. Lipids, buoyancy and the seasonal vertical migration of Calanus finmarchicus. Fish. Oceanogr. 8, 100–106 (1999).
    Article  Google Scholar 

    61.
    Scott, C. L., Kwasniewski, S., Falk-Petersen, S. & Sargent, J. R. Lipids and life strategies of Calanus finmarchicus, Calanus glacialis and Calanus hyperboreus in late autumn, Kongsfjorden, Svalbrad. Polar Biol. 23, 510–516 (2000).
    Article  Google Scholar 

    62.
    Heide-Jørgensen, M. P., Laidre, K. L., Logsdon, M. L. & Nielsen, T. G. Springtime coupling between chlorophyll a, sea ice and sea surface temperature in Disko Bay, West Greenland. Prog. Oceanogr. 73, 79–95 (2007).
    ADS  Article  Google Scholar 

    63.
    Baumgartner, M. F. Comparisons of Calanus finmarchicus fifth copepodite abundance estimates from nets and an optical plankton counter. J. Plankton Res. 25, 855–868 (2003).
    Article  Google Scholar 

    64.
    Herman, A. W. Design and calibration of a new optical plankton counter capable of sizing small zooplankton. Deep Sea Res. A 39, 395–415 (1992).
    ADS  Article  Google Scholar 

    65.
    Falk-Petersen, S. et al. Vertical migration in high Arctic waters during autumn 2004. Deep Sea Res. II(55), 2275–2284 (2008).
    ADS  Article  Google Scholar 

    66.
    Baumgartner, M. F., Lysiak, N. S. J., Schuman, C., Urban-Rich, J. & Wenzel, F. W. Diel vertical migration behavior of Calanus finmarchicus and its influence on right and sei whale occurrence. Mar. Ecol. Prog. Ser. 423, 167–184 (2011).
    ADS  Article  Google Scholar 

    67.
    Bollens, S. M. & Frost, B. W. Predator-induced diet vertical migration in a planktonic copepod. J. Plankton Res. 11, 1047–1065 (1989).
    Article  Google Scholar 

    68.
    Hays, G. C. Ontogenetic and seasonal variation in the diel vertical migration of the copepods Metridia lucens and Metridia longa. Limnol. Oceanogr. 40, 1461–1465 (1995).
    ADS  Article  Google Scholar 

    69.
    Huntley, M. & Brooks, E. R. Effects of age and food availability on diel vertical migration of Calanus pacificus. Mar. Biol. 71, 23–31 (1982).
    Article  Google Scholar 

    70.
    Simon, M., Johnson, M. J., Tyack, P. & Madsen, P. T. Behavior and kinematics of continous ram filtration in bowhead wahles (Balaena mysticetus). Proc. R. Soc. Lond. B. 276, 3819–3828 (2009).
    Article  Google Scholar 

    71.
    van der Hoop, J. M. et al. Foraging rates of ram-filtering North Atlantic right whales. Funct. Ecol. 33, 1290–1306 (2019).
    Article  Google Scholar 

    72.
    Goldbogen, J. A. et al. Prey density and distribution drive the three-dimensional foraging strategies of the largest filter feeder. Funct. Ecol. 29, 951–961 (2015).
    Article  Google Scholar 

    73.
    Kooyman, G. L., Wahrenbrock, E. A., Castellini, M. A., Davis, R. W. & Sinnett, E. E. Aerobic and anaerobic metabolism during voluntary diving in Weddell seals: evidence of preferred pathways from blood chemsitry and behavior. J. Comp. Physiol. B 138, 335–346 (1980).
    CAS  Article  Google Scholar 

    74.
    Kooyman, G. L., Castellini, M. A., Davis, R. W. & Maue, R. A. Aerobic diving limits of immature Weddell seals. J. Comp. Physiol. B 151, 171–174 (1983).
    Article  Google Scholar 

    75.
    Dyke, A. S., Hooper, J. & Savelle, J. M. A history of sea ice in the Canadian Arctic archipelago based on postglacial remains of the bowhead whale (Balaena mysticetus). Arctic 49, 235–255 (1996).
    Article  Google Scholar 

    76.
    Baumgartner, M. F., Hammar, T. & Robbins, J. Development and assessment of a new dermal attachment for short-term tagging studies of baleen whales. Methods Ecol. Evol. 6, 289–297 (2015).
    Article  Google Scholar 

    77.
    Reinhart, N. R. et al. Occurrence of killer whale Orcinus orca rake marks on Eastern Canada-West Greenland bowhead whales Balaena mysticetus. Polar Biol. 36, 1133–1146 (2013).
    Article  Google Scholar 

    78.
    Fortune, S. M. E. et al. Evidence of molting and the function of “rock-nosing” behavior in bowhead whales in the eastern Canadian Arctic. PLoS ONE 12, 1–15 (2017).
    MathSciNet  Article  CAS  Google Scholar 

    79.
    Silva, M. A. et al. Assessing performance of Bayesian state-space models fit to argos satellite telemetry locations processed with kalman filtering. PLoS ONE 9, e92277 (2014).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    80.
    Lowther, A. D., Lydersen, C., Fedak, M. A., Lovell, P. & Kovacs, K. M. The argos-CLS kalman filter: Error structures and state-space modelling relative to fastloc GPS data. PLoS ONE 10, e0124754 (2015).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    81.
    R Development Core Team. R: A Language and Environment for Statistical Computing. R Development Core Team, Vienna (2016). https://doi.org/10.1038/sj.hdy.6800737.

    82.
    Jonsen, I. D., Flemming, J. M. & Myers, R. A. Robust state-space modeling of animal movement data. Ecology 86, 2874–2880 (2005).
    Article  Google Scholar 

    83.
    Jonsen, I. D. et al. State-space models for bio-loggers: a methodological road map. Deep. Res. II(88–89), 34–46 (2013).
    ADS  Google Scholar 

    84.
    Tinbergen, N., Impekoven, M. & Franck, D. An experiment on spacing-out as a defence against predation. Behaviour 28, 307–320 (1967).
    Article  Google Scholar 

    85.
    Kareiva, P. & Odell, G. Swarms of predators exhibit ‘preytaxis’ if individual predators use area-restricted search. Am. Nat. 130, 233–270 (1987).
    Article  Google Scholar 

    86.
    Haskell, D. G. Experiments and a model examining learning in the area-restricted search behavior of ferrets (Mustela putorius furo). Behav. Ecol. 8, 448–455 (1997).
    Article  Google Scholar 

    87.
    Fauchald, P. & Tveraa, T. Using first-passage time in the analysis of area-restricted search and habitat selection. Ecology 84, 282–288 (2003).
    Article  Google Scholar 

    88.
    Anderwald, P. et al. Spatial scale and environmental determinants in minke whale habitat use and foraging. Mar. Ecol. Prog. Ser. 450, 259–274 (2012).
    ADS  Article  Google Scholar 

    89.
    Jonsen, I. D., Myers, R. A. & James, M. C. Identifying leatherback turtle foraging behaviour from satellite telemetry using a switching state-space model. Mar. Ecol. Prog. Ser. 337, 255–264 (2007).
    ADS  Article  Google Scholar 

    90.
    Pinheiro, J. C. & Bates, D. M. Linear mixed-effects models. in Mixed-effects models in S and S-Plus 1–56 (Springer, New York, 2000). https://doi.org/10.1198/tech.2001.s574.

    91.
    Sverdrup, H. U. On conditions for the vernal blooming of phytoplankton. ICES J. Mar. Sci. 18, 287–295 (1953).
    Article  Google Scholar 

    92.
    Thomson, R. E. & Fine, I. V. Estimating mixed layer depth from oceanic profile data. J. Atmos. Ocean. Technol. 20, 319–329 (2003).
    ADS  Article  Google Scholar 

    93.
    Smith, W. O. & Jones, R. M. Vertical mixing, critical depths, and phytoplankton growth in the Ross Sea. ICES J. Mar. Sci. 72, 1952–1960 (2015).
    Article  Google Scholar 

    94.
    Suthers, I. M., Taggart, C. T., Rissik, D. & Baird, M. E. Day and night ichthyoplankton assemblages and zooplankton biomass size spectrum in a deep ocean island wake. Mar. Ecol. Prog. Ser. 322, 225–238 (2006).
    ADS  CAS  Article  Google Scholar 

    95.
    Grainger, E. H. The copepods Calanus glacial is Jaschnov and Calanus finmarchicus (Gunnerus) in Canadian Arctic-Subarctic waters. J. Fish. Res. Board Can. 18, 663–678 (1961).
    Article  Google Scholar 

    96.
    Jaschnov, W. A. Distribution of Calanus Species in the Seas of the Northern Hemisphere. Int. Rev. Hydrobiol. Hydrogr. 55, 197–212 (1970).
    Article  Google Scholar 

    97.
    Hirche, H. J. & Mumm, N. Distribution of dominant copepods in the Nansen Basin, Arctic Ocean, in summer. Deep Sea Res. A 39, 485–505 (1992).
    ADS  Article  Google Scholar 

    98.
    Breteler, W. C. M. K., Fransz, H. G. & Gonzalez, S. R. Growth and development of four calanoid copepod species under experimental and natural conditions. Neth. J. Sea Res. 16, 195–207 (1982).
    Article  Google Scholar  More

  • in

    Analyzing long-term impacts of ungulate herbivory on forest-recruitment dynamics at community and species level contrasting tree densities versus maximum heights

    1.
    Crawley, M. Herbivory: The Dynamics of Animal–Plant Interactions (Blackwell Scientific, Oxford, 1983).
    Google Scholar 
    2.
    Putman, R. Grazing in Temperate Ecosystems: Large Herbivores and the Ecology of the New Forest (Springer, Berlin, 1986).
    Google Scholar 

    3.
    Huntly, N. Herbivores and the dynamics of communities and ecosystems. Annu. Rev. Ecol. Syst. 1, 477–503 (1991).
    Article  Google Scholar 

    4.
    Skarpe, C. Impact of grazing in savanna ecosystems. Ambio 20, 351–356 (1991).
    Google Scholar 

    5.
    Agrawal, A. A. Macroevolution of plant defense strategies. Trends Ecol. Evol. 22, 103–109 (2007).
    PubMed  Article  Google Scholar 

    6.
    Bruce, T. C. Interplay between insects and plants–dynamic and complex interactions that have coevolved over millions of years but act in milliseconds. J. Exp. Bot. 2, 391 (2014).
    Google Scholar 

    7.
    Mason, N. W. H., Peltzer, D. A., Richardson, S. J., Bellingham, P. J. & Allen, R. B. Stand development moderates effects of ungulate exclusion on foliar traits in the forests of New Zealand: Ungulate impacts on foliar traits. J. Ecol. 98, 1422–1433 (2010).
    Article  Google Scholar 

    8.
    Faison, E. K., DeStefano, S., Foster, D. R., Motzkin, G. & Rapp, J. M. Ungulate browsers promote herbaceous layer diversity in logged temperate forests. Ecol. Evol. 6, 4591–4602 (2016).
    PubMed  PubMed Central  Article  Google Scholar 

    9.
    Simončič, T., Bončina, A., Jarni, K. & Klopčič, M. Assessment of the long-term impact of deer on understory vegetation in mixed temperate forests. J. Veg. Sci. 30, 108–120 (2019).
    Article  Google Scholar 

    10.
    Schmitz, O. J. Herbivory from individuals to ecosystems. Annu. Rev. Ecol. Evol. Syst. 39, 133–152 (2008).
    Article  Google Scholar 

    11.
    Riginos, C. & Grace, J. B. Savanna tree density, herbivores, and the herbaceous community: Bottom-up vs. top-down effects. Ecology 89, 2228–2238 (2008).
    PubMed  Article  Google Scholar 

    12.
    Turkington, R. Top-down and bottom-up forces in mammalian herbivore–vegetation systems: An essay review. Botany 87, 723–739 (2009).
    Article  Google Scholar 

    13.
    Kos, M. et al. Relative importance of plant-mediated bottom-up and top-down forces on herbivore abundance on Brassica oleracea: Bottom-up and top-down effects on herbivores. Funct. Ecol. 25, 1113–1124 (2011).
    Article  Google Scholar 

    14.
    Kuijper, D. P. J. et al. Bottom-up versus top-down control of tree regeneration in the Białowieża Primeval Forest, Poland: Abiotic and biotic control of tree regeneration. J. Ecol. 98, 888–899 (2010).
    Article  Google Scholar 

    15.
    Churski, M., Bubnicki, J. W., Jędrzejewska, B., Kuijper, D. P. J. & Cromsigt, J. P. G. M. Brown world forests: Increased ungulate browsing keeps temperate trees in recruitment bottlenecks in resource hotspots. New Phytol. 214, 158–168 (2017).
    PubMed  Article  Google Scholar 

    16.
    Fretwell, S. D. Food chain dynamics: The central theory of ecology?. Oikos 50, 291–301 (1987).
    Article  Google Scholar 

    17.
    Reimoser, F. & Putman, R. Impacts of wild ungulates on vegetation: Costs and benefits. In Ungulate Management in Europe—Problems and Practices (eds Putman, R. et al.) 144–191 (Cambridge University Press, Cambridge, 2011).
    Google Scholar 

    18.
    Bellingham, P. J. & Allan, C. N. Forest regeneration and the influences of white-tailed deer (Odocoileus virginianus) in cool temperate New Zealand rain forests. For. Ecol. Manag. 175, 71–86 (2003).
    Article  Google Scholar 

    19.
    Russell, F. L. & Fowler, N. L. Effects of white-tailed deer on the population dynamics of acorns, seedlings and small saplings of Quercus buckleyi. Plant Ecol. 173, 59–72 (2004).
    Article  Google Scholar 

    20.
    Casabon, C. & Pothier, D. Browsing of tree regeneration by white-tailed deer in large clearcuts on Anticosti Island, Quebec. For. Ecol. Manag. 253, 112–119 (2007).
    Article  Google Scholar 

    21.
    Pellerin, M. et al. Impact of deer on temperate forest vegetation and woody debris as protection of forest regeneration against browsing. For. Ecol. Manag. 260, 429–437 (2010).
    Article  Google Scholar 

    22.
    Tschöpe, O., Wallschläger, D., Burkart, M. & Tielbörger, K. Managing open habitats by wild ungulate browsing and grazing: A case-study in North-Eastern Germany: Managing open habitats by wild ungulate browsing and grazing. Appl. Veg. Sci. 14, 200–209 (2011).
    Article  Google Scholar 

    23.
    Millett, J. & Edmondson, S. The impact of 36 years of grazing management on vegetation dynamics in dune slacks. J. Appl. Ecol. 50, 1367–1376 (2013).
    Article  Google Scholar 

    24.
    Beck, H., Snodgrass, J. W. & Thebpanya, P. Long-term exclosure of large terrestrial vertebrates: Implications of defaunation for seedling demographics in the Amazon rainforest. Biol. Conserv. 163, 115–121 (2013).
    Article  Google Scholar 

    25.
    Charles, G. K., Porensky, L. M., Riginos, C., Veblen, K. E. & Young, T. P. Herbivore effects on productivity vary by guild: Cattle increase mean productivity while wildlife reduce variability. Ecol. Appl. 27, 143–155 (2017).
    PubMed  Article  Google Scholar 

    26.
    Castleberry, S. B., Ford, W. M., Miller, K. V. & Smith, W. P. Influences of herbivory and canopy opening size on forest regeneration in a southern bottomland hardwood forest. For. Ecol. Manag. 131, 57–64 (2000).
    Article  Google Scholar 

    27.
    Filazzola, A., Tanentzap, A. J. & Bazely, D. R. Estimating the impacts of browsers on forest understories using a modified index of community composition. For. Ecol. Manag. 313, 10–16 (2014).
    Article  Google Scholar 

    28.
    Nishizawa, K., Tatsumi, S., Kitagawa, R. & Mori, A. S. Deer herbivory affects the functional diversity of forest floor plants via changes in competition-mediated assembly rules. Ecol. Res. 31, 569–578 (2016).
    CAS  Article  Google Scholar 

    29.
    Boulanger, V. et al. Ungulates increase forest plant species richness to the benefit of non-forest specialists. Glob. Change Biol. 24, e485–e495 (2018).
    Article  Google Scholar 

    30.
    McGarvey, J. C., Bourg, N. A., Thompson, J. R., McShea, W. J. & Shen, X. Effects of twenty years of deer exclusion on woody vegetation at three life-history stages in a mid-atlantic temperate deciduous forest. Northeast. Nat. 20, 451–468 (2013).
    Article  Google Scholar 

    31.
    Kabeya, D. & Sakai, S. The role of roots and cotyledons as storage organs in early stages of establishment in Quercus crispula: a quantitative analysis of the nonstructural carbohydrate in cotyledons and roots. Ann. Bot. 92, 537–545 (2003).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    32.
    Boege, K. & Marquis, R. J. Facing herbivory as you grow up: The ontogeny of resistance in plants. Trends Ecol. Evol. 20, 441–448 (2005).
    PubMed  Article  Google Scholar 

    33.
    Hanley, M. E., Lamont, B. B., Fairbanks, M. M. & Rafferty, C. M. Plant structural traits and their role in anti-herbivore defence. Perspect. Plant Ecol. Evol. Syst. 8, 157–178 (2007).
    Article  Google Scholar 

    34.
    Diggle, P. J. Statistical Analysis of Spatial Point Patterns. (Arnold, 2003).

    35.
    Gratzer, G. & Waagepetersen, R. Seed dispersal, microsites or competition—What drives gap regeneration in an old-growth forest? An application of spatial point process modelling. Forests 9, 230 (2018).
    Article  Google Scholar 

    36.
    Szwagrzyk, J., Gratzer, G., Stępniewska, H., Szewczyk, J. & Veselinovic, B. High reproductive effort and low recruitment rates of European beech: Is there a limit for the superior competitor?. Pol. J. Ecol. 63, 198–212 (2015).
    Article  Google Scholar 

    37.
    Nopp-Mayr, U., Kempter, I., Muralt, G. & Gratzer, G. Herbivory on young tree seedlings in old-growth and managed mountain forests. Ecol. Res. 30, 479–491 (2015).
    CAS  Article  Google Scholar 

    38.
    Shugart, H. H. A theory of forest dynamics. (Springer, 1984).

    39.
    Lertzman, K. B. Patterns of gap-phase replacement in a subalpine, old-growth forest. Ecology 73, 657–669 (1992).
    Article  Google Scholar 

    40.
    Kneeshaw, D. D. & Bergeron, Y. Canopy gap characteristics and tree replacement in the Southeastern Boreal forest. Ecology 79, 783–794 (1998).
    Article  Google Scholar 

    41.
    Wakeling, J. L., Staver, A. C. & Bond, W. J. Simply the best: The transition of savanna saplings to trees. Oikos 120, 1448–1451 (2011).
    Article  Google Scholar 

    42.
    Kobe, R. K., Pacala, S. W., Silander, J. A. Jr. & Canham, C. D. Juvenile tree survivorship as a component of shade tolerance. Ecol. Appl. 5, 517–532 (1995).
    Article  Google Scholar 

    43.
    Zuidema, P. A., Brienen, R. J. W., During, H. J. & Güneralp, B. Do persistently fast-growing juveniles contribute disproportionately to population growth? A new analysis tool for matrix models and its application to rainforest trees. Am. Nat. 174, 709–719 (2009).
    PubMed  Article  Google Scholar 

    44.
    Tremblay, J.-P., Huot, J. & Potvin, F. Density-related effects of deer browsing on the regeneration dynamics of boreal forests. J. Appl. Ecol. 44, 552–562 (2007).
    Article  Google Scholar 

    45.
    Speed, J. D. M., Austrheim, G., Hester, A. J., Solberg, E. J. & Tremblay, J.-P. Regional-scale alteration of clear-cut forest regeneration caused by moose browsing. For. Ecol. Manag. 289, 289–299 (2013).
    Article  Google Scholar 

    46.
    Shelton, A. L., Henning, J. A., Schultz, P. & Clay, K. Effects of abundant white-tailed deer on vegetation, animals, mycorrhizal fungi, and soils. For. Ecol. Manag. 320, 39–49 (2014).
    Article  Google Scholar 

    47.
    Reimoser, F. & Reimoser, S. Ergebnisse aus dem Vergleichsflächenverfahren (‘Wildschaden-Kontrollzäune’) – ein Beitrag zur Objektivierung der Wildschadensbeurteilung. In Ist die natürliche Verjüngung des Bergwaldes gesichert? (ed. Müller, F.) 151–159 (Austrian Research Centre for Forests, Vienna, 2003).
    Google Scholar 

    48.
    ZAMG. Klimadaten von Österreich 1971–2000. (2013).

    49.
    Mucina, L., Grabherr, G. & Wallnöfer, S. Die Pflanzengesellschaften Österreichs. Teil III – Wälder und Gebüsche (Gustav Fischer Verlag, Stuttgart, 1993).
    Google Scholar 

    50.
    Reimoser, F., Schodterer, H. & Reimoser, S. Beurteilung des Schalenwildeinflusses auf die Waldverjüngung – Vergleich verschiedener Methoden des Wildeinfluss-Monitorings („WEM – Methodenvergleich”) (Austrian Research Centre for Forests, Vienna, 2014).
    Google Scholar 

    51.
    Reimoser, F., Armstrong, H. & Suchant, R. Measuring forest damage of ungulates: What should be considered. For. Ecol. Manag. 120, 47–58 (1999).
    Article  Google Scholar 

    52.
    Long, Z. T., Pendergast, T. H. & Carson, W. P. The impact of deer on relationships between tree growth and mortality in an old-growth beech-maple forest. For. Ecol. Manag. 252, 230–238 (2007).
    Article  Google Scholar 

    53.
    Van den Brink, P. J. & Ter Braak, C. J. Principal response curves: Analysis of time-dependent multivariate responses of biological community to stress. Environ. Toxicol. Chem. 18, 138–148 (1999).
    Article  Google Scholar 

    54.
    van den Brink, P. J., den Besten, P. J., de Bij, V. A. & ter Braak, C. J. F. Principal response curves technique for the analysis of multivariate biomonitoring time series. Environ. Monit. Assess. 152, 271–281 (2009).
    CAS  PubMed  Article  Google Scholar 

    55.
    Poulin, M., Andersen, R. & Rochefort, L. A new approach for tracking vegetation change after restoration: A case study with peatlands. Restor. Ecol. 21, 363–371 (2013).
    Article  Google Scholar 

    56.
    Borcard, D., Gillet, F. & Legendre, P. Numerical Ecology with R (Springer International Publishing, Berlin, 2018).
    Google Scholar 

    57.
    Van den Brink, P. J., Van den Brink, N. W. & Ter Braak, C. J. Multivariate analysis of ecotoxicological data using ordination: demonstrations of utility on the basis of various examples. Austr. J. Ecotoxicol. 9, 141–156 (2003).
    Google Scholar 

    58.
    R Development Core Team. R: A language and environment for statistical computing. (R Foundation for Statistical Computing, 2020).

    59.
    Oksanen, J. et al. vegan: Community Ecology Package. (2019).

    60.
    RStudio Team. RStudio: Integrated Development Environment for R. (RStudio, Inc., 2016).

    61.
    Wickham, H., François, R., Henry, L. & Müller, K. dplyr: A Grammar of Data Manipulation. (2019).

    62.
    Wickham, H. forcats: Tools for Working with Categorical Variables (Factors). (2018).

    63.
    Wickham, H. ggplot2: Elegant Graphics for Data Analysis (Springer, Berlin, 2016).
    Google Scholar 

    64.
    Henry, L. & Wickham, H. purrr: Functional Programming Tools. (2019).

    65.
    Wickham, H., Hester, J. & Francois, R. readr: Read Rectangular Text Data. (2018).

    66.
    Wickham, H. stringr: Simple, Consistent Wrappers for Common String Operations. (2019).

    67.
    Müller, K. & Wickham, H. tibble: Simple Data Frames. (2019).

    68.
    Wickham, H. & Henry, L. tidyr: Easily Tidy Data with ‘spread()’ and ‘gather()’ Functions. (2019).

    69.
    McNamara, A., Rubia, E. A. de la, Zhu, H., Ellis, S. & Quinn, M. skimr: Compact and Flexible Summaries of Data. (2019).

    70.
    Allaire, J. J., Wickham, H., Ushey, K. & Ritchie, G. rstudioapi: Safely Access the RStudio API. (2017).

    71.
    Allaire, J. J. et al. rmarkdown: Dynamic Documents for R. (2018).

    72.
    Xie, Y. knitr: A General-Purpose Package for Dynamic Report Generation in R. (2018).

    73.
    Baumgartner, J. hues: Distinct Colours Palettes Based on ‘iwanthue’. (2017).

    74.
    Jones, O. R. et al. Diversity of ageing across the tree of life. Nature 505, 169–173 (2014).
    ADS  CAS  PubMed  Article  Google Scholar 

    75.
    Pacala, S. W. et al. Forest models defined by field measurements: Estimation error analysis and dynamics. Ecol. Monogr. 66, 1–43 (1996).
    Article  Google Scholar 

    76.
    Beckage, B. & Clark, J. S. Seedling survival and growth of three forest tree species: The role of spatial heterogeneity. Ecology 84, 1849–1861 (2003).
    Article  Google Scholar 

    77.
    Peltzer, D. A. et al. Disentangling drivers of tree population size distributions. For. Ecol. Manag. 331, 165–179 (2014).
    Article  Google Scholar 

    78.
    Reimoser, F., Odermatt, O., Roth, R. & Suchant, R. Die Beurteilung von Wildverbiss durch SOLL-IST-Vergleich. Allg Forst Jagdztg 168, 214–227 (1997).
    Google Scholar 

    79.
    Pépin, D. et al. Relative impact of browsing by red deer on mixed coniferous and broad-leaved seedlings—An enclosure-based experiment. For. Ecol. Manag. 222, 302–313 (2006).
    Article  Google Scholar 

    80.
    Jurena, P. N. & Archer, S. Woody plant establishment and spatial heterogeneity in grasslands. Ecology 84, 907–919 (2003).
    Article  Google Scholar 

    81.
    Cramer, M. D., Chimphango, S. B. M., Cauter, A. V., Waldram, M. S. & Bond, W. J. Grass competition induces N2 fixation in some species of African Acacia. J. Ecol. 95, 1123–1133 (2007).
    CAS  Article  Google Scholar 

    82.
    Hegland, S. J., Lilleeng, M. S. & Moe, S. R. Old-growth forest floor richness increases with red deer herbivory intensity. For. Ecol. Manag. 310, 267–274 (2013).
    Article  Google Scholar 

    83.
    Lilleeng, M. S., Hegland, S. J., Rydgren, K. & Moe, S. R. Red deer mediate spatial and temporal plant heterogeneity in boreal forests. Ecol. Res. 31, 777–784 (2016).
    Article  Google Scholar 

    84.
    Laurent, L., Mårell, A., Balandier, P., Holveck, H. & Saïd, S. Understory vegetation dynamics and tree regeneration as affected by deer herbivory in temperate hardwood forests. IForest – Biogeosciences For. 10, 837–844 (2017).
    Article  Google Scholar 

    85.
    Holladay, C.-A., Kwit, C. & Collins, B. Woody regeneration in and around aging southern bottomland hardwood forest gaps: Effects of herbivory and gap size. For. Ecol. Manag. 223, 218–225 (2006).
    Article  Google Scholar 

    86.
    Smit, C., Gusberti, M. & Müller-Schärer, H. Safe for saplings; safe for seeds?. For. Ecol. Manag. 237, 471–477 (2006).
    Article  Google Scholar 

    87.
    Pröll, G., Darabant, A., Gratzer, G. & Katzensteiner, K. Unfavourable microsites, competing vegetation and browsing restrict post-disturbance tree regeneration on extreme sites in the Northern Calcareous Alps. Eur. J. For. Res. 134, 293–308 (2015).
    Article  Google Scholar 

    88.
    Stephan, J. G. et al. Long-term deer exclosure alters soil properties, plant traits, understorey plant community and insect herbivory, but not the functional relationships among them. Oecologia 184, 685–699 (2017).
    ADS  PubMed  PubMed Central  Article  Google Scholar 

    89.
    Hidding, B., Tremblay, J.-P. & Côté, S. D. A large herbivore triggers alternative successional trajectories in the boreal forest. Ecology 94, 2852–2860 (2013).
    PubMed  Article  Google Scholar 

    90.
    Augustine, D. J. & McNaughton, S. J. Ungulate effects on the functional species composition of plant communities: Herbivore selectivity and plant tolerance. J. Wildl. Manag. 62, 1165–1183 (1998).
    Article  Google Scholar 

    91.
    Owen-Smith, N. R. Adaptive Herbivore Ecology. From Resources to Populations in Variable Environments (Cambridge University Press, Cambridge, 2002).
    Google Scholar 

    92.
    Reimoser, F. & Reimoser, S. Richtiges Erkennen von Wildschäden am Wald (Zentralstelle Österr, Landesjagdverbände, 2017).
    Google Scholar 

    93.
    Ramirez, J. I. et al. Above- and below-ground cascading effects of wild ungulates in temperate forests. Ecosystems https://doi.org/10.1007/s10021-020-00509-4 (2020).
    Article  Google Scholar 

    94.
    Kral, F. Spät- und postglaziale Waldgeschichte der Alpen aufgrund der bisherigen Pollenanalysen (Österreichischer Agrarverlag, Vienna, 1979).
    Google Scholar 

    95.
    Mayer, H. & Ott, E. Gebirgswaldbau, Schutzwaldpflege: ein waldbaulicher Beitrag zur Landschaftsökologie und zum Umweltschutz (G. Fischer, Mumbai, 1991).
    Google Scholar 

    96.
    Mayer, M., Keßler, D. & Katzensteiner, K. Herbivory modulates soil CO2 fluxes after windthrow: A case study in temperate mountain forests. Eur. J. For. Res. 139, 383–391 (2020).
    CAS  Article  Google Scholar  More