More stories

  • in

    Bayesian Network Analysis reveals resilience of the jellyfish Aurelia aurita to an Irish Sea regime shift

    1.
    Lynam, C. P. et al. Have jellyfish in the Irish Sea benefited from climate change and overfishing?. Glob. Change Biol. 17, 767–782 (2011).
    ADS  Article  Google Scholar 
    2.
    Condon, R. H. et al. Recurrent jellyfish blooms are a consequence of global oscillations. Proc. Natl. Acad. Sci. 110, 1000–1005 (2013).
    ADS  CAS  PubMed  Article  Google Scholar 

    3.
    Lynam, C. P., Hay, S. J. & Brierley, A. S. Jellyfish abundance and climatic variation: Contrasting responses in oceanographically distinct regions of the North Sea, and possible implications for fisheries. J. Mar. Biol. Assoc. 85, 435–450 (2005).
    Article  Google Scholar 

    4.
    Lynam, C. P. et al. Jellyfish overtake fish in a heavily fished ecosystem. Curr. Biol. 16, R492-493 (2006).
    CAS  PubMed  Article  Google Scholar 

    5.
    Hays, G. C., Doyle, T. K. & Houghton, J. D. R. A paradigm shift in the trophic importance of jellyfish?. Trends Ecol. Evol. 33, 874–884 (2018).
    PubMed  Article  Google Scholar 

    6.
    Gibbons, M. J. & Richardson, A. J. Patterns of jellyfish abundance in the North Atlantic. In Jellyfish Blooms: Causes, Consequences, and Recent Advances: Proceedings of the Second International Jellyfish Blooms Symposium, held at the Gold Coast, Queensland, Australia, 24–27 June, 2007 (eds. Pitt, K. A. & Purcell, J. E.) 51–65 (Springer Netherlands, 2009). https://doi.org/10.1007/978-1-4020-9749-2_4.

    7.
    Attrill, M. J., Wright, J. & Edwards, M. Climate-related increases in jellyfish frequency suggest a more gelatinous future for the North Sea. Limnol. Oceanogr. 52, 480–485 (2007).
    ADS  Article  Google Scholar 

    8.
    Brodeur, R. D., Sugisaki, H. & Hunt, G. L. Jr. Increases in jellyfish biomass in the Bering Sea: Implications for the ecosystem. Mar. Ecol. Prog. Ser. 233, 89–103 (2002).
    ADS  Article  Google Scholar 

    9.
    Pauly, D., Christensen, V., Dalsgaard, J., Froese, R. & Torres, F. Fishing down marine food webs. Science 279, 860–863 (1998).
    ADS  CAS  PubMed  Article  Google Scholar 

    10.
    Purcell, J. E. & Arai, M. N. Interactions of pelagic cnidarians and ctenophores with fish: A review. Hydrobiologia 451, 27–44 (2001).
    Article  Google Scholar 

    11.
    Robinson, K. L. et al. Jellyfish, forage fish, and the world’s major fisheries. Oceanography 27, 104–115 (2014).
    Article  Google Scholar 

    12.
    Uye, S. Blooms of the giant jellyfish Nemopilema nomurai: A threat to the fisheries sustainability of the East Asian Marginal Seas. Plankton Benthos Res. 3, 125–131 (2008).
    Article  Google Scholar 

    13.
    Wright, R. M., Le Quéré, C., Buitenhuis, E., Pitois, S. & Gibbons, M. Unique role of jellyfish in the plankton ecosystem revealed using a global ocean biogeochemical model. Biogeosci. Discuss. https://doi.org/10.5194/bg-2020-136 (2020).
    Article  Google Scholar 

    14.
    Jackson, J. B. C. Ecological extinction and evolution in the brave new ocean. Proc. Natl. Acad. Sci. 105, 11458–11465 (2008).
    ADS  CAS  PubMed  Article  Google Scholar 

    15.
    Kintner, A. & Brierley, A. S. Cryptic hydrozoan blooms pose risks to gill health in farmed North Atlantic salmon (Salmo salar). J. Mar. Biol. Assoc. 99, 539–550 (2019).
    Article  Google Scholar 

    16.
    Flynn, B. A. et al. Temporal and spatial patterns in the abundance of jellyfish in the northern Benguela upwelling ecosystem and their link to thwarted pelagic fishery recovery. Afr. J. Mar. Sci. 34, 131–146 (2012).
    Article  Google Scholar 

    17.
    Luo, J. Y. et al. Gelatinous zooplankton-mediated carbon flows in the global oceans: A data-driven modeling study. Glob. Biogeochem. Cycles 34, e2020GB006704 (2020).
    ADS  CAS  Article  Google Scholar 

    18.
    Behrenfeld, M. J. et al. Climate-driven trends in contemporary ocean productivity. Nature 444, 752–755 (2006).
    ADS  CAS  PubMed  Article  Google Scholar 

    19.
    Hays, G. C., Richardson, A. J. & Robinson, C. Climate change and marine plankton. Trends Ecol. Evol. 20, 337–344 (2005).
    PubMed  Article  Google Scholar 

    20.
    Richardson, A. J. & Schoeman, D. S. Climate impact on plankton ecosystems in the northeast Atlantic. Science 305, 1609–1612 (2004).
    ADS  CAS  PubMed  Article  Google Scholar 

    21.
    Suikkanen, S. et al. Climate change and eutrophication induced shifts in northern summer plankton communities. PLoS ONE 8, e66475 (2013).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    22.
    Wiafe, G., Yaqub, H. B., Mensah, M. A. & Frid, C. L. J. Impact of climate change on long-term zooplankton biomass in the upwelling region of the Gulf of Guinea. ICES J. Mar. Sci. 65, 318–324 (2008).
    Article  Google Scholar 

    23.
    Reid, P. C., Colebrook, J. M., Matthews, J. B. L. & Aiken, J. The Continuous Plankton Recorder: Concepts and history, from Plankton Indicator to undulating recorders. Prog. Oceanogr. 58, 117–173 (2003).
    ADS  Article  Google Scholar 

    24.
    Edwards, M. et al. Plankton, jellyfish and climate in the North-East Atlantic. MCCIP Sci. Rev. 2020, 322–353. https://doi.org/10.14465/2020.arc15.plk (2020).
    Article  Google Scholar 

    25.
    ICES. Report of the Working Group on the Celtic Seas Ecoregion (WGCSE), 11–19 May 2011, Copenhagen, Denmark. (2011).

    26.
    Bartolino, V. et al. Herring assessment working group for the area south of 62° N (HAWG). (2019) https://doi.org/10.17895/ices.pub.5460.

    27.
    ICES Advice Book 5. https://www.ices.dk/sites/pub/Publication%20Reports/Advice/2007/may/her-nirs.pdf (2007).

    28.
    Beaugrand, G. The North Sea regime shift: Evidence, causes, mechanisms and consequences. Prog. Oceanogr. 60, 245–262 (2004).
    ADS  Article  Google Scholar 

    29.
    Gregory, B., Christophe, L. & Martin, E. Rapid biogeographical plankton shifts in the North Atlantic Ocean. Glob. Change Biol. 15, 1790–1803 (2009).
    ADS  Article  Google Scholar 

    30.
    deYoung, B. et al. Regime shifts in marine ecosystems: Detection, prediction and management. Trends Ecol. Evol. 23, 402–409 (2008).
    PubMed  Article  Google Scholar 

    31.
    Bastian, T. et al. Large-scale sampling reveals the spatio-temporal distributions of the jellyfish Aurelia aurita and Cyanea capillata in the Irish Sea. Mar. Biol. 158, 2639–2652 (2011).
    Article  Google Scholar 

    32.
    Houghton, J. D. R., Doyle, T. K., Davenport, J. & Hays, G. C. Developing a simple, rapid method for identifying and monitoring jellyfish aggregations from the air. Mar. Ecol. Prog. Ser. 314, 159–170 (2006).
    ADS  Article  Google Scholar 

    33.
    Bastian, T., Lilley, M. K. S., Beggs, S. E., Hays, G. C. & Doyle, T. K. Ecosystem relevance of variable jellyfish biomass in the Irish Sea between years, regions and water types. Estuar. Coast. Shelf Sci. 149, 302–312 (2014).
    ADS  Article  Google Scholar 

    34.
    Heckerman, D., Geiger, D. & Chickering, D. M. Learning Bayesian networks: The combination of knowledge and statistical data. Mach. Learn. 20, 197–243 (1995).
    MATH  Google Scholar 

    35.
    Milns, I., Beale, C. M. & Smith, V. A. Revealing ecological networks using Bayesian network inference algorithms. Ecology 91, 1892–1899 (2010).
    PubMed  Article  Google Scholar 

    36.
    Mitchell, E. G. & Neutel, A.-M. Feedback spectra of soil food webs across a complexity gradient, and the importance of three-species loops to stability. Theor. Ecol. 5, 153–159 (2012).
    Article  Google Scholar 

    37.
    Olff, H. et al. Parallel ecological networks in ecosystems. Philos. Trans. R. Soc. B Biol. Sci. 364, 1755–1779 (2009).
    Article  Google Scholar 

    38.
    Mitchell, E. G., Whittle, R. & Griffths, H. J. Benthic ecosystem cascade effects in Antarctica using Bayesian network inference. Commun. Biol. 3, 582 (2020).
    PubMed  PubMed Central  Article  Google Scholar 

    39.
    Yu, J., Smith, V. A., Wang, P. P., Hartemink, A. J. & Jarvis, E. D. Advances to Bayesian network inference for generating causal networks from observational biological data. Bioinformatics 20, 3594–3603 (2004).
    CAS  PubMed  Article  Google Scholar 

    40.
    Yu, J., Smith, V. A., Wang, P. P., Hartemink, E. J. & Jarvis, E. D. Using Bayesian network inference algorithms to recover molecular genetic regulatory networks. In Prof. of Int. (2002).

    41.
    Smith, V. A., Yu, J., Smulders, T. V., Hartemink, A. J. & Jarvis, E. D. Computational inference of neural information flow networks. PLOS Comput. Biol. 2, e161 (2006).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    42.
    Mitchell, E. G. & Butterfield, N. J. Spatial analyses of Ediacaran communities at Mistaken Point. Paleobiology 44, 40–57 (2018).
    Article  Google Scholar 

    43.
    Mitchell, E. G., Durden, J. M. & Ruhl, H. A. First network analysis of interspecific associations of abyssal benthic megafauna reveals potential vulnerability of abyssal hill community. Prog. Oceanogr. 187, 102401 (2020).
    Article  Google Scholar 

    44.
    Mitchell, E. G. & Harris, S. Mortality, population and community dynamics of the glass sponge dominated community “The Forest of the Weird” from the RIDGE seamount, Johnston Atoll, Pacific Ocean. Front. Mar. Sci. 7, 872 (2020).
    Article  Google Scholar 

    45.
    Reid, P. C., Borges, M. D. F. & Svendsen, E. A regime shift in the North Sea circa 1988 linked to changes in the North Sea horse mackerel fishery. Fish. Res. 50, 163–171 (2001).
    Article  Google Scholar 

    46.
    Brierley, A. S. et al. Acoustic observations of jellyfish in the Namibian Benguela. Mar. Ecol. Prog. Ser. 210, 55–66 (2001).
    ADS  Article  Google Scholar 

    47.
    Brierley, A. S. et al. Towards the acoustic estimation of jellyfish abundance. Mar. Ecol. Prog. Ser. 295, 105–111 (2005).
    ADS  Article  Google Scholar 

    48.
    MacLennan, D. N. & Simmonds, E. J. Fisheries Acoustics (Springer, Berlin, 2013).
    Google Scholar 

    49.
    Planque, B. & Fromentin, J. Calanus and environment in the eastern North Atlantic. I. Spatial and temporal patterns of C. finmarchicus and C. helgolandicus. Mar. Ecol. Prog. Ser. 134, 101–109 (1996).
    ADS  Article  Google Scholar 

    50.
    Batten, S. D. et al. CPR sampling: The technical background, materials and methods, consistency and comparability. Prog. Oceanogr. 58, 193–215 (2003).
    ADS  Article  Google Scholar 

    51.
    Richardson, A. J. et al. Using continuous plankton recorder data. Prog. Oceanogr. 68, 27–74 (2006).
    ADS  Article  Google Scholar 

    52.
    John, E. H. et al. Continuous plankton records stand the test of time: Evaluation of flow rates, clogging and the continuity of the CPR time-series. J. Plankton Res. 24, 941–946 (2002).
    Article  Google Scholar 

    53.
    Beaugrand, G., Ibañez, F., Lindley, J. A. & Reid, P. C. Diversity of calanoid copepods in the North Atlantic and adjacent seas: Species associations and biogeography. Mar. Ecol. Prog. Ser. 232, 179–195 (2002).
    ADS  Article  Google Scholar 

    54.
    Yu, J. Developing Bayesian Network Inference Algorithms to Predict Causal Functional Pathways in Biological Systems (Duke University, Durham, 2005).
    Google Scholar 

    55.
    Chickering, D. M. Learning Bayesian Networks is NP-Complete. In Learning from Data: Artificial Intelligence and Statistics V (eds. Fisher, D. & Lenz, H.-J.) 121–130 (Springer, Berlin, 1996). https://doi.org/10.1007/978-1-4612-2404-4_12.

    56.
    Rayner, N. A. et al. Global analyses of sea surface temperature, sea ice, and night marine air temperature since the late nineteenth century. J. Geophys. Res. Atmos. 108, (2003).

    57.
    Jones, P. D., Jonsson, T. & Wheeler, D. Extension to the North Atlantic oscillation using early instrumental pressure observations from Gibraltar and south-west Iceland. Int. J. Climatol. 17, 1433–1450 (1997).
    Article  Google Scholar 

    58.
    Mitchell, E. G. Functional programming through deep time: Modeling the first complex ecosystems on earth. ACM SIGPLAN Not. 46, 28–31 (2011).
    MATH  Article  Google Scholar 

    59.
    Jones, S. P. Haskell 98 Language and Libraries: The Revised Report (Cambridge University Press, Cambridge, 2003).
    Google Scholar 

    60.
    Friedman, N., Geiger, D. & Goldszmidt, M. Bayesian network classifiers. Mach. Learn. 29, 131–163 (1997).
    MATH  Article  Google Scholar 

    61.
    Dickey-Collas, M., Nash, R. D. M. & Brown, J. The location of spawning of Irish sea herring (Clupea harengus). J. Mar. Biol. Assoc. 81, 713–714 (2001).
    Article  Google Scholar 

    62.
    Nash, R. D. M. & Geffen, A. J. Seasonal and interannual variation in abundance of Calanus finmarchicus (Gunnerus) and Calanus helgolandicus (Claus) in inshore waters (west coast of the Isle of Man) in the central Irish Sea. J. Plankton Res. 26, 265–273 (2004).
    Article  Google Scholar 

    63.
    Hurrell, J. W. & Deser, C. North Atlantic climate variability: The role of the North Atlantic Oscillation. J. Mar. Syst. 78, 28–41 (2009).
    Article  Google Scholar 

    64.
    Brotz, L., Cheung, W. W. L., Kleisner, K., Pakhomov, E. & Pauly, D. Increasing jellyfish populations: trends in Large Marine Ecosystems. In Jellyfish Blooms IV: Interactions with Humans and Fisheries (eds. Purcell, J. et al.) 3–20 (Springer Netherlands, 2012). https://doi.org/10.1007/978-94-007-5316-7_2.

    65.
    Purcell, J. E. Jellyfish and ctenophore blooms coincide with human proliferations and environmental perturbations. Annu. Rev. Mar. Sci. 4, 209–235 (2012).
    ADS  Article  Google Scholar 

    66.
    Pitt, K. A., Lucas, C. H., Condon, R. H., Duarte, C. M. & Stewart-Koster, B. Claims that anthropogenic stressors facilitate jellyfish blooms have been amplified beyond the available evidence: A systematic review. Front. Mar. Sci. 5, 451 (2018).
    Article  Google Scholar 

    67.
    Sanz-Martín, M. et al. Flawed citation practices facilitate the unsubstantiated perception of a global trend toward increased jellyfish blooms. Glob. Ecol. Biogeogr. 25, 1039–1049 (2016).
    Article  Google Scholar 

    68.
    Dunne, J. A., Williams, R. J. & Martinez, N. D. Network structure and biodiversity loss in food webs: Robustness increases with connectance. Ecol. Lett. 5, 558–567 (2002).
    Article  Google Scholar 

    69.
    Gardner, M. R. & Ashby, W. R. Connectance of large dynamic (cybernetic) systems: Critical values for stability. Nature 228, 784–784 (1970).
    ADS  CAS  PubMed  Article  Google Scholar 

    70.
    Lawton, J. H. & Brown, V. K. Redundancy in ecosystems. In Biodiversity and Ecosystem Function (eds. Schulze, E.-D. & Mooney, H. A.) 255–270 (Springer, Berlin, 1994). https://doi.org/10.1007/978-3-642-58001-7_12.

    71.
    Thébault, E. & Loreau, M. Trophic interactions and the relationship between species diversity and ecosystem stability. Am. Nat. 166, E95–E114 (2005).
    PubMed  Article  Google Scholar 

    72.
    Wang, S. & Loreau, M. Biodiversity and ecosystem stability across scales in metacommunities. Ecol. Lett. 19, 510–518 (2016).
    PubMed  PubMed Central  Article  Google Scholar 

    73.
    Van Voris, P., O’Neill, R. V., Emanuel, W. R. & Shugart, H. H. Functional complexity and ecosystem stability. Ecology 61, 1352–1360 (1980).
    Article  Google Scholar 

    74.
    Graham, W. M. & Kroutil, R. M. Size-based prey selectivity and dietary shifts in the jellyfish, Aurelia aurita. J. Plankton Res. 23, 67–74 (2001).
    Article  Google Scholar 

    75.
    Marques, R., Bonnet, D., Carré, C., Roques, C. & Darnaude, A. M. Trophic ecology of a blooming jellyfish (Aurelia coerulea) in a Mediterranean coastal lagoon. Limnol. Oceanogr. n/a.

    76.
    Anninsky, B. E., Finenko, G. A., Datsyk, N. A. & Kıdeyş, A. E. Trophic ecology and assessment of the predatory impact of the Moon jellyfish Aurelia aurita (Linnaeus, 1758) on zooplankton in the Black Sea (2020) https://doi.org/10.21411/cbm.a.96dd01aa.

    77.
    Widmer, C. L., Fox, C. J. & Brierley, A. S. Effects of temperature and salinity on four species of northeastern Atlantic scyphistomae (Cnidaria: Scyphozoa). Mar. Ecol. Prog. Ser. 559, 73–88 (2016).
    ADS  Article  Google Scholar 

    78.
    Watson, D. I. & Barnes, D. K. A. Temporal and spatial components of variability in benthic recruitment, a 5-year temperate example. Mar. Biol. 145, 201–214 (2004).
    Article  Google Scholar 

    79.
    Arrhenius, F. & Hansson, S. Food consumption of larval, young and adult herring and sprat in the Baltic Sea. Mar. Ecol. Prog. Ser. 96, 125–137 (1993).
    ADS  Article  Google Scholar 

    80.
    Williams, R., Conway, D. V. P. & Hunt, H. G. The role of copepods in the planktonic ecosystems of mixed and stratified waters of the European shelf seas. Hydrobiologia 292, 521–530 (1994).
    Article  Google Scholar 

    81.
    Gowen, R. J., Mills, D. K., Trimmer, M. & Nedwell, D. B. Production and its fate in two coastal regions of the Irish Sea: The influence of anthropogenic nutrients. Mar. Ecol. Prog. Ser. 208, 51–64 (2000).
    ADS  Article  Google Scholar 

    82.
    Scorrano, S., Aglieri, G., Boero, F., Dawson, M. N. & Piraino, S. Unmasking Aurelia species in the Mediterranean Sea: An integrative morphometric and molecular approach. Zool. J. Linn. Soc. 180, 243–267 (2017).
    Google Scholar 

    83.
    Haussermann, V., Dawson, M. N. & Forsterra, G. First record of the moon jellyfish, Aurelia for Chile. Spixana 32, 3–7 (2009).
    Google Scholar  More

  • in

    Rainfall as a trigger of ecological cascade effects in an Australian groundwater ecosystem

    Study area
    The field work was carried out at the Sturt Meadows calcrete aquifer (28°41′ S 120°58′ E) located on Sturt Meadows pastoral station, Western Australia, ~ 42 km from the settlement of Leonora (833 km northeast of Perth, see Fig. 1a). The study area is a calcrete aquifer lying in the Raeside paleodrainages in the Yilgarn region of Western Australia (Fig. 1a). The vegetation of the area is Acacia woodlands, primarily Acacia aneura (F.Muell. ex Benth.), and is subjected to combined grazing pressure from domestic stock, feral animals and macropods. The aquifer is accessible through a bore grid comprising 115 bore holes of between 5 and 11 m depth (Fig. 1b).
    Figure 1

    Map of the Sturt Meadows calcrete. (a) Location within the Yilgarn craton region and detailed paleodrainages/cacretes in the area and (b) the grid map showing the location of the boreholes sampled for stygofauna together with probe samples, water samples (in light blue) and the combination of both. Map was produced in ArcGIS Desktop 10.679 and edited in Adobe Illustrator 25.080.

    Full size image

    Three sampling campaigns were carried out, two of them (LR1: 26/07/2017 and LR2: 07/11/2017) corresponding to low rainfall periods24 and one during the wet season (high rainfall, HR; two consecutive days of sampling collection: 17-18/03/2018) (Supplementary Fig. 1). The well-studied stygofaunal community of the area is composed of 11 main stygofaunal taxa belonging to five Classes: Oligochaeta (family Tubificidae (Vejdovský 1884)), subcohort Hydrachnidia, Maxillopoda (two species of harpacticoids: Novanitocrella cf. aboriginesi (Karanovic, 2004), Schizopera cf. austindownsi (Karanovic, 2004) and four species of cyclopoids: Halicyclops kieferi (Karanovic, 2004), Halicyclops cf. ambiguous (Kiefer, 1967), Schizopera slenderfurca (Karanovic & Cooper, 2012) and Fierscyclops fiersi (De Laurentiis et al., 2001)), Malacostraca: Amphipoda (species Scutachiltonia axfordi (King, 2012), Yilgarniella sturtensis (King, 2012) and Stygochiltonia bradfordae (King, 2012)) and Insecta: Coleoptera: Dytiscidae (species Paroster macrosturtensis (Watts & Humphreys 2006), Paroster mesosturtensis (Watts & Humphreys 2006) and Paroster microsturtensis (Watts & Humphreys 2006) and respective larvae).
    Field work procedures and sample preparation
    Given the sensitivity of the hydrological dynamics in shallow calretes23,25, extensive water extraction along the bores was avoided, and preliminary tests on the bores with the highest water depth were carried out to quantify potential risk of dewatering the calcrete. During the field campaigns LR2 and HR, 20 water samples in total (two samples for stable isotope analysis on DOC (Dissolved Organic Carbon) and DIC (Dissolved Inorganic Carbon), three samples for radiocarbon analysis on DOC, one sample for radiocarbon analysis on DIC, and two samples for stable isotope and radiocarbon analyses on POC (Particulate Organic Carbon)) were collected from bores D13 and W4 (Fig. 1b), which are representative of the two main geological conformations of the area—calcrete (W4) and clay (D13) (Supplementary Fig. 2)—and host stable hydrological and biotic conditions7. Water samples were collected using a submersible centrifugal pump (GEOSub 12 V Purging Pump) after wells were purged of three well-volumes and stabilisation of in-field parameters was observed, according to the methodology in Bryan et al.26.
    Samples for 14CDIC analysis were filtered through 0.45 μm filters and collected in 1 L high density poly-ethylene (HDPE) bottles. δ13CDIC samples were filtered through 0.2 μm filters, collected in 12 mL glass vials (Exetainers) and refrigerated after sampling. δ 13CDOC samples were filtered through 0.2 μm filters, collected in 60 mL HDPE bottles and frozen after sampling.14CDOC samples were filtered through 0.2 μm filters, collected in 3 L HDPE bottles and frozen after sampling.
    In order to investigate 14C and δ 13C content of POC, two extra liters were collected from the same bores (D13, W4) and kept frozen (− 20 °C) until further analyses. 14CPOC δ13CPOC samples were then filtered through pre-combusted GF/F filters (12 h at 450 °C), washed with 1.2 N HCl to remove any inorganic carbon, and subsequently dried at 60 °C for 24 h. All samples were closed with sealing tape after collection to limit atmospheric exchange and kept in darkness.
    Temperature, pH, ORP, salinity, DO and depth were measured in situ using a portable Hydrolab Quanta Multi‐Probe Meter across 30 bores during LR1, LR2 and HR23 (presented in Supplementary Table 8). Adult and larval stygofaunal specimens were collected from the same bores by hauling a weighted plankton net (mesh 100 μm27) five times through the water column (Fig. 1b). All biological samples were kept frozen (− 20 °C) in darkness until laboratory processing. Individual organisms were counted and identified (and consequently separated) to the lowest taxonomic level via optical microscopy and reference to specific taxonomic keys. Plant material, sediment samples and fauna were each separated during the sorting in the laboratory and each taxon pooled according to sampling campaign (LR1, LR2 or HR) and subsequently washed with Milli-Q water to remove surface impurities from their bodies. Sediment samples were soaked in acid (0.1 N HCl) to remove inorganic carbon, and together with the other samples were then oven dried at 60 °C overnight and ground until obtaining a homogeneous fine powder and stored at − 20 °C until further analyses.
    Previous investigations on the ecological niche trends at Sturt Meadows indicated that all stygofauna characterize similar niche occupations under low rainfall regimes (LR1 and LR2)23. Stygofaunal specimens from the two low rainfall sampling events were combined to form sample LR to address the competing requirements between isotopic detection limits, analytical replicates and cost, while maintaining the main taxonomic and biological classifications7.
    Bulk isotope and 14C analyses
    Water δ13CDIC and δ13CPOC isotopic ratios were analysed by Isotope Ratio Mass Spectrometer—WABC at The University of Western Australia using a GasBench II coupled with a Delta XL Mass Spectrometer (Thermo-Fisher Scientific)—and the results, with a precision of ± 0.10‰, were reported as ‰ deviation from the NBS19 and NSB18 international carbonate standard28.
    δ13CDOC isotopic ratios of waters were analysed via Liquid Chromatography Isotope Ratio Mass Spectrometer (LC-IRMS) at the La Trobe Institute for Molecular Sciences (LIMS, La Trobe University, Melbourne, Australia) comprising an Accela 600 pump connected to a Delta V Plus Isotope Ratio Mass Spectrometer via a Thermo Scientific LC Isolink (Thermo Scientific).
    C and N bulk SIA on homogenised samples of sediment, roots, stygofauna and copepods (cyclopoids and harpacticoids) were performed at the Australian Nuclear Science and Technology Organisation (ANSTO, Sydney, Australia). Samples were loaded into tin capsules and analysed with a continuous flow isotope ratio mass spectrometer (CF-IRMS), model Delta V Plus (Thermo Scientific Corporation, U.S.A.), interfaced with an elemental analyser (Thermo Fisher Flash 2000 HT EA, Thermo Electron Corporation, USA) following the procedure published by Mazumder et al.29.
    For radiocarbon analyses, samples (sediment, roots, copepods, ants, stygofauna) were treated with 1 M HCl for 2 h to remove all possible carbonate contamination. These pre-treated samples together with 14CPOC, 14CDOC and 14CDIC samples were subjected to CO2 extraction and graphitization following the methodology published by Hua et al.30. 14C content of samples was determined by means of the Accelerator Mass Spectrometry (AMS) at ANSTO.
    Carbon CSIA
    Carbon CSIA followed the procedure described in Saccò et al.7. Samples of roots and stygofaunal specimens were hydrolysed under vacuum with 0.5 to 1 mL of amino acid-free 6 M HCl (Sigma-Aldrich) at 110 °C for 24 h. The protein hydrolysates were dried overnight in a rotary vacuum concentrator and stored in a freezer. Prior to analysis, the samples were dissolved in Milli-Q water and 10 μL of 1-mmol solution of 2-aminoisobutyric acid (Sigma-Aldrich) as internal standard. The sample stock had a concentration of approximately 8 to 10 mg/mL, which was further diluted as needed. Single amino acid carbon isotope analysis was carried out at the La Trobe Institute for Molecular Sciences (LIMS, La Trobe University, Melbourne, Australia) using an Accela 600 pump connected to a Delta V Plus Isotope Ratio Mass Spectrometer via a Thermo Scientific LC Isolink (Thermo Scientific).
    The amino acids were separated using a mixed mode (reverse phase/ion exchange) Primesep A column (2.1 × 250 mm, 100 °C, 5 μm, SIELC Technologies) following the chromatographic method described in Mora et al.31 after Smith et al.32. Mobile phases are those described in Mora et al.33. Percentage of Phases B and C in the conditioning run, as well as flow rate of the analytical run and timing of onset of 100% Phase C were adjusted as needed. Samples were injected onto the column in the 15 μL—partial loop or no waste—injection mode, and measured in duplicate or triplicate.
    To elucidate carbon flows through the stygofaunal community we focused on the essential amino acids Valine (Val), Phenylalanine (Phe) and Arginine (Arg), as these compounds must be integrated through diet and cannot be synthetised internally by the fauna14,34. In addition, to distinguish between terrestrial and aquatic carbon sources, the ratio between Val and Phe signals (δ13CVal-Phe), a widely employed index in archaeology and freshwater biology35, was calculated for roots, water mites, aquatic worms, amphipods and beetles (larvae and adults).
    Microbial taxonomic and functional gene analyses
    Consumer amphipods (Scutachiltonia axfordi (AM1), Yilgarniella sturtensis (AM2), S. bradfordae (AM3)), cyclopoids and harpacticoids, together with predator stygobiotic beetles (Paroster macrosturtensis (B), P. mesosturtensis (M) and P. microsturtensis (S)) (see Saccò el at 7. for further details on the trophic characterisation of the stygofaunal community at Sturt Meadows), were used for gut microbiome bacterial 16S metabarcoding and microbial functional analysis. A total of 16 AM1, 16 AM2, 16 AM3, 20 cyclopoids and 20 harpaticoids and 20 of each one of the three Paroster species (B, M and S), were sorted into duplicates of stygobiotic pools of 3–5 individuals from both LR and HR events for DNA extraction. Prior to DNA extraction stygobitic animals (3–5 individuals per pool; n = 40) were placed in a petri dish containing ultrapure water and UV sterilized in a UV oven for 10 min to eliminate any bacterial species that may be contained on the exoskeleton as this study targeted the gut microbiome. Immediately post-UV treatment, the animals were placed into tissue lysis tubes with 180 μL tissue lysis buffer (ATL) and 20 μL Proteinase K and homogenised using Minilys tissue homogeniser (ThermoFisher Scientific, Australia) at high speed for 30 s. Lysis tubes, inclusive of two laboratory controls, were incubated at 56 °C using an agitating heat block (Eppendorf ThermoStat C, VWR, Australia) for 5 h.
    Following the incubation, the analytical procedure was adapted from Saccò et al.22 and DNA extraction was carried out using DNeasy Blood and Tissue Kit (Qiagen; Venlo, Netherlands) and eluted off the silica column in 30–50 μL AE buffer. The quality and quantity of DNA extracted from each stygobitic pool was measured using quantitative PCR (qPCR), targeting the bacterial 16S gene. PCR reactions were used to assess the quality and quantity of the DNA target of interest via qPCR (Applied Biosystems [ABI], USA) in 25 μL reaction volumes consisting of 2 mM MgCl2 (Fisher Biotec, Australia), 1 × PCR Gold Buffer (Fisher Biotec, Australia), 0.4 μM dNTPs (Astral Scientific, Australia), 0.1 mg bovine serum albumin (Fisher Biotec, Australia), 0.4 μM of each primer (Bact16S_515F and Bact16S_806R36,37), and 0.2 μL of AmpliTaq Gold (AmpliTaq Gold, ABI, USA), and 2 μL of template DNA (Neat, 1/10, 1/100 dilutions). The cycling conditions were: initial denaturation at 95 °C for 5 min, followed by 40 cycles of 95 °C for 30 s, 52 °C for 30 s, 72 °C for 30 s, and a final extension at 72 °C for 10 min.
    DNA extracts that successfully yielded DNA of sufficient quality, free of inhibition, as determined by the initial qPCR screen (detailed above), were assigned a unique 6–8 bp multiplex identifier tag (MID-tag) with the bacterial 16S primer set. Independent MID-tag qPCR for each stygobiotic pool were carried out in 25 μL reactions containing 1 × PCR Gold Buffer, 2.5 mM MgCl2, 0.4 mg/mL BSA, 0.25 mM of each dNTP, 0.4 μM of each primer, 0.2 μL AmpliTaq Gold and 2–4 μL of DNA as determined by the initial qPCR screen. The cycling conditions for qPCR using the MID-tag primer sets were as described above. MID-tag PCR amplicons were generated in duplicate and the library was pooled in equimolar ratio post-PCR for DNA sequencing. The final library was size selected (160–600 bp) using Pippin Prep (Sage Sciences, USA) to remove any MID-tag primer-dimer products that may have formed during amplification. The final library concentration was determined using a QuBitTM 4 Fluorometer (Thermofischer, Australia) and sequenced using a 300 cycle V2 kit on an Illumina MiSeq platform (Illumina, USA).
    MID-tag bacterial 16S sequence reads obtained from the MiSeq were sorted (filtered) back to the stygobitic pool based on the MID-tags assigned to each DNA extract using Geneious v10.2.538. MID-tag and primer sequences were trimmed from the sequence reads allowing for no mismatch in length or base composition.
    Filtered reads were then input into a containerised workflow comprising USEARCH39 and BLASTN40, which was run on a high-throughput HPC cluster at Pawsey supercomputing centre. The fastx-uniques, unoise3 (with minimum abundance of 8) and otutab commands of USEARCH were applied to generate unique sequences, ZOTUs (zero-radius Operational Taxonomic Units) and abundance table, respectively. The ZOTUs were compared against the nucleotide database using the following parameters in BLASTN: perc_identity ≥ 94, evalue ≤ 1e−3, best_hit_score_edge 0.05, best_hit_overhang 0.25, qcov_hsp_perc 100, max_target_seqs = 5. An in-house Python script was used to assign the ZOTUs to their lowest common ancestor (LCA)41. The threshold for dropping a taxonomic assignment to LCA was set to perc_identity ≥ 96 and the difference between the % of identity of the two hits when their query coverage is equal was set to 1. Results on the microbial taxonomic diversity detected in ground water samples from a previous study on carbon inputs in the aquifer22 were incorporated in this work to allow qualitative comparison with the stygofaunal gut diversity.
    To investigate functional activity involved in carbon cycling, the 16S metabarcoding data were fed to the Phylogenetic Investigation of Communities by Reconstruction of Unobserved States 2 (PICRUSt2) software package to generate predicted metagenome profiles42. These profiles were clustered into Kyoto Encyclopedia of Genes and Genomes (KEGG) Orthologs (KOs)43 and MetaCyc pathway abundances44 focusing on the relative abundances of four carbon metabolisms: carbon fixation in prokaryotes, carbohydrates, lipids and amino acid metabolisms. Relative abundance of pathways linked with methane, nitrogen and sulfur metabolisms were also investigated.
    Statistical analyses
    The Phyloseq package in R45,46 was used to plot the ZOTU abundance for the styfofaunal specimens at the order level under LR and HR periods. The Statistical Analysis of Metagenomic Profiles (STAMP) bioinformatics software package was used to carry out Principal Components Analysis (PCA) to assess the differences between functional genomic fingerprints based on the KEGG orthologs function prediction between copepods (C and H) and amphipods (AM1, AM2 and AM3), and determine statistically significant results from the PICRUSt2 output47. Clustering patterns according to rainfall periods (LR and HR) and major consumers taxonomic groups (cyclopoids, harpacticoids and amphipods) were assessed through Permutational multivariate analysis of variance (PERMANOVA, R-package46 ‘vegan’) and pairwise post hoc pairwise multilevel comparisons48.
    For comparison of potential shifts in abundances of microbial metabolic pathways within groundwater samples, copepods and amphipods across rainfall periods, analysis of variance (ANOVA) was performed on the abundance data (two replicates per each group) on the predicted pathways depicting carbon fixation, carbohydrate, lipid, amino acid, methane, nitrogen and sulfur metabolisms. ANOVAs coupled with Tukey’s HSD pairwise comparisons (R-package46 ‘stats’) were employed to inspect significant differences between bulk SIA (δ13C and δ15N) and essential amino acid (δ13CPhe, δ13CArg, δ13CVal and δ13CVal-Phe) data from the stygofaunal taxa within the different rainfall conditions (LR and HR). PERMANOVAs (R-package46 ‘vegan’) were also performed to investigate the potential clustering trends within the stygofaunal taxa across rainfall periods from the combination of radiocarbon (Δ14C) and carbon SIA (δ13C) isotopic fingerprints.
    SIMM (Stable Isotope Mixing Models, R-package46 ‘simmr’) were used to estimate dietary proportions of copepods and amphipods within a Bayesian framework. Due to the lack of species-specific trophic discrimination factors for stygofauna, we employed the widely accepted values of 3.4 ± 2‰ for nitrogen and 0.5 ± 1‰ for carbon49. Markov chain Monte Carlo (MCMC) algorithms were used for simulating posterior distributions in SIMM, and MCMC convergence was evaluated using the Gelman-Rubin diagnostic by using 1.1 as a threshold value for analysis validation. More

  • in

    Seasonal biological carryover dominates northern vegetation growth

    Interseasonal VGC dominates peak-to-late-season growth
    We first examined the VGC effect at the seasonal scale, using satellite-derived Normalized Difference Vegetation Index (NDVI, see “Methods”) for the 1982–2016 period. We defined the dormancy season (DS) and three periods of the growing season, i.e., EGS, peak growing season (PGS), and late growing season (LGS), based on phenological metrics (see “Methods”). The partial autocorrelation calculated for NDVI time series of two consecutive seasons, after factoring out concurrent and preceding climatic impacts, provides an estimate of the interseasonal VGC effects (see “Methods”). At the hemispheric scale, our analyses show a significant (p  More

  • in

    An equation of state for insect swarms

    State variables
    The first step in describing the macroscopic properties of the swarm is to define a set of state variables that fully characterizes the state of the system. The equation of state then links these state variables in a functional relation. In classical thermodynamics, a complete set of state variables is given by the conjugate pairs of pressure P and volume V, temperature T and entropy S, and, if the number of particles is not fixed, chemical potential μ and number of particles N. We use an analogous set of state variables here to characterize swarms. The most straightforward state variable to define is the number of individuals N, which is given simply by the number of midges in the swarm at a given time (note that midges that are not swarming simply sit on the walls or floor of the laboratory enclosure). The volume V of the swarm can be straightforwardly defined and computed as the volume of the convex hull enclosing all the midges. Note that while N and V are not independently controllable quantities, the ratio N/V is empirically approximately constant in large swarms24, meaning that the “thermodynamic” limit (that is, (N to infty) and (V to infty) with (frac{N}{V} to rho)) is approached in our swarms25. In typical swarming events, N changes on a time scale that is very slow compared to the swarm dynamics; thus, a chemical potential is not needed to describe the instantaneous state of the swarm. Note, though, that since the number of midges varies between measurements that may be separated by many days, N remains a relevant state variable for capturing swarm-to-swarm variability.
    The remaining three state variables are somewhat more subtle, but can be defined by building on previous work. It has been explicitly shown26 that a virial relation based on the kinetic energy and an effective potential energy holds for laboratory swarms of Chironomus riparius. For particles moving in a potential, this virial relation can be used to define a pressure26. As we have shown previously, swarming midges behave as if they are trapped in a harmonic potential well that binds them to the swarm, with a spring constant k(N) that depends on the swarm size24,26 (Fig. 1b). The difference between the kinetic energy and this harmonic potential energy thus allows us to compute a pressure4,26,27, which is conceptually similar to the swim pressure defined in other active systems28. The virial theorem thus provides a link between kinetic energy, potential energy, and a field that plays the role of a pressure, when coupled with the observation that individual midges to a good approximation behave as if they are moving in a harmonic potential24,26. We can write this virial pressure P (per unit mass, assuming a constant mass per midge) as

    $$P = frac{1}{3NV}mathop sum limits_{i = 1}^{N} left( {v_{i}^{2} – frac{1}{2}langle krangle r_{i}^{2} } right),$$

    where N is the number of midges in the swarm, V is the swarm volume, vi is the velocity of midge i, ri is its distance from the swarm centre of mass, and (langle krangle = langle – {mathbf{a}}_{i} cdot hat{mathbf{r}}_{i} / r_{i} rangle) is the effective spring constant of the emergent potential well that binds midges to the swarm. In this expression, ai is the acceleration of midge i, (hat{mathbf{r}}_{i}) is the unit vector pointing from a midge towards the instantaneous centre of mass of the swarm (defined as (1/Nsumnolimits_{i = 1}^{N} {{mathbf{r}}_i })) and averages are taken over the individuals in the swarm. This spring constant depends on the swarm size N (Fig. 1b). We note that we have previously simply used the directly computed potential energy (- langle {mathbf{a}}_{i} cdot {mathbf{r}}_{i} rangle) to define the pressure4,27; here, we instead average the potential terms and fit them to a power law in N (Fig. 1b) to mitigate the contribution of spurious instantaneous noise in the individual positions that would be enhanced by differentiating them twice to compute accelerations. We use this power law to determine the spring constant k instantaneously at each time step.
    The results from the two methods for computing the pressure are similar and consistent, but the method we use here is less prone to noise. Physically, this pressure P can be interpreted as the additional spatially variable energy density required to keep the midges bound to the swarm given that their potential energy varies in space but their mean velocity (and therefore kinetic energy) does not. Thus, compared to a simple passive particle moving in a harmonic well, midges have more kinetic energy than expected at the swarm edges; this pressure compensates for the excess kinetic energy. This pressure should be viewed as a manifestation of the active nature of the midges (similar to a swim pressure28), since the kinetic energy is an active property of each individual midge and the potential energy is an emergent property of the swarm.
    We can define a Shannon-like entropy S via its definition in terms of the joint probability distributions of position and velocity. This entropy is defined as

    $$S = – mathop smallint limits_{ – infty }^{infty } p(x,;v)log_{2} p(x,;v)dxdv,$$

    where p(x,v) is the joint probability density function (PDF) of midge position and velocity. S here is measured in bits, as it is naturally an information entropy. Empirically, we find that the position and velocity PDFs are nearly statistically independent for all components and close to Gaussian, aside from the vertical component of the position (Fig. 1c–f). However, the deviation from Gaussianity in this component (which occurs because of the symmetry breaking due to the ground) does not significantly affect the estimate of the entropy; thus, we approximate it as Gaussian as well. Making these approximations, we can thus analytically write the (extensive) entropy as

    $$S = frac{3N}{{ln 2}}ln left( {2pi esigma_{x} sigma_{v} } right),$$

    where (sigma_{x}) and (sigma_{v}) are the standard deviations of the midge positions and velocities, respectively. In practice, we calculated (sigma_{v}) by averaging the instantaneous root-mean-square values of all three velocity components rather than a time-averaged value; the difference between these components was always less than 10%. This expression makes it more clear why the Gaussian approximation for the vertical component of the position is reasonable here: only the mean and variance of the PDFs are required to compute the entropy, and these low moments are very similar for the true data and the Gaussian estimate.
    Although there is no obvious definition of temperature for a swarm, we can define one starting from the entropy, since temperature (when scaled by a Boltzmann constant) can be defined as the increase in the total physical energy of the system due to the addition of a single bit of entropy. Given our definitions, adding a single bit of entropy (that is, setting (S to S + 1)) for constant (sigma_{x}) and N (that is, a swarm of fixed number and spatial size) is equivalent to setting (sigma_{v} to 2^{1/(3N)} sigma_{v} .) Adding this entropy changes the total energy of the system by an amount

    $$frac{3}{2}sigma_{v}^{2} Nleft( {2^{frac{2}{3N}} – 1} right) equiv k_{B}^{*} T,$$

    which we thus define as the temperature (k_{B}^{*} T). Even though this temperature is nominally a function of the swarm size N, it correctly yields an intensive temperature as expected in the limit of large N, as the explicit N-dependence vanishes in that limit since (lim_{n to infty } k_{B}^{*} T = sigma_{v}^{2} ln 2). In practice, this limit is achieved very rapidly: we find that this temperature is nearly independent of N for N larger than about 20, consistent with our earlier results on the effective “thermodynamic limit” for swarms25. The effective Boltzmann constant (k_{B}^{*}) is included here to convert between temperature and energy, though we note that we cannot set its value, as there is no intrinsically preferred temperature scale.
    Equipartition
    With these definitions in hand, we can evaluate the suitability of these quantities for describing the macroscopic state of midge swarms. First, we note that proper state variables ought to be independent of the swarm history; that is, they ought to describe only the current state of the system rather than the protocol by which that state was prepared. Although this property is difficult to prove incontrovertibly, none of the definitions of our state variables have history dependence. We further find that when these state variables are modulated (see below), their correlation times are very short, lending support to their interpretation as true state variables. We can also compare the relationships between these state variables and the swarm behaviour to what would be expected classically. In equilibrium thermodynamics, for example, temperature is connected to the number of degrees of freedom (d.o.f.) in a system via equipartition, such that each d.o.f. contributes an energy of (frac{1}{2}k_{B}^{*} T). We can write the total energy E of a swarm as the sum of the kinetic energy (E_{k} (t) = frac{1}{2}v^{2}) and potential energy (E_{p} (t) = frac{1}{2}k(N)r(t)^{2}) for all the individuals, where r is the distance of a midge to the swarm centre of mass, v is the velocity of a midge, and k(N) is the effective spring constant. Surprisingly, even though individual midges are certainly not in equilibrium due to their active nature, we find that the total energy is linear in both T and N (Fig. 2a), and that there is no apparent anisotropy, suggesting that equipartition holds for our swarms. This result is highly nontrivial, especially given that our definition of T does not contain the spring constant k(N), which is only determined empirically from our data. Moreover, the slope of the (E/k_{B}^{*} T) curve is well approximated as (9/2)N, implying that each midge has 9 effective d.o.f. (or 6 after discounting the factor of ({text{ln}}2) in our definition of (k_{B}^{*} T)) These d.o.f. can be identified as 3 translational and 3 potential modes, given that the potential well in which the midges reside is three-dimensional. These results demonstrate the surprising applicability of equilibrium thermodynamics for describing the macroscopic state of swarms29.
    Figure 2

    Equipartition and the equation of state. (a) The total energy of the system (E) normalized by (k_{B}^{*} T) as a function of swarm size (blue) along with the kinetic energy (E_{k}) (yellow) and potential energy (E_{p}) (blue). The total normalized energy of the system is well approximated by (9/2)N (black dashed line), indicating that each individual midge contributes ((9/2)k_{B}^{*} T) to E and thus has 9 degrees of freedom (6 after discounting the factor of ({text{ln}}2) in our definition of (k_{B}^{*} T)). The deviations from that behaviour for the largest swarms can be attributed to a growing uncertainty in the energy due to the smaller number of experiments with such large swarms. (b) A portion of our ensemble of data of the measured pressure (blue). The yellow line is the reconstruction of the pressure from our equation of state. The inset shows a zoomed-in portion of the data to highlight the quality of the reconstruction. (c) PDF of the pressure for our entire data ensemble23. The statistics of the directly measured pressure (blue) and reconstructed pressure from the equation of state have nearly identical statistics for the full dynamic range of the signal.

    Full size image

    Equation of state
    The fundamental relation in any thermodynamic system is the equation of state that expresses how the state variables co-vary. Equations of state are thus the foundation for the design and control of thermodynamic systems, because they describe how the system will respond when a subset of the state variables are modulated. Any equation of state can be written in the form (P = f(V, ;T,;N)) for some function f. Although the form of f is a priori unknown, it can typically be written as a power series in V, T, and N, in the spirit of a virial expansion. We fit the equation of state to our data assuming the functional form

    $$P = f(V, ;k_{B}^{*} T,;N) = c_{4} V^{{c_{1} }} (k_{B}^{*} T)^{{c_{2} }} N^{{c_{3} }},$$

    and using nonlinear least-squares regression. We chose to fit to the pressure for convenient analogy with a thermodynamic framework, but any other variable would have been an equivalent possibility. We note that when fitting, we normalized all the state variables by their root-mean-square values so that they were all of the same order of magnitude. These normalization pre-factors do not change the exponents, but are instead simply absorbed into (c_{4}). Thus, to leading order, we assume (P = f(V,;k_{B}^{*} T,;N) propto V^{{c_{1} }} (k_{B}^{*} T)^{{c_{2} }} N^{{c_{3} }}) and fit this relation to the swarm pressure (Fig. 2b,c), obtaining c1 = − 1.7, c2 = 2, and c3 = 1, with uncertainties on the order of 1%. Although the expression for the pressure does depend on three parameters in a nonlinear fashion, the resulting estimates for these parameters are remarkably stable and consistent across all measurements. Hence, we arrive at the equation of state (PV^{1.7} propto N(k_{B}^{*} T)^{2} .)
    This equation of state reveals aspects of the nature of swarms, particularly when compared with the linear equation of state for an ideal gas (where (PV = Nk_{B} T)). In both cases, for example, to maintain a fixed pressure and volume, smaller systems need to be hotter; but this requirement is less severe for swarms since the temperature is squared, meaning that midges have to speed up less than ideal gas molecules do. Likewise, to maintain a fixed temperature, volume expansion must be counteracted by a reduction in pressure; but midges must lower the pressure more than a corresponding ideal gas, which is reflective of the decrease of the swarm spring constant with size.
    Thermodynamic cycling
    Beyond such reasoning, however, the true power of an equation of state in thermodynamics lies in specifying how the state variables will change when some are varied but the system remains in the same state, such as in an engine. To demonstrate that our equation of state similarly describes swarms, it is thus necessary to drive them away from their natural state. Although it is impossible to manipulate the state variables directly in this system of living organisms as one would do with a mechanical system, we have shown previously that time-varying acoustic30 and illumination27 stimulation lead to macroscopic changes in swarm behaviour. Here we therefore build on these findings and use interlaced illumination changes and acoustic signals to drive swarms along four distinct paths in pressure–volume space, analogous to a thermodynamic engine cycle. The stimulation protocol is sketched in Fig. 3a. The “on” state of the acoustic signal is telegraph noise (see Experimental details), while the “off” state is completely quiet. The illumination signal simply switches between two different steady light levels. Switching between the four states of “light-high and sound-on,” “light-high and sound-off,” “light-low and sound-off,” and “light-low and sound-on” with a 40-s period (Fig. 3a) produces the pressure–volume cycle shown in Fig. 3b. We suspect that the loops in the cycle stem from the swarm’s typical “startle” response after abrupt changes in environmental conditions, followed by a rapid relaxation to a steady state27,30.
    Figure 3

    Thermodynamic cycling of a midge swarm with (langle Nrangle = 27). Schematic of the perturbation cycle showing the illumination (solid) and sound (dashed) signal timings. The symbols indicate the switching points identified in (b). (b,c) Phase-averaged swarm behaviour during the perturbation cycle plotted in the pressure–volume phase plane for (b) the pressure signal as measured and (c) as reconstructed using our equation of state. (leftlangle {,} rightrangle_{phi }) denotes a phase average of a quantity over a full cycle. The four different states of the perturbation signal are indicated. The data has been averaged using a moving 3.5-s window for clarity. The swarm behaviour moves in a closed loop in this phase plane during this cycling, as would be expected for an engine in equilibrium thermodynamics, and the equation of state holds throughout even though it was developed only for unperturbed swarms. (d) Phase-averaged pressure (langle Prangle_{phi }) of the swarm during a continuous cycle through the four light and sound states. The blue line shows the directly measured pressure and the yellow line shows the reconstruction using the equation of state.

    Full size image

    In addition to the pressure and volume, we can also measure the other state variables as we perturb the swarms. Given that we do not observe any evidence of a phase transition, we would expect that our equation of state, if valid, should hold throughout this cycle. To check this hypothesis, we used the measured V, T, and N values during unperturbed experiments along with the equation of state to predict the scaling exponents, and in turn the pressure P. We then use these baseline, unperturbed exponents and V, T, and N during the interlaced perturbations to predict a pressure P. This pressure prediction matches the measured signal exceptionally well (Fig. 3c,d) even though the equation of state was formulated only using data from unperturbed swarms, highlighting the quality of this thermodynamic analogy. Although we might expect that a strong enough perturbation might lead to qualitatively different behaviour (if the swarm went through the analog of a phase transition31), our results give strong support to the hypothesis that our equation of state should hold for any perturbation that does not drive such a transition. More

  • in

    Humans push a hulking fish with a chainsaw nose towards oblivion

    The knifetooth sawfish (Anoxypristis cuspidata) is one of five sawfish species, all of which are facing extinction. Credit: Norbert Wu/Minden/Nature Picture Library

    Conservation biology
    12 February 2021

    The strange-looking sawfish, itself a predator, falls prey to overfishing and habitat destruction.

    The long-snouted, shark-like predators called sawfish have vanished from nearly 60% of their historical habitat and are nearing extinction, according to a mathematical method of counting marine species.
    Sawfish (family Pristidae) are rays that live near seashores and in mangrove areas: habitats that are rapidly disappearing. To study what this means for the five species in the family, Helen Yan at Simon Fraser University in Burnaby, Canada, and her colleagues reviewed sawfish research conducted between 2014 and 2019 in 64 countries. From this, they estimated the distribution patterns of these fishes, which are known for their large size — some species routinely reach 5 metres in length — and their chainsaw-like noses.
    The team found the creatures’ distribution patterns had changed in places affected by fishing pressures and habitat loss. These pressures have driven the fishes to extinction in 55 of the 90 countries whose waters they once occupied.
    Saving the sawfish, the researchers say, will require countries to protect the creatures’ offshore habitats and to ban fishers from keeping sawfish that are caught in their nets. More

  • in

    Photosynthetic parameters of a sedge-grass marsh as a big-leaf: effect of plant species composition

    1.
    IPCC. Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. (Cambridge University Press, Cambridge, 2013).
    2.
    Reichstein, M. et al. Climate extremes and the carbon cycle. Nature 500, 287–295 (2013).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    3.
    Matson, P. A., Harriss, a. R. C. (ed.) Biogenic Trace Gases: Measuring Emissions from Soil and Water. Methods in ecology (Blackwell Science, Oxford [England]; Cambridge, Mass., USA, 1995).

    4.
    Baldocchi, D. et al. FLUXNET: a new tool to study the temporal and spatial variability of ecosystem-scale carbon dioxide, water vapor, and energy flux densities. Bull. Am. Meteorl. Soc. 82, 2415–2434 (2001).
    ADS  Article  Google Scholar 

    5.
    Sellers, P. J., Berry, J. A., Collatz, G. J., Field, C. B. & Hall, F. G. Canopy reflectance, photosynthesis, and transpiration. III. A reanalysis using improved leaf models and a new canopy integration scheme. Remote Sens. Environ. 42, 187–216 (1992).
    ADS  Article  Google Scholar 

    6.
    de Pury, D. & Farquhar, G. D. Simple scaling of photosynthesis from leaves to canopies without the errors of big-leaf models. Plant Cell Environ. 20, 537–557 (1997).
    Article  Google Scholar 

    7.
    Lasslop, G. et al. Separation of net ecosystem exchange into assimilation and respiration using a light response curve approach: critical issues and global evaluation. Glob. Change Biol. 16, 187–208 (2010).
    ADS  Article  Google Scholar 

    8.
    Naeem, S. & Li, S. Biodiversity enhances ecosystem reliability. Nature 390, 507–509 (1997).
    ADS  CAS  Article  Google Scholar 

    9.
    Tilman, D. Biodiversity: population versus ecosystem stability. Ecology 77, 350–363 (1995).
    Article  Google Scholar 

    10.
    Walker, B., Kinzig, A. & Langridge, J. Plant attribute diversity, resilience, and ecosystem function: the nature and significance of dominant and minor species. Ecosystems 1999, 95–113 (1999).
    Article  Google Scholar 

    11.
    Isbell, F. I., Polley, H. W. & Wilsey, B. J. Biodiversity, productivity and the temporal stability of productivity: patterns and processes. Ecol. Lett. 12, 443–451 (2009).
    PubMed  Article  PubMed Central  Google Scholar 

    12.
    Tilman, D. Functional diversity. Ecyclopedia Biodivers. 2001, 109–121 (2001).
    Article  Google Scholar 

    13.
    Larcher, W. Physiological Plant Ecology: Ecophysiology and Stress Physiology of Functional Groups (Springer, Berlin, 2003).
    Google Scholar 

    14.
    Crawford, R. M. M. Studies in Plant Survival: Ecological Case Histories of Plant Adaptation to Adversity (Blackwell Scientific Publications, Oxford, 1989).
    Google Scholar 

    15.
    Kaplan, Z. (ed.) Klíč ke květeně: České republiky. Second editions. (Academia, Prague, 2019).

    16.
    Jeník, J., Kurka, R. & Husák, Š. Wetlands of the Třeboň Basin Biosphere Reserve in the central European context, in Freshwater Wetlands and their Sustainable Future. A Case Study of the Třeboň Basin Biosphere Reserve 11–18 (CRC Press, New York, 2002).

    17.
    Holubičková, B. Příspěvek ke studiu rašeliništní vegetace. I. Mokré louky u Třeboně (A contribution to the study of peatland vegetation. I. Mokré louky near Třeboň). (1959).

    18.
    Blažková, D. Pflanzensoziologische Studie über die Wiesen der Südböhmischen Becken. Stud. CSAV 73, 1–172 (1973).
    Google Scholar 

    19.
    Prach, K. Vegetational changes in a wet meadow complex, south-bohemia, Czech Republic. Folia Geobot. Phytotaxon. 28, 1–13 (1993).
    ADS  Article  Google Scholar 

    20.
    Prach, K. Vegetation changes in a wet meadow complex during the past half-century. Folia Geobot. 43, 119–130 (2008).
    Article  Google Scholar 

    21.
    Prach, K. & Soukupová, L. Alterations in the Wet Meadows vegetation pattern, in Freshwater Wetlands and their Sustainable Future. A Case Study of the Třeboň Basin Biosphere Reserve 243–254 (CRC Press, 2002).

    22.
    Balátová-Tuláčková, E. Die Nass- und Feuchtwiesen Nordwest-Böhmens mit besonderer Berücksichtigung Der Magnocaricetalia-Gesellschaften. Rozpr. Českoslov. Akad. Věd, Řada Mat. Přír. Věd 1978.

    23.
    Květ, J. (ed.) Freshwater Wetlands and their Sustainable Future: A Case of the Třeboň Basin Biosphere Reserve, Czech Republic. Man and the biosphere series 28. (UNESCO, Paris, 2002).

    24.
    Honissová, M. et al. Seasonal dynamics of biomass partitioning in a tall sedge Carex acuta L. Aquat. Bot. 125, 64–71 (2015).
    Article  Google Scholar 

    25.
    Hejný, S. Dynamic changes in the macrophyte vegetation of South Bohemian fishponds after 35 years. Folia Geobot. Phytotaxon. 25, 245–255 (1990).
    Article  Google Scholar 

    26.
    Káplová, M., Edwards, K. R. & Květ, J. The effect of nutrient level on plant structure and production in a wet grassland: a field study. Plant Ecol. 212, 809–819 (2011).
    Article  Google Scholar 

    27.
    Chambers, J. M. Software for Data Analysis: Programming with R (Springer, Berlin, 2008).
    Google Scholar 

    28.
    Koyama, K. & Takemoto, S. Morning reduction of photosynthetic capacity before midday depression. Sci. Rep. 4, 4389 (2015).
    Article  CAS  Google Scholar 

    29.
    Nobel, P. S. Physicochemical and Environmental Plant Physiology (Academic Press, New York, 2009).
    Google Scholar 

    30.
    Gilmanov, T. G. et al. Gross primary production and light response parameters of four Southern Plains ecosystems estimated using long-term CO2-flux tower measurements: GPP OF SOUTHERN PLAINS ECOSYSTEMS. Glob. Biogeochem. Cycles 17, 1–15 (2003).
    Google Scholar 

    31.
    Bates, D. M. & Watts, D. G. Nonlinear Regression Analysis and Its Applications (Wiley, New York, 1988).
    Google Scholar 

    32.
    Ogren, E. Convexity of the photosynthetic light-response curve in relation to intensity and direction of light during growth. Plant Physiol. 101, 7 (1993).
    Article  Google Scholar 

    33.
    Hollander, M., Wolfe, D. A. & Chicken, E. Nonparametric Statistical Methods (Wiley, New York, 2014).
    Google Scholar 

    34.
    Best, D. J. & Roberts, D. E. Algorithm AS 89: the upper tail probabilities of Spearman’s Rho. Appl. Stat. 24, 377 (1975).
    Article  Google Scholar 

    35.
    Busch, J. & Losch, R. The gas exchange of Carex species from eutrophic wetlands and its dependence on microclimatic and soil wetness conditions. Phys. Chem. Earth Part B Hydrol. Oceans Atmos. 24, 117–120 (1999).
    ADS  Article  Google Scholar 

    36.
    Ondok, J. & Gloser, J. Leaf photosynthesis and dark respiration in a sedge-grass marsh. 1. model for mid-summer conditions. Photosynthetica 17, 77–88 (1983).
    Google Scholar 

    37.
    Caudle, K. L. & Maricle, B. R. Physiological relationship between oil tolerance and flooding tolerance in marsh plants. Environ. Exp. Bot. 107, 7–14 (2014).
    CAS  Article  Google Scholar 

    38.
    Ge, Z.-M. et al. Measured and modeled biomass growth in relation to photosynthesis acclimation of a bioenergy crop (Reed canary grass) under elevated temperature, CO2 enrichment and different water regimes. Biomass Bioenerg. 46, 251–262 (2012).
    CAS  Article  Google Scholar 

    39.
    Waring, E. F. & Maricle, B. R. Photosynthetic variation and carbon isotope discrimination in invasive wetland grasses in response to flooding. Environ. Exp. Bot. 77, 77–86 (2012).
    CAS  Article  Google Scholar 

    40.
    Gloser, J. Net photosynthesis and dark respiration of reed estimated by gas-exchange measurements, in Pond Littoral Ecosystems. Structure and Functioning, vol. 1978, 227–234 (Springer, Berlin, 1978).

    41.
    Zhou, X., Liu, X., Wallace, L. L. & Luo, Y. Photosynthetic and respiratory acclimation to experimental warming for four species in a tallgrass prairie ecosystem. J. Integr. Plant Biol. 49, 270–281 (2007).
    CAS  Article  Google Scholar 

    42.
    Jones, H. G. Plants and Microclimate: A Quantitative Approach to Environmental Plant Physiology (Cambridge University Press, Cambridge, 1992).
    Google Scholar 

    43.
    Smith, M. & Houpis, J. L. J. Gas exchange responses of the wetland plant Schoenoplectus hallii to irradiance and vapor pressure deficit. Aquat. Bot. 79, 267–275 (2004).
    CAS  Article  Google Scholar 

    44.
    Li, M., Yang, D. & Li, W. Leaf gas exchange characteristics and chlorophyll fluorescence of three wetland plants in response to long-term soil flooding. Photosynthetica 45, 222–228 (2007).
    Google Scholar 

    45.
    Li, M., Hou, G., Yang, D., Deng, G. & Li, W. Photosynthetic traits of Carex cinerascens in flooded and nonflooded conditions. Photosynthetica 48, 370–376 (2010).
    CAS  Article  Google Scholar 

    46.
    Vervuren, P., Beurskens, S. & Blom, C. Light acclimation, CO2 response and long-term capacity of underwater photosynthesis in three terrestrial plant species. Plant Cell Environ. 22, 959–968 (1999).
    Article  Google Scholar 

    47.
    Bouma, T. J., De Visser, R., Van Leeuwen, P. H., De Kock, M. J. & Lambers, H. The respiratory energy requirements involved in nocturnal carbohydrate export from starch-storing mature source leaves and their contribution to leaf dark respiration. J. Exp. Bot. 46, 1185–1194 (1995).
    CAS  Article  Google Scholar 

    48.
    McCutchan, C. L. & Monson, R. K. Night-time respiration rate and leaf carbohydrate concentrations are not coupled in two alpine perennial species. New Phytol. 149, 419–430 (2002).
    Article  Google Scholar 

    49.
    Emerson, R. The quantum yield of photosynthesis. Annu. Rev. Plant Physiol. 9, 1–24 (1958).
    CAS  Article  Google Scholar 

    50.
    Singsaas, E. L., Ort, D. R. & DeLucia, E. H. Variation in measured values of photosynthetic quantum yield in ecophysiological studies. Oecologia 128, 15–23 (2001).
    ADS  PubMed  Article  PubMed Central  Google Scholar 

    51.
    de Lobo, F. A. et al. Fitting net photosynthetic light-response curves with Microsoft Excel—a critical look at the models. Photosynthetica 51, 445–456 (2013).
    CAS  Article  Google Scholar 

    52.
    Busch, J. Characteristic values of key ecophysiological parameters in the genus Carex. FLORA 196, 405–430 (2001).
    Article  Google Scholar 

    53.
    Hull, J. C. Photosynthetic induction dynamics to sunflecks of four deciduous forest understory herbs with different phenologies1. Int. J. Plant Sci. 163, 913–924 (2002).
    ADS  Article  Google Scholar 

    54.
    Wayne, E. R. & Van Auken, O. W. Light responses of Carex planostachys from various microsites in a Juniperus community. J. Arid Environ. 73, 435–443 (2009).
    ADS  Article  Google Scholar 

    55.
    Colmer, T. D. & Pedersen, O. Underwater photosynthesis and respiration in leaves of submerged wetland plants: gas films improve CO2 and O2 exchange. New Phytol. 177, 918–926 (2008).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    56.
    Mommer, L. & Visser, E. J. W. Underwater photosynthesis in flooded terrestrial plants: a matter of leaf plasticity. Ann. Bot. 96, 581–589 (2005).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    57.
    Dušek, J. Seasonal dynamic of nonstructural saccharides in a rhizomatous grass Calamagrostis epigeios. Biol. Plant. 45, 383–387 (2002).
    Article  Google Scholar 

    58.
    Mitsch, W. J. (ed.) Wetland Ecosystems (Wiley, Hoboken, NJ, 2009).

    59.
    Soukupová, L. Life strategy of graminoid populations in the wet meadows, in Freshwater Wetlands and their Sustainable Future. A Case Study of the Třeboň Basin Biosphere Reserve 255–267 (CRC Press, New York, 2002).

    60.
    Polechová, J. & Storch, D. Ecological Niche, in Encyclopedia of Ecology 72–80 (Elsevier, Amsterdam, 2019). https://doi.org/10.1016/B978-0-12-409548-9.11113-3.

    61.
    Dykyjová, D. Production ecology of Acorus calamus. Folia Geobot. Phytotaxon. 15, 29–57 (1980).
    Article  Google Scholar 

    62.
    Westlake, D. F., Květ, J., Andrzej Szczepański, a International Biological Programme. (ed.) The Production Ecology of Wetlands: The IBP Synthesis (Cambridge University Press, Cambridge, UK; New York, NY, USA, 1998).

    63.
    Pai, A. & McCarthy, B. C. Variation in shoot density and rhizome biomass of Acorus calamus L. With respect to environment. Castanea 70, 263–275 (2005).
    Article  Google Scholar 

    64.
    Pai, A. & McCarthy, B. C. Suitability of the medicinal plant, Acorus calamus L, for wetland restoration. Nat. Areas J. 30, 380–386 (2010).
    Article  Google Scholar 

    65.
    Květ, J., Lukavská, J. & Tetter, M. Biomass and net primary production in graminoid vegetation. in Freshwater Wetlands and their Sustainable Future. A Case Study of the Třeboň Basin Biosphere Reserve 293–299 (CRC Press, Boca Raton, 2002).

    66.
    Hejný, S. The dynamic characteristics of littoral vegetation with respect to changes of water level. Hidrobiol. Bucur. 1971, 71–85 (1971).
    Google Scholar  More

  • in

    Reduced nest development of reared Bombus terrestris within apiary dense human-modified landscapes

    1.
    Ollerton, J., Winfree, R. & Tarrant, S. How many flowering plants are pollinated by animals?. Oikos 120, 321–326. https://doi.org/10.1111/j.1600-0706.2010.18644.x (2011).
    Article  Google Scholar 
    2.
    Klein, A. M. et al. Importance of pollinators in changing landscapes for world crops. Proc. R. Soc. B Biol. Sci. 274, 303–313. https://doi.org/10.1098/rspb.2006.3721 (2007).
    Article  Google Scholar 

    3.
    Kremen, C., Williams, N. M. & Thorp, R. W. Crop pollination from native bees at risk from agricultural intensification. Proc. Natl. Acad. Sci. U.S.A. 99, 16812–16816. https://doi.org/10.1073/pnas.262413599 (2002).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    4.
    Potts, S. G. et al. Global pollinator declines: Trends, impacts and drivers. Trends Ecol. Evol. 25, 345–353. https://doi.org/10.1016/j.tree.2010.01.007 (2010).
    Article  PubMed  Google Scholar 

    5.
    Tscharntke, T. et al. Landscape moderation of biodiversity patterns and processes—Eight hypotheses. Biol. Rev. 87, 661–685. https://doi.org/10.1111/j.1469-185X.2011.00216.x (2012).
    Article  PubMed  Google Scholar 

    6.
    Winfree, R., Aguilar, R., Vazquez, D. P., LeBuhn, G. & Aizen, M. A. A meta-analysis of bees’ responses to anthropogenic disturbance. Ecology 90, 2068–2076. https://doi.org/10.1890/08-1245.1 (2009).
    Article  PubMed  Google Scholar 

    7.
    Isaacs, R. et al. Integrated crop pollination: Combining strategies to ensure stable and sustainable yields of pollination-dependent crops. Basic Appl. Ecol. 22, 44–60. https://doi.org/10.1016/j.baae.2017.07.003 (2017).
    Article  Google Scholar 

    8.
    Steffan-Dewenter, I. & Tscharntke, T. Resource overlap and possible competition between honey bees and wild bees in central Europe. Oecologia 122, 288–296. https://doi.org/10.1007/s004420050034 (2000).
    ADS  CAS  Article  Google Scholar 

    9.
    Paini, D. R. & Roberts, J. D. Commercial honey bees (Apis mellifera) reduce the fecundity of an Australian native bee (Hylaeus alcyoneus). Biol. Cons. 123, 103–112. https://doi.org/10.1016/j.biocon.2004.11.001 (2005).
    Article  Google Scholar 

    10.
    Schaffer, W. M. et al. Competition for nectar between introduced honey bees and native North American bees and ants. Ecology 64, 564–577. https://doi.org/10.2307/1939976 (1983).
    Article  Google Scholar 

    11.
    Dupont, Y. L., Hansen, D. M., Valido, A. & Olesen, J. M. Impact of introduced honey bees on native pollination interactions of the endemic Echium wildpretii (Boraginaceae) on Tenerife, Canary Islands. Biol. Cons. 118, 301–311. https://doi.org/10.1016/j.biocon.2003.09.010 (2004).
    Article  Google Scholar 

    12.
    Garibaldi, L. A. et al. Wild pollinators enhance fruit set of crops regardless of honey bee abundance. Science 339, 1608–1611. https://doi.org/10.1126/science.1230200 (2013).
    ADS  CAS  Article  PubMed  Google Scholar 

    13.
    Thomson, D. M. Local bumble bee decline linked to recovery of honey bees, drought effects on floral resources. Ecol. Lett. 19, 1247–1255. https://doi.org/10.1111/ele.12659 (2016).
    Article  PubMed  Google Scholar 

    14.
    Thomson, D. Competitive interactions between the invasive European honey bee and native bumble bees. Ecology 85, 458–470. https://doi.org/10.1890/02-0626 (2004).
    Article  Google Scholar 

    15.
    Goulson, D. & Sparrow, K. Evidence for competition between honeybees and bumblebees; effects on bumblebee worker size. J. Insect. Conserv. 13, 177–181. https://doi.org/10.1007/s10841-008-9140-y (2009).
    Article  Google Scholar 

    16.
    Paini, D. R. Impact of the introduced honey bee (Apis mellifera) (Hymenoptera : Apidae) on native bees: A review. Austral. Ecol. 29, 399–407. https://doi.org/10.1111/j.1442-9993.2004.01376.x (2004).
    Article  Google Scholar 

    17.
    Gross, C. L. The effect of introduced honeybees on native bee visitation and fruit-set in Dillwynia juniperina (Fabaceae) in a fragmented ecosystem. Biol. Cons. 102, 89–95. https://doi.org/10.1016/s0006-3207(01)00088-x (2001).
    Article  Google Scholar 

    18.
    Nielsen, A., Reitan, T., Rinvoll, A. W. & Brysting, A. K. Effects of competition and climate on a crop pollinator community. Agric. Ecosyst. Environ. 246, 253–260. https://doi.org/10.1016/j.agee.2017.06.006 (2017).
    Article  Google Scholar 

    19.
    Lindström, S. A. M., Herbertssön, L., Rundlof, M., Bommarco, R. & Smith, H. G. Experimental evidence that honeybees depress wild insect densities in a flowering crop. Proc. R. Soc. B Biol. Sci. 283, 8. https://doi.org/10.1098/rspb.2016.1641 (2016).
    Article  Google Scholar 

    20.
    Magrach, A., González-Varo, J. P., Boiffier, M., Vilà, M. & Bartomeus, I. Honeybee spillover reshuffles pollinator diets and affects plant reproductive success. Nat. Ecol. Evol. https://doi.org/10.1038/s41559-017-0249-9 (2017).
    Article  PubMed  Google Scholar 

    21.
    González-Varo, J. P. & Vilà, M. Spillover of managed honeybees from mass-flowering crops into natural habitats. Biol. Conserv. 212, 376–382. https://doi.org/10.1016/j.biocon.2017.06.018 (2017).
    Article  Google Scholar 

    22.
    Begon, M., Harper, J. L. & Townsend, C. R. Ecology: Individuals, Populations, and Communities 3rd edn. (Blackwell Science Ltd, Hoboken, 1996).
    Google Scholar 

    23.
    United Nations. (United Nations, Department of Economic and Social Affairs, Population Division, New York, 2012).

    24.
    Goulson, D., Nicholls, E., Botias, C. & Rotheray, E. L. Bee declines driven by combined stress from parasites, pesticides, and lack of flowers. Science 347, 1435. https://doi.org/10.1126/science.1255957 (2015).
    CAS  Article  Google Scholar 

    25.
    Scheper, J. et al. Local and landscape-level floral resources explain effects of wildflower strips on wild bees across four European countries. J. Appl. Ecol. 52, 1165–1175. https://doi.org/10.1111/1365-2664.12479 (2015).
    Article  Google Scholar 

    26.
    McCune, F., Normandin, E., Mazerolle, M. J. & Fournier, V. Response of wild bee communities to beekeeping, urbanization, and flower availability. Urban Ecosyst. https://doi.org/10.1007/s11252-019-00909-y (2019).
    Article  Google Scholar 

    27.
    Samuelson, A. E., Gill, R. J., Brown, M. J. F. & Leadbeater, E. Lower bumblebee colony reproductive success in agricultural compared with urban environments. Proc. R. Soc. B Biol. Sci. 285, 9. https://doi.org/10.1098/rspb.2018.0807 (2018).
    Article  Google Scholar 

    28.
    Steffan-Dewenter, I. & Kuhn, A. Honeybee foraging in differentially structured landscapes. Proc. R. Soc. B Biol. Sci. 270, 569–575. https://doi.org/10.1098/rspb.2002.2292 (2003).
    Article  Google Scholar 

    29.
    Couvillon, M. J., Schurch, R. & Ratnieks, F. L. W. Dancing bees communicate a foraging preference for rural lands in high-level agri-environment schemes. Curr. Biol. 24, 1212–1215. https://doi.org/10.1016/j.cub.2014.03.072 (2014).
    CAS  Article  PubMed  Google Scholar 

    30.
    Bänsch, S., Tscharntke, T., Ratnieks, F. L. W., Härtel, S. & Westphal, C. Foraging of honey bees in agricultural landscapes with changing patterns of flower resources. Agric. Ecosyst. Environ. 291, 106792. https://doi.org/10.1016/j.agee.2019.106792 (2020).
    Article  Google Scholar 

    31.
    Walther-Hellwig, K. & Frankl, R. Foraging distances of Bombus muscorum, Bombus lapidarius, and Bombus terrestris (Hymenoptera, Apidae). J. Insect Behav. 13, 239–246. https://doi.org/10.1023/A:1007740315207 (2000).
    Article  Google Scholar 

    32.
    Chauzat, M. P. et al. Demographics of the European apicultural industry. PLoS ONE 8, e79018. https://doi.org/10.1371/journal.pone.0079018 (2013).
    ADS  Article  PubMed  PubMed Central  Google Scholar 

    33.
    Stanley, D. A., Gunning, D. & Stout, J. C. Pollinators and pollination of oilseed rape crops (Brassica napus L.) in Ireland: ecological and economic incentives for pollinator conservation. J. Insect Conserv. 17, 1181–1189. https://doi.org/10.1007/s10841-013-9599-z (2013).
    Article  Google Scholar 

    34.
    Westphal, C. et al. Measuring bee diversity in different European habitats and biogeographical regions. Ecol. Monogr. 78, 653–671. https://doi.org/10.1890/07-1292.1 (2008).
    Article  Google Scholar 

    35.
    Lebuhn, G., Droege, S., Connor, E., Gemmill-Herren, B. & Azzu, N. in Guidance for practioners 64 pp. (FAO, Rome, 2016).

    36.
    De Saeger, S. et al. (ed Rapporten van het Instituut voor Natuur- en Bosonderzoek 2016) (Instituut voor Natuur- en Bosonderzoek, Brussel, 2016).

    37.
    3QGIS_Development_Team. QGIS Geographic Information System, 2018).

    38.
    Oksanen, J. et al. Community Ecology Package ‘Vegan’. (2016). https://github.com/vegandevs/vegan.

    39.
    Meeus, I., de Graaf, D. C., Jans, K. & Smagghe, G. Multiplex PCR detection of slowly-evolving trypanosomatids and neogregarines in bumblebees using broad-range primers. J. Appl. Microbiol. 109, 107–115. https://doi.org/10.1111/j.1365-2672.2009.04635.x (2010).
    CAS  Article  PubMed  Google Scholar 

    40.
    Ravoet, J. et al. Widespread occurrence of honey bee pathogens in solitary bees. J. Invertebr. Pathol. 122, 55–58. https://doi.org/10.1016/j.jip.2014.08.007 (2014).
    Article  PubMed  Google Scholar 

    41.
    De Smet, L. et al. BeeDoctor, a versatile MLPA-based diagnostic tool for screening bee viruses. PLoS ONE 7, e47953. https://doi.org/10.1371/journal.pone.0047953 (2012).
    ADS  CAS  Article  PubMed  PubMed Central  Google Scholar 

    42.
    Parmentier, L. et al. Commercial bumblebee hives to assess an anthropogenic environment for pollinator support: A case study in the region of Ghent (Belgium). Environ. Monit. Assess. 186, 2357–2367. https://doi.org/10.1007/s10661-013-3543-2 (2014).
    CAS  Article  PubMed  Google Scholar 

    43.
    Rundlöf, M. et al. Seed coating with a neonicotinoid insecticide negatively affects wild bees. Nature 521, 77–80. https://doi.org/10.1038/nature14420 (2015).
    ADS  CAS  Article  PubMed  Google Scholar 

    44.
    Goulson, D. Bumblebees: Their Behaviour and Ecology (Oxford University Press, Oxford, 2003).
    Google Scholar 

    45.
    Bates, D., Machler, M., Bolker, B. M. & Walker, S. C. Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67, 1–48 (2015).
    Article  Google Scholar 

    46.
    Hedges, L. & Olkin, I. Statistical Methods for Meta-Analysis (Academic Press, Cambridge, 1985).
    Google Scholar 

    47.
    DeBach, P. The competitive displacement and coexistence principles. Annu. Rev. Entomol. 11, 183–212. https://doi.org/10.1146/annurev.en.11.010166.001151 (1966).
    Article  Google Scholar 

    48.
    Balfour, N. J., Gandy, S. & Ratnieks, F. L. W. Exploitative competition alters bee foraging and flower choice. Behav. Ecol. Sociobiol. 69, 1731–1738. https://doi.org/10.1007/s00265-015-1985-y (2015).
    Article  Google Scholar 

    49.
    Herbertssön, L., Lindström, S. A. M., Rundlof, M., Bornmarco, R. & Smith, H. G. Competition between managed honeybees and wild bumblebees depends on landscape context. Basic Appl. Ecol. 17, 609–616. https://doi.org/10.1016/j.baae.2016.05.001 (2016).
    Article  Google Scholar 

    50.
    Ropars, L., Dajoz, I., Fontaine, C., Muratet, A. & Geslin, B. Wild pollinator activity negatively related to honey bee colony densities in urban context. PLoS ONE 14, 16. https://doi.org/10.1371/journal.pone.0222316 (2019).
    CAS  Article  Google Scholar 

    51.
    Ellis, C., Park, K. J., Whitehorn, P., David, A. & Goulson, D. The neonicotinoid insecticide Thiacloprid impacts upon bumblebee colony development under field conditions. Environ. Sci. Technol. 51, 1727–1732. https://doi.org/10.1021/acs.est.6b04791 (2017).
    ADS  CAS  Article  PubMed  Google Scholar 

    52.
    Geslin, B., Gauzens, B., Thebault, E. & Dajoz, I. Plant pollinator networks along a gradient of urbanisation. PLoS ONE 8, e63421 (2013).
    ADS  Article  Google Scholar 

    53.
    Neame, L. A., Griswold, T. & Elle, E. Pollinator nesting guilds respond differently to urban habitat fragmentation in an oak-savannah ecosystem. Insect Conserv. Divers. 6, 57–66 (2013).
    Article  Google Scholar 

    54.
    Glaum, P., Simao, M.-C., Vaidya, C., Fitch, G. & Iulinao, B. Big city Bombus: Using natural history and land-use history to find significant environmental drivers in bumble-bee declines in urban development. R. Soc. Open Sci. 4, 170156 (2017).
    ADS  Article  Google Scholar 

    55.
    Normandin, E., Vereecken, N. J., Buddle, C. M. & Fournier, V. Taxonomic and functional trait diversity of wild bees in two urban settings. PeerJ 5, e3051 (2017).
    Article  Google Scholar 

    56.
    Moerman, R., Vanderplanck, M., Fournier, D., Jacquemart, A. L. & Michez, D. Pollen nutrients better explain bumblebee colony development than pollen diversity. Insect Conserv. Divers. 10, 171–179. https://doi.org/10.1111/icad.12213 (2017).
    Article  Google Scholar  More

  • in

    Long-term movements and activity patterns of platypus on regulated rivers

    1.
    Grant, T. & Denny, M. J. S. Distribution of the platypus in Australia with guidelines for management. A report to Australian National Parks and Wildlife Service. Mount King Ecological Surveys, Oberon, New South Wales (1991).
    2.
    Serena, M., Worley, M., Swinnerton, M. & Williams, G. Effect of food availability and habitat on the distribution of platypus (Ornithorhynchus anatinus) foraging activity. Aust. J. Zool. 49, 263–277 (2001).
    Article  Google Scholar 

    3.
    Serena, M., Thomas, J., Williams, G. & Officer, R. Use of stream and river habitats by the platypus, Ornithorhynchus anatinus, in an urban fringe environment. Aust. J. Zool. 46, 267–282 (1998).
    Article  Google Scholar 

    4.
    Bino, G. et al. The platypus: evolutionary history, biology, and an uncertain future. J. Mammal. 100, 308–327 (2019).
    PubMed  PubMed Central  Article  Google Scholar 

    5.
    Gardner, J. & Serena, M. Spatial-organization and movement patterns of adult male platypus, Ornithorhynchus anatinus (Monotremata, Ornithorhynchidae). Aust. J. Zool. 43, 91–103 (1995).
    Article  Google Scholar 

    6.
    Bino, G., Kingsford, R. T., Grant, T., Taylor, M. D. & Vogelnest, L. Use of implanted acoustic tags to assess platypus movement behaviour across spatial and temporal scales. Sci. Rep. 8, 5117 (2018).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    7.
    Grant, T., Grigg, G. C., Beard, L. & Augee, M. Movements and burrow use by platypuses, Ornithorhynchus anatinus, in the Thredbo River, New South Wales. In Platypus and Echidnas (ed. Augee, M. L.) 263–267 (The Royal Zoological Society of NSW, Sydney, 1992).
    Google Scholar 

    8.
    Griffiths, J., Kelly, T. & Weeks, A. Impacts of high flows on platypus movements and habitat use in an urban stream. Report to Melbourne Water. Cesar, VIC, Australia (2014).

    9.
    Gust, N. & Handasyde, K. Seasonal-variation in the ranging behavior of the platypus (Ornithorhynchus anatinus) on the Goulburn River, Victoria. Aust. J. Zool. 43, 193–208 (1995).
    Article  Google Scholar 

    10.
    Serena, M. Use of time and space by platypus (Ornithorhynchus anatinus: Monotremata) along a Victorian stream. J. Zool. 232, 117–131 (1994).
    Article  Google Scholar 

    11.
    Temple-Smith, P. & Grant, T. Uncertain breeding: a short history of reproduction in monotremes. Reprod. Fertil. Dev. 13, 487–497 (2001).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    12.
    Temple-Smith, P. D. Seasonal breeding biology of the platypus, Ornithorhynchus anatinus (Shaw, 1799), with special reference to the male (1973).

    13.
    Bino, G., Grant, T. R. & Kingsford, R. T. Life history and dynamics of a platypus (Ornithorhynchus anatinus) population: four decades of mark-recapture surveys. Sci. Rep. 5, 16073 (2015).
    ADS  PubMed  PubMed Central  Article  CAS  Google Scholar 

    14.
    Serena, M. & Williams, G. A. Movements and cumulative range size of the platypus (Ornithorhynchus anatinus) inferred from mark-recapture studies. Aust. J. Zool. 60, 352–359 (2012).
    Article  Google Scholar 

    15.
    Grant, T. Body temperatures of free-ranging platypuses, Ornithorhynchus anatinus (Monotremata), with observations on their use of burrows. Aust. J. Zool. 31, 117–122 (1983).
    Article  Google Scholar 

    16.
    Bethge, P., Munks, S., Otley, H. & Nicol, S. Activity patterns and sharing of time and space of platypuses, Ornithorhynchus anatinus, in a subalpine Tasmanian lake. J. Mammal. 90, 1350–1356 (2009).
    Article  Google Scholar 

    17.
    Macgregor, J. et al. Novel use of in-stream microchip readers to monitor wild platypuses. Pac. Conserv. Biol. 21, 101–101 (2015).
    Article  Google Scholar 

    18.
    Grant, T. Captures, capture mortality, age and sex ratios of platypuses, Ornithorhynchus anatinus, during studies over 30 years in the upper Shoalhaven River in New South Wales. Proc. Linn. Soc. NSW 125, 217–226 (2004).
    Google Scholar 

    19.
    Grant, T. & Temple-Smith, P. Conservation of the platypus, Ornithorhynchus anatinus: threats and challenges. Aquat. Ecosyst. Health Manag. 6, 5–18 (2003).
    Article  Google Scholar 

    20.
    Griffiths, J., Kelly, T. & Weeks, A. Net-avoidance behaviour in platypuses. Aust. Mammal. 35, 245–247 (2013).
    Article  Google Scholar 

    21.
    Griffiths, J. & Weeks, A. Impact of environmental flows on platypuses in a regulated river. Report to Melbourne Water. Cesar, VIC, Australia (2015).

    22.
    Hawke, T. et al. Fine-scale movements and interactions of platypuses, and the impact of an environmental flushing flow. Freshw. Biol. 2020(00), 1–12 (2020).
    Google Scholar 

    23.
    Erskine, W. et al. Bedform maintenance and pool destratification by the new environmental flows on the Snowy River downstream of the Jindabyne Dam, New South Wales. J. Proc. R. Soc. N. S. W. 150, 152–171 (2017).
    Google Scholar 

    24.
    Rohlfs, A. M. Rehabilitating the Snowy River: The influence of environmental flow releases on dissolved organic carbon supply and utilisation. PhD thesis, University of Technology, Sydney, Australia (2016).

    25.
    Pigram, J. J. Options for rehabilitation of Australia’s Snowy River: an economic perspective. Regul. Rivers Res. Manag. 16, 363–373 (2000).
    Article  Google Scholar 

    26.
    Davey, C. Mitta Mitta Biological Monitoring Program 2013–2014 Annual Report. Annual report prepared for the Murray Darling Basin Authority by The Murray–Darling Freshwater Research Centre, MDFRC Publication 45/2014, September, 88pp. Canberra, Australia (2014).

    27.
    Watts, R. J., Ryder, D. S. & Allan, C. Environmental monitoring of variable flow trials conducted at Dartmouth Dam, 2001/02-07/08-Synthesis of key findings and operational guidelines. Institute for Land, Water and Society Report Number 50, Charles Sturt University, Albury, NSW (2009).

    28.
    Vogelnest, L. & Woods, R. Medicine of Australian Mammals (CSIRO Publishing, Clayton, 2008).
    Google Scholar 

    29.
    McCullagh, P. & Nelder, J. A. Generalized Linear Models 2nd edn. (Chapman and Hall, London, 1989).
    Google Scholar 

    30.
    Bates, D. et al. Package “lme4”. Convergence 12, 2 (2015).
    Google Scholar 

    31.
    R Development Core Team. R: a language and environment for statistical computing (R Foundation for Statistical Computing, 2020).

    32.
    Lenth, R. emmeans: Estimated Marginal Means, aka Least-Squares Means. R package version 1.3.0. https://CRAN.R-project.org/package=emmean.

    33.
    Wood, S. N. & Augustin, N. H. GAMs with integrated model selection using penalized regression splines and applications to environmental modelling “Manuscript”. Ecol. Model. 157, 157–177 (2002).
    Article  Google Scholar 

    34.
    Wood, S. N. Generalized Additive Models: An Introduction with R (Chapman and Hall/CRC, Boca Raton, 2017).
    Google Scholar 

    35.
    Wood, S., Scheipl, F. & Wood, M. S. Package ‘gamm4’. Am. Stat. 45, 339 (2017).
    Google Scholar 

    36.
    Hazen, E. L. et al. Ontogeny in marine tagging and tracking science: technologies and data gaps. Mar. Ecol. Prog. Ser. 457, 221–240 (2012).
    ADS  Article  Google Scholar 

    37.
    McClellan, C. M. & Read, A. J. Complexity and variation in loggerhead sea turtle life history. Biol. Lett. 3, 592–594 (2007).
    PubMed  PubMed Central  Article  Google Scholar 

    38.
    Taylor, M. D., Babcock, R. C., Simpfendorfer, C. A. & Crook, D. A. Where technology meets ecology: acoustic telemetry in contemporary Australian aquatic research and management. Mar. Freshw. Res. 68, 1397–1402 (2017).
    Article  Google Scholar 

    39.
    Pillans, R., Fry, G., Steven, A. & Patterson, T. Environmental influences on long-term movement patterns of a euryhaline Elasmobranch (Carcharhinus leucas) within a subtropical estuary. Estuar. Coasts 43, 2152–2169 (2020).
    CAS  Article  Google Scholar 

    40.
    Otley, H. M., Munks, S. A. & Hindell, M. A. Activity patterns, movements and burrows of platypuses (Ornithorhynchus anatinus) in a sub-alpine Tasmanian lake. Aust. J. Zool. 48, 701–713 (2000).
    Article  Google Scholar 

    41.
    Hendry, A. P. et al. Evolution Illuminated: Salmon and Their Relatives (Oxford University Press, Oxford, 2003).
    Google Scholar 

    42.
    Stevick, P. et al. Population spatial structuring on the feeding grounds in North Atlantic humpback whales (Megaptera novaeangliae). J. Zool. 270, 244–255 (2006).
    Article  Google Scholar 

    43.
    Cutts, C., Brembs, B., Metcalfe, N. & Taylor, A. Prior residence, territory quality and life-history strategies in juvenile Atlantic salmon (Salmo salar L.). J. Fish Biol. 55, 784–794 (1999).
    Article  Google Scholar 

    44.
    Haley, M. P. Resource-holding power asymmetries, the prior residence effect, and reproductive payoffs in male northern elephant seal fights. Behav. Ecol. Sociobiol. 34, 427–434 (1994).
    Article  Google Scholar 

    45.
    Lindstedt, S. L., Miller, B. J. & Buskirk, S. W. Home range, time, and body size in mammals. Ecology 67, 413–418 (1986).
    Article  Google Scholar 

    46.
    Turner, F. B., Jennrich, R. I. & Weintraub, J. D. Home ranges and body size of lizards. Ecology 50, 1076–1081 (1969).
    Article  Google Scholar 

    47.
    Slavenko, A., Itescu, Y., Ihlow, F. & Meiri, S. Home is where the shell is: predicting turtle home range sizes. J. Anim. Ecol. 85, 106–114 (2016).
    PubMed  Article  Google Scholar 

    48.
    McNab, B. K. Bioenergetics and the determination of home range size. Am. Nat. 97, 133–140 (1963).
    Article  Google Scholar 

    49.
    Greenwood, P. J. Mating systems, philopatry and dispersal in birds and mammals. Anim. Behav. 28, 1140–1162 (1980).
    Article  Google Scholar 

    50.
    Soderquist, T. R. & Serena, M. Juvenile behaviour and dispersal of chuditch (Dasyurus geoffroii) (Marsupialia: Dasyuridae). Aust. J. Zool. 48, 551–560 (2000).
    Article  Google Scholar 

    51.
    Cockburn, A., Scott, M. P. & Scotts, D. J. Inbreeding avoidance and male-biased natal dispersal in Antechinus spp. (Marsupialia: Dasyuridae). Anim. Behav. 33, 908–915 (1985).
    Article  Google Scholar 

    52.
    Furlan, E. et al. Dispersal patterns and population structuring among platypuses, Ornithorhynchus anatinus, throughout south-eastern Australia. Conserv. Genet. 14, 837–853 (2013).
    CAS  Article  Google Scholar 

    53.
    Kolomyjec, S. H. Relationships of northern platypus (Ornithorhynchus anatinus) populations: a molecular approach (2010).

    54.
    Faragher, R., Grant, T. & Carrick, F. Food of the platypus (Ornithorhynchus anatinus) with notes on the food of brown trout (Salmo trutta) in the Shoalhaven River, NSW. Aust. J. Ecol. 4, 171–179 (1979).
    Article  Google Scholar 

    55.
    Todd, C. R., Ryan, T., Nicol, S. J. & Bearlin, A. R. The impact of cold water releases on the critical period of post-spawning survival and its implications for Murray cod (Maccullochella peelii peelii): a case study of the Mitta Mitta River, southeastern Australia. River Res. Appl. 21, 1035–1052 (2005).
    Article  Google Scholar 

    56.
    Koehn, J., Doeg, T., Harrington, D. & Milledge, G. The effects of Dartmouth Dam on the aquatic fauna of the Mitta Mitta River. Report to the Murray-Darling Basin Commission. Arthur Rylah Institute for Environmental Research, Melbourne (1995).

    57.
    Hawke, T., Bino, G., & Kingsford, R. T. Damming insights: impacts and implications of river regulation on platypus populations. Aquat Conserv Mar Freshw Ecosystems. in press. (2021).

    58.
    Hawke, T., Bino, G. & Kingsford, R. T. Platypus: paucities and peril. A response to: limitations on the use of historical and database sources to identify changes in distribution and abundance of the platypus. Glob. Ecol. Conserv. 21, e00857 (2019).
    Article  Google Scholar 

    59.
    Hawke, T., Bino, G. & Kingsford, R. T. A silent demise: historical insights into population changes of the iconic platypus (Ornithorhynchus anatinus). Glob. Ecol. Conserv. 20, e00720 (2019).
    Article  Google Scholar 

    60.
    Bino, G., Kingsford, R. T. & Wintle, B. A. A stitch in time—synergistic impacts to platypus metapopulation extinction risk. Biol. Conserv. 242, 108399 (2020).
    Article  Google Scholar  More