More stories

  • in

    Rainfall affects interactions between plant neighbours

    Lebrija-Trejos, E., Hernández, A. & Wright, S. J. Nature https://doi.org/10.1038/s41586-023-05717-1 (2023).Article 

    Google Scholar 
    Chesson, P. J. Ecol. 106, 1773–1794 (2018).Article 

    Google Scholar 
    Barabás, G., D’Andrea, R. & Stump, S. M. Ecol. Monogr. 88, 277–303 (2018).Article 

    Google Scholar 
    Broekman, M. J. E. et al. Ecol. Lett. 22, 1957–1975 (2019).Article 
    PubMed 

    Google Scholar 
    Freckleton, R. P. & Lewis, O. T. Proc. R. Soc. B 273, 2909–2916 (2006).Article 
    PubMed 

    Google Scholar 
    Bagchi, R. et al. Nature 506, 85–88 (2014).Article 
    PubMed 

    Google Scholar 
    Chen, L. et al. Science 366, 124–128 (2019).Article 
    PubMed 

    Google Scholar 
    Milici, V. R., Dalui, D., Mickley, J. G. & Bagchi, R. J. Ecol. 108, 1800–1809 (2020).Article 

    Google Scholar 
    Song, X. & Corlett, R. T. Oikos 2022, e08509 (2022).Article 

    Google Scholar 
    Engelbrecht, B. M. J. et al. Nature 447, 80–82 (2007).Article 
    PubMed 

    Google Scholar 
    Krishnadas, M. & Stump, S. M. J. Ecol. 109, 2137–2151 (2021).Article 

    Google Scholar 
    Van Dyke, M. N., Levine, J. M. & Kraft, N. J. B. Nature 611, 507–511 (2022).Article 
    PubMed 

    Google Scholar  More

  • in

    Seasonal variation in the lipid content of Fraser River Chinook Salmon (Oncorhynchus tshawytscha) and its implications for Southern Resident Killer Whale (Orcinus orca) prey quality

    Caughley, G. Directions in conservation biology. J. Anim. Ecol. 63, 215 (1994).Article 

    Google Scholar 
    Fisheries and Oceans Canada. National recovery strategy for northern and southern resident killer whales (Orcinus orca) in Canada [proposed]. vol. Species at (2018).National Marine Fisheries Service. Recovery Plan for Southern Resident Killer Whales (Orcinus orca). (2008).Barrett-Lennard, L. G. & Ellis, G. M. Population Structure and Genetic Variability in Northeastern Pacific Killer Whales: Towards an Assessment of Population Viability. DFO Can. Sci. Advis. Secr. Res. Deocument 2001/065 65 (2001).DFO. Evaluation of the scientific evidence to inform the probability of effectiveness of mitigation measures in reducing shipping-related noise levels received by southern resident killer whales. CSAS Science Advisory Report vol. 2017/041 (2017).Ross, P. S., Ellis, G. M., Ikonomou, M. G. & Addison, R. F. High PCB concentrations in free-ranging Pacific Killer Whales, Orcinus orca: Effects of age, sex and dietary preference. Mar. Pollut. Bull. 40, 504–515 (2000).Article 
    CAS 

    Google Scholar 
    Ward, E. J., Holmes, E. E. & Balcomb, K. C. Quantifying the effects of prey abundance on killer whale reproduction. J. Appl. Ecol. 46, 632–640 (2009).Article 

    Google Scholar 
    Ford, J. K. B., Ellis, G. M., Olesiuk, P. F. & Balcomb, K. C. Linking killer whale survival and prey abundance: Food limitation in the oceans’ apex predator ?. Biol. Lett. 6, 139–142 (2010).Article 
    PubMed 

    Google Scholar 
    Ford, J. K. B. et al. Dietary specialization in two sympatric populations of killer whales (Orcinus orca) in coastal British Columbia and adjacent waters. Can. J. Zool. 76, 1456–1471 (1998).Article 

    Google Scholar 
    Ford, J. K. B., Ellis, G. M. & Olesiuk, P. F. Linking Prey and Population Dynamics Did Food Limitation Cause Recent Declines of RKW in BC, vol. 3848 (2005).O’Neill, S. M., Ylitalo, G. M. & West, J. E. Energy content of Pacific salmon as prey of northern and southern resident killer whales. Endanger. Species Res. 25, 265–281 (2014).Article 

    Google Scholar 
    Ford, J. K. B. & Ellis, G. M. Selective foraging by fish-eating killer whales Orcinus orca in British Columbia. Mar. Ecol. Prog. Ser. 316, 185–199 (2006).Article 
    ADS 

    Google Scholar 
    Jeffrey, K. M., Côté, I. M., Irvine, J. R. & Reynolds, J. D. Changes in body size of Canadian Pacific salmon over six decades. Can. J. Fish. Aquat. Sci. 74, 191–201 (2017).Article 

    Google Scholar 
    Ohlberger, J., Schindler, D. E., Ward, E. J., Walsworth, T. E. & Essington, T. E. Resurgence of an apex marine predator and the decline in prey body size. Proc. Natl. Acad. Sci. USA https://doi.org/10.1073/pnas.1910930116 (2019).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Ohlberger, J., Ward, E. J., Schindler, D. E. & Lewis, B. Demographic changes in Chinook salmon across the Northeast Pacific Ocean. Fish Fish. 19, 533–546 (2018).Article 

    Google Scholar 
    Bigler, B. S., Welch, D. W. & Helle, J. H. A review of size trends among North Pacific salmon (Oncorhynchus spp.). Can. J. Fish. Aquat. Sci. 53, 455–465 (2011).Article 

    Google Scholar 
    Hanson, M. B. et al. Species and stock identification of prey consumed by endangered southern resident killer whales in their summer range. Endanger. Species Res. 11, 69–82 (2010).Article 
    ADS 

    Google Scholar 
    Losee, J. P., Kendall, N. W. & Dufault, A. Changing salmon: An analysis of body mass, abundance, survival, and productivity trends across 45 years in Puget Sound. Fish Fish. 20, 934–951 (2019).Article 

    Google Scholar 
    Riddell, B. et al. Assessment of Status and Factors for Decline of Southern BC Chinook Salmon: Independent Panel’s Report (2013).DFO. Integrated Biological Status of Southern British Columbia Chinook Salmon (Oncorhynchus tshawytscha) Under the Wild Salmon Policy. DFO Can. Sci. Advis. Sec. Sci. Advis. Rep. 2016/042, 15 (2016).
    Google Scholar 
    COSEWIC. COSEWIC assessment and status report on the Chinook Salmon Oncorhynchus tshawytscha, Designatable Units in Southern British Columbia, in Canada. (2019).Pacific Salmon Commission Joint Chinook Technical Committee. Annual Report of Catch and Escapement for 2021. Tcchinook (13)-01 (2021).Quinn, T. P. Behavior and ecology of Pacific Salmon and trout. Fish Fish. 7, 75–76 (2004).
    Google Scholar 
    Brett, J. R. Energetics. In Phsyiological Ecology of Pacific Salmon (eds Groot, C. et al.) 1–68 (UBC Press, 1995).
    Google Scholar 
    Chamberlain, M. W. & Parken, C. Utilizing the Albion test fishery as an in-season predictor of run size of the Fraser River spring and summer age 52 Chinook. DFO Can. Sci. Advis. Sec. Res. Doc. 2012, 42 (2012).
    Google Scholar 
    Schoener, T. W. Theory of feeding strategies. Annu. Rev. Ecol. Syst. 2, 369–404 (1971).Article 

    Google Scholar 
    Williams, R. et al. Competing conservation objectives for predators and prey: Estimating Killer Whale prey requirements for Chinook Salmon. PLoS ONE 6, e26738 (2011).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Courtney, K. R., Falke, J. A., Cox, M. K. & Nichols, J. Energetic status of Alaskan Chinook Salmon: Interpopulation comparisons and predictive modeling using bioelectrical impedance analysis. North Am. J. Fish. Manag. https://doi.org/10.1002/nafm.10398 (2019).Article 

    Google Scholar 
    Pothoven, S. A. et al. Reliability of bioelectrical impedance analysis for estimating whole-fish energy density and percent lipids. Trans. Am. Fish. Soc. 137, 1519–1529 (2008).Article 

    Google Scholar 
    Crossin, G. T. & Hinch, S. G. A Nonlethal, rapid method for assessing the somatic energy content of migrating adult pacific salmon. Trans. Am. Fish. Soc. 134, 184–191 (2005).Article 

    Google Scholar 
    Colt, J. & Shearer, K. D. Evaluation of the Use of the Torry Fish Fatmeter to Non-Lethally Estimate Lipid in Adult Salmon (2001).Hanson, K. C., Ostrand, K. G., Gannam, A. L. & Ostrand, S. L. Comparison and validation of nonlethal techniques for estimating condition in Juvenile Salmonids. Trans. Am. Fish. Soc. 139, 1733–1741 (2010).Article 

    Google Scholar 
    Naughton, G., Caudill, C. & Clabough, T. Migration Behavior and Spawning Success of Spring Chinook Salmon in Fall Creek and the North Fork Middle Fork Willamette River: Relationship Among Fate, Fish Condition, and Environmental Factors, 2011. (2012).Folch, J., Lees, M. & Sloane Stanley, G. A simple method for the isolation and purification of total lipides from animal tissues. J. Biol. Chem. 226, 497–509 (1957).Article 
    CAS 
    PubMed 

    Google Scholar 
    Post, J. R. & Parkinson, E. A. Energy allocation strategy in young fish: Allometry and survival. Ecology 82, 1040–1051 (2001).Article 

    Google Scholar 
    Arrington, D. A., Davidson, B. K., Winemiller, K. O. & Layman, C. A. Influence of life history and seasonal hydrology on lipid storage in three neotropical fish species. J. Fish Biol. 68, 1347–1361 (2006).Article 
    CAS 

    Google Scholar 
    Holty, B. L. & Ciruna, K. A. Conservation units for Pacific Salmon under the Wild Salmon Policy. DFO Can. Sci. Advis. Sec. Res. Doc 2007/070, 350 (2007).
    Google Scholar 
    PSC. Catch and Escapement of Chinook Under Pacific Salmon Commission Jurisdiction, 2001 (PSC, 2002).
    Google Scholar 
    Waples, R. S., Teel, D. J., Myers, J. M. & Marshall, A. R. Life-history divergence in Chinook Salmon: Historic contingency and parallel evolution. Evolution 58, 386–403 (2004).PubMed 

    Google Scholar 
    Beacham, T. D. et al. Pacific rim population structure of chinook salmon as determined from microsatellite analysis. Trans. Am. Fish. Soc. 135, 1604–1621 (2006).Article 
    CAS 

    Google Scholar 
    Crossin, G. T. et al. Energetics and morphology of sockeye salmon: Effects of upriver migratory distance and elevation. J. Fish Biol. 65, 788–810 (2004).Article 

    Google Scholar 
    MacDonald, B. In-Season Forecasting of Fraser Chinook Salmon Using Genetic Stock Identification of Test Fishery Data By (2016).Parken, C. K., Candy, J. R., Irvine, J. R. & Beacham, T. D. Genetic and coded wire tag results combine to allow more-precise management of a complex Chinook salmon aggregate. North Am. J. Fish. Manag. 28, 328–340 (2008).Article 

    Google Scholar 
    Mann, R., Peery, C., Pinson, A. & Anderson, C. Energy use, migration times, and spawning success of adult spring–summer Chinook salmon returning to spawning areas in the South Fork Salmon River in Central Idaho: 2002–2007. Technical report 2009–4 http://www.cnr.uidaho.edu/uiferl/pdfreports/SFS_Tech_Report_2009-4_Final.pdf (2009).Hearsey, J. W. & Kinziger, A. P. Diversity in sympatric chinook salmon runs: Timing, relative fat content and maturation. Environ. Biol. Fishes 98, 413–423 (2015).Article 

    Google Scholar 
    Arimitsu, M. L. et al. Heatwave-induced synchrony within forage fish portfolio disrupts energy flow to top pelagic predators. Glob. Chang. Biol. 27, 1859–1878 (2021).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Lloret-Lloret, E. et al. Small pelagic fish fitness relates to local environmental conditions and trophic variables. Prog. Oceanogr. 202, 102745 (2022).Article 

    Google Scholar 
    Mesa, M. G. & Magie, C. D. Evaluation of energy expenditure in adult spring Chinook salmon migrating upstream in the Columbia River Basin: An assessment based on sequential proximate analysis. River Res. Appl. 22, 1085–1095 (2006).Article 

    Google Scholar 
    Crossin, G. T., Hinch, S. G., Farrell, A. P., Higgs, D. A. & Healey, M. C. Somatic energy of sockeye salmon Oncorhynchus nerka at the onset of upriver migration: A comparison among ocean climate regimes. Fish. Oceanogr. 13, 345–349 (2004).Article 

    Google Scholar 
    Roni, P. & Quinn, T. P. Geographic variation in size and age of North American Chinook salmon. North Am. J. Fish. Manag. 15, 325–345 (1995).Article 

    Google Scholar 
    Hendry, A. P., Berg, O. K., Quinn, T. P. & Condition, T. P. Condition dependence and adaptation-by-time: Breeding date, life history, and energy allocation within a population of salmon. Oikos 85, 499–514 (1999).Article 

    Google Scholar 
    Hanson, M. B. et al. Endangered predators and endangered prey: Seasonal diet of Southern Resident killer whales. PLoS ONE 16, e0247031 (2021).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Weitkamp, L. A. Marine distributions of Chinook Salmon from the West Coast of North America determined by coded wire tag recoveries. Trans. Am. Fish. Soc. 139, 147–170 (2010).Article 

    Google Scholar 
    Shields, M. W., Lindell, J. & Woodruff, J. Declining spring usage of core habitat by endangered fish-eating killer whales reflects decreased availability of their primary prey. Pac. Conserv. Biol. 24, 189–193 (2018).Article 

    Google Scholar 
    Brown, G. S. et al. Pre-COSEWIC review of southern British Columbia Chinook Salmon (Oncorhynchus tshawytscha) conservation units Part I: Background. Can. Sci. Advis. Sec. Res. Doc. 2019/11, 67 (2019).
    Google Scholar 
    NOAA Fisheries West Coast & Washington Department of Fish and Wildlife. Southern Resident Killer Whale Priority Chinook Stocks Report. https://www.westcoast.fisheries.noaa.gov/publications/protected_species/marine_mammals/killer_whales/recovery/srkw_priority_chinook_stocks_conceptual_model_report___list_22june2018.pdf (2018).Chalifour, L. et al. Chinook salmon exhibit long-term rearing and early marine growth in the fraser river, british columbia, a large urban estuary. Can. J. Fish. Aquat. Sci. 78, 539–550 (2021).Article 
    CAS 

    Google Scholar 
    Lamperth, J. S., Quinn, T. P. & Zimmerman, M. S. Levels of stored energy but not marine foraging patterns differentiate seasonal ecotypes of wild and hatchery steelhead (Oncorhynchus mykiss) returning to the Kalama river, Washington. Can. J. Fish. Aquat. Sci. 74, 157–167 (2017).Article 
    CAS 

    Google Scholar 
    Von Biela, V. R. et al. Extreme reduction in nutritional value of a key forage fish during the pacific marine heatwave of 2014–2016. Mar. Ecol. Prog. Ser. 613, 171–182 (2019).Article 
    ADS 

    Google Scholar 
    Healey, M. C. Life history of Chinook Salmon (Oncorhynchus tshawytscha). In Pacific Salmon Life Histories (eds Groot, C. & Margolis, L.) 313–393 (University of British Columbia Press, 1991).
    Google Scholar 
    Freshwater, C. et al. An integrated model of seasonal changes in stock composition and abundance with an application to Chinook salmon. PeerJ 9, 1–27 (2021).Article 

    Google Scholar 
    Couture, F., Oldford, G., Christensen, V., Barrett-lennard, L. & Walters, C. Requirements and availability of prey for northeastern pacific southern resident killer whales. PLoS ONE 17, e0270523 (2022).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    DFO. Government of Canada Takes Action to Address Fraser River Chinook Decline (DFO, 2019).
    Google Scholar 
    Brown, R. F. & Musgrave, M. M. Preliminary Catalogue of Salmon Steams and Escapements of Misson-Harrison Sub District. Fisheries and Marine Service Data Report No. 133 (1979).Manzon, C. I. & Marshall, D. E. Catalogue of salmon streams and spawning escapements of Cariboo subdistrict. Can. Data Rep. Fish. Aquat. Sci. 211, 51 (1980).
    Google Scholar 
    Marshall, D. E. & Manzon, C. I. Catalogue of Salmon Streams and Spawning Escapements of the Prince George Subdistrict (Department of Fisheries and Oceans Fisheries and Marine Services Data Report N0o. 79, 1980).
    Google Scholar 
    Olmsted, W., Whelen, M. & Stewart, R. 1980 Investigations of fall-spawning chinook salmon (Oncorhynchus tshawytscha), Quesnel, blackwater (west road) and cottonwood river drainages, B.C. 34, 131–134 (1981).Brown, R. F., Musgrave, M. M. & Marshall, D. E. Catalogue of salmon streams and spawning escapements for Kamloops sub-district. Fish. Mar. Serv. Data Rep. 151, 226 (1979).
    Google Scholar 
    DFO. Information Document to Assist Development of a Fraser Chinook Management Plan 56 (DFO, 2006).
    Google Scholar 
    Kosakoski, G. T. & Hamilton, R. E. Water Requirements for the Fisheries Resource of the Nicola River, B.C. Can. Manuscr. Rep. Fish. Aquat. Sci. 140 (1982). More

  • in

    Two odorant receptors regulate 1-octen-3-ol induced oviposition behavior in the oriental fruit fly

    Insect rearingWT B. dorsalis were collected from Haikou, Hainan province, China, in 2008. They were maintained at the Key Laboratory of Entomology and Pest Control Engineering in Chongqing at 27 ± 1 °C, 70 ± 5% relative humidity, with a 14-h photoperiod. Adult flies were reared on an artificial diet containing honey, sugar, yeast powder, and vitamin C. Newly hatched larvae were transferred to an artificial diet containing corn and wheat germ flour, yeast powder, agar, sugar, sorbic acid, linoleic acid, and filter paper.Behavioral assaysDouble trap lure assays were set up to compare the olfactory preferences of gravid and virgin females in a 20 × 20 × 20 cm transparent cage with evenly distributed holes (diameter = 1.5 mm) on the side walls. The traps were refitted from inverted 50-mL centrifuge tubes and were placed along the diagonal of the cage. The top of each trap was pierced with a 1-mL pipette tip, which was shortened to ensure flies could access the trap from the pipette. For the olfactory preference assay with mango, one trap was loaded with 60 mg mango flesh and the other trap with 20 μL MO in the cap of a 200-μL PCR tube. For the olfactory preference assay with 1-octen-3-ol (≥98%, sigma, USA), one trap was loaded with 20 μL 10% (v/v) 1-octen-3-ol diluted in MO, and the other with 20 μL MO. A cotton ball soaked in water was placed at the center of the cage to provide water for the flies. Groups of 30 female flies were introduced into the cage for each experiment, and each experiment was repeated to provide eight biological replicates. All experiments commenced at 10 am to ensure circadian consistency. The number of flies in each trap was counted every 2 h for 24 h. We compared the preferences of 3-day-old immature females, 15-day-old virgin females, and 15-day-old mated females. The olfactory preference index was calculated using the following formula41: (number of flies in mango/odorant trap – number of flies in control trap)/total number of flies.Oviposition behavior was monitored in a 10 × 10 × 10 cm transparent cage with evenly distributed holes on the side walls as above. A 9-cm Petri dish filled with 1% agar was served as an oviposition substrate, and the mango flesh, 10% (v/v) 1-octen-3-ol or MO were added at opposite edges of the dish. We tested the preference of flies for different substrates: (1) ~60 mg of mango flesh on one edge and 20 μL of MO on the other; (2) 20 μL of 1-octen-3-ol on one edge and 20 μL of MO on the other; (3) ~60 mg mango flesh on one edge and 20 μL of 1-octen-3-ol on the other; and (4) ~60 mg mango flesh plus 20 μL 1-octen-3-ol on one side and ~60 mg of mango flesh plus 20 μL MO on the other. The agar disc was covered in a pierced plastic wrap to mimic fruit skin, encouraging flies extend their ovipositor into the plastic wrap to lay eggs. The agar disc was placed at the center of the cage, and we introduced eight 15-day-old gravid females. Two Sony FDR-AX40 cameras recorded the behavior of the flies for 24 h, one fixed above the cage to record the tracks and the other placed in front of the cage to record the oviposition behavior. Based on the results from double traps luring assays, a 3 h duration (6–9 h) of the videos was selected to analyze the tracks and spent time of all flies in observed area (the surface of Petri dish). The videos were analyzed using EthoVision XT v16 (Noldus Information Technology) to determine the total time of all flies spent on each side in seconds and the total distance of movement in centimeters, and the tracks were visualized in the form of heat maps17. The number of eggs laid by the eight flies in each experiment was counted under a CNOPTEC stereomicroscope, and each experimental group comprised 7–16 replicates.Annotation of B. dorsalis OR genesD. melanogaster amino acid sequences downloaded from the National Center for Biotechnology Information (https://www.ncbi.nlm.nih.gov/) were used as BLASTP queries against the B. dorsalis amino acid database with an identity cut-off of 30%. The candidate OR genes were compared with deep transcriptome data from B. dorsalis antennae42, maxillary palps and proboscis, and other tissues.Cloning of candidate B. dorsalis OR genesHigh-fidelity PrimerSTAR Max DNA polymerase (TaKaRa, Dalian, China) was used to amplify the full open reading frame of BdorOR genes by nested PCR using primers (Supplementary Table 2) designed according to B. dorsalis genome data. Each 25-μL reaction comprised 12.5 μL 2 × PrimerSTAR Max Premix (TaKaRa), 10.5 μL ultrapure water, 1 μL of each primer (10 μM), and 1 μL of the cDNA template. An initial denaturation step at 98 °C for 2 min was followed by 35 cycles of 10 s at 98 °C, 15 s at 55 °C and 90 s at 72 °C, and a final extension step of 10 min at 72 °C. Purified PCR products were transferred to the vector pGEM-T Easy (Promega, Madison, WI) for sequencing (BGI, Beijing, China).Transcriptional profilingTotal RNA was extracted from (i) male and female antennae, maxillary palps, head cuticle (without antenna, maxillary palps, and proboscis), proboscis, legs, wings and ovipositors, and (ii) from the heads of 15-day-old virgin and mated females using TRIzol reagent (Invitrogen, Carlsbad, CA). Genomic DNA was eliminated with RNase-free DNase I (Promega) and first-strand cDNA was synthesized from 1 µg total RNA using the PrimeScript RT reagent kit (TaKaRa). Standard curves were used to evaluate primer efficiency (Supplementary Table 3) with fivefold serial dilutions of cDNA. Quantitative real-time PCR (qRT-PCR) was carried out using a CFX Connect Real-Time System (Bio-Rad, Hercules, CA) in a total reaction volume of 10 µL containing 5 μL SYBR Supermix (Novoprotein, Shanghai, China), 3.9 μL nuclease-free water, 0.5 μL cDNA (~200 ng/μL) and 0.3 μL of the forward and reverse primers (10 μM). We used α-tubulin (GenBank: GU269902) and ribosomal protein S3 (GenBank: XM_011212815) as internal reference genes. Four biological replicates were prepared for each experiment. Relative expression levels were determined using the 2−∆∆Ct method43, and data were analyzed using SPSS v20.0 (IBM).Two-electrode voltage clamp electrophysiological recordingsVerified PCR products representing candidate B. dorsalis OR genes and BdorOrco were transferred to vector pT7Ts for expression in oocytes. The plasmids were linearized for the synthesis of cRNAs using the mMESSAGE mMACHINE T7 Kit (Invitrogen, Lithuania). The purified cRNA was diluted to 2 µg/µL and ~60 ng cRNA was injected into X. laevis oocytes. The oocytes were pre-treated with 1.5 mg/mL collagenase I (GIBCO, Carlsbad, CA) in washing buffer (96 mM NaCl, 5 mM MgCl2, 2 mM KCl, 5 mM HEPES, pH 7.6) for 30–40 min at room temperature before injection. After incubation for 2 days at 18 °C in Ringer’s solution (96 mM NaCl, 5 mM MgCl2, 2 mM KCl, 5 mM HEPES, 0.8 mM CaCl2), the oocytes were exposed to different concentrations of 1-octen-3-ol diluted in Ringer’s solution from a 1 M stock in DMSO. Odorant-induced whole-cell inward currents were recorded from injected oocytes using a two-electrode voltage clamp and an OC-725C amplifier (Warner Instruments, Hamden, CT) at a holding potential of –80 mV. The signal was processed using a low-pass filter at 50 Hz and digitized at 1 kHz. Oocytes injected with nuclease-free water served as a negative control. Data were acquired using a Digidata 1550 A device (Warner Instruments, Hamden, CT) and analyzed using pCLAMP10.5 software (Axon Instruments Inc., Union City, CA).Calcium imaging assayVerified PCR products representing candidate B. dorsalis OR genes and BdorOrco were transferred to vector pcDNA3.1(+) along with an mCherry tag that confers red fluorescence to confirm transfection. High-quality plasmid DNA was prepared using the Qiagen plasmid MIDIprep kit (QIAgen, Düsseldorf, Germany). The B. dorsalis OR and BdorOrco plasmids were co-transfected into HEK 293 cell using TransIT-LT1 transfection reagent (Mirus Bio LLC, Japan) in 96-well plates. The fluorescent dye Fluo-4 AM (Invitrogen) was prepared as a 1 mM stock in DMSO and diluted to 2.5 μM in Hanks’ balanced salt solution (HBSS, Invitrogen, Lithuania) to serve as a calcium indicator. The cell culture medium was removed 24–30 h after transfection and cells were rinsed three times with HBSS before adding Fluo 4-AM and incubating the cells for 1 h in the dark. After three rinses in HBSS, 99 μL of fresh HBSS was added to each well before testing in the dark with 1 μL of diluted 1-octen-3-ol. Fluorescent images were acquired on a laser scanning confocal microscope (Zeiss, Germany). Fluo 4-AM was excited at 488 nm and mCherry at 555 nm. The relative change in fluorescence (ΔF/F0) was used to represent the change in Ca2+, where F0 is the baseline fluorescence and ΔF is the difference between the peak fluorescence induced by 1-octen-3-ol stimulation and the baseline. The healthy and successfully transfected cells (red when excited at 555 nm) were used for analysis. The final concentration of 10−4 M was initially used to screen corresponding ORs, and then to determine the response of screened ORs to stimulation with different concentrations of 1-octen-3-ol. Each concentration of 1-octen-3-ol was tested in triplicate. Concentration–response curves were prepared using GraphPad Prism v8.0 (GraphPad Software).Genome editingThe exon sequences of BdorOR7a-6 and BdorOR13a were predicted using the high-quality B. dorsalis genome assembly. Each gRNA sequence was 20 nucleotides in length plus NGG as the protospacer adjacent motif (PAM). The potential for off-target mutations was evaluated by using CasOT to screen the B. dorsalis genome sequence. Each gRNA was synthesized using the GeneArt Precision gRNA Synthesis Kit (Invitrogen) and purified using the GeneArt gRNA Clean-up Kit (Invitrogen). Embryos were microinjected as previously described20. Purified gRNA and Cas9 protein from the GeneArt Platinum Cas9 Nuclease Kit (Invitrogen) were mixed and diluted to final concentrations of 600 and 500 ng/µL, respectively. Fresh eggs (laid within 20 min) were collected and exposed to 1% sodium hypochlorite for 90 s to soften the chorion. The eggs were fixed on glass slides and injected with the mix of gRNA and Cas9 protein at the posterior pole using an IM-300 device (Narishige, Tokyo, Japan) and needles prepared using a Model P-97 micropipette puller (Sutter Instrument Co, Novato, CA). Eggs were injected with nuclease-free water as a negative control. Injection was completed within 2 h. The injected embryos were cultured in a 27 °C incubator and mortality was recorded during subsequent development.G0 mutants were screened as previously described20. G0 adult survivors were individually backcrossed to WT flies (single pair) to collect G1 offspring. Genomic DNA was extracted from G0 individuals after oviposition using the DNeasy Blood & Tissue Kit (Qiagen). The region surrounding each gRNA target was amplified by PCR using the extracted DNA as a template, the specific primers listed in Supplementary Table 2, and 2 × Taq PCR MasterMix (Biomed, Beijing, China). PCR products were analyzed by capillary electrophoresis using the QIAxcel DNA High Resolution Kit (Qiagen). PCR products differing from the WT alleles were purified and transferred to the vector pGEM-T Easy for sequencing. To confirm the mutation was inherited, genomic DNA was also extracted from one mesothoracic leg of G1 flies using InstaGene Matrix (Bio-Rad, Hercules, CA) and was analyzed as above. To avoid potential off-target mutations, heterozygous G1 mutants were backcrossed to WT flies more than 10 generations before self-crossing to generate homozygous mutant flies.Electroantennogram (EAG) recordingThe antennal responses of 15-day-old B. dorsalis adults to 1-octen-3-ol were determined by EAG recording (Syntech, the Netherlands) as previously reported20. Briefly, antennae were fixed to two electrodes using Spectra 360 electrode gel (Parker, Fairfield, NJ, USA). The signal response was amplified using an IDAC4 device and collected using EAG-2000 software (Syntech). Before each experiment, 1-octen-3-ol and other three volatiles (ethyl tiglate, ethyl acetate, ethyl butyrate) were diluted to 10%, 1% and 0.1% (v/v) with MO to serve as the electrophysiological stimulus, and MO was used as a negative control. A constant air flow (100 mL/min) was produced using a controller (Syntech) to stimulate the antenna. We then placed 10 µL of each dilution or MO onto a piece of filter paper (5 × 1 cm), and the negative control (MO) was applied before and after the diluted odorants to calibrate the response signal. The EAG responses at each concentration were recorded for 15–20 antennae, and each concentration was recorded twice. Each test lasted 1 s, and the interval between tests was 30 s. EAG response data from WT and mutant flies for the diluted odorants were analyzed using Student’s t test with SPSS v20.0.Molecular docking and site-directed mutagenesisThe three dimensional-structures of BdorOR7a-6 and BdorOR13a were modeled using AlphaFold 2.044. The quality and rationality of each protein structure was evaluated online using a PROCHECK Ramachandran plot in SAVES 6.0 (https://saves.mbi.ucla.edu/). AutoDock Vina 1.1.2 was used for docking analysis, and the receptor protein structure and ligand molecular structure were pre-treated using AutoDock 4.2.6. The docking parameters were set according to the protein structure and active sites, and the optimal docking model was selected based on affinity (kcal/mol). Docking models were imported into Pymol and Discovery Studio 2016 Client for analysis and image processing. Based on the molecular docking data, three residues (Asn86 in OR7a-6, Asp320, and Lys323 in OR13a) were replaced with alanine by site-directed mutagenesis45 using the primers listed in Supplementary Table 2. Calcium imaging assays and molecular docking of mutated proteins were then carried out as described above.Statistics reproducibilityAll of the olfactory preference assays, oviposition bioassays, expression profiles analysis, EAG recording assays were analyzed using Student’s t-test (*p  More

  • in

    The degree of urbanisation reduces wild bee and butterfly diversity and alters the patterns of flower-visitation in urban dry grasslands

    Sánchez-Bayo, F. & Wyckhuys, K. A. Worldwide decline of the entomofauna: A review of its drivers. Biol. Conserv. 232, 8–27. https://doi.org/10.1016/j.biocon.2019.01.020 (2019).Article 

    Google Scholar 
    van Klink, R. et al. Meta-analysis reveals declines in terrestrial but increases in freshwater insect abundances. Science 368, 417–420. https://doi.org/10.1126/science.aax9931 (2020).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Wagner, D. L. Insect declines in the anthropocene. Annu. Rev. Entomol. 65, 457–480. https://doi.org/10.1146/annurev-ento-011019-025151 (2020).Article 
    CAS 
    PubMed 

    Google Scholar 
    Goulson, D. The insect apocalypse, and why it matters. Curr. Biol. 29, R967–R971. https://doi.org/10.1016/j.cub.2019.06.069 (2019).Article 
    CAS 
    PubMed 

    Google Scholar 
    Cardoso, P. et al. Scientists’ warning to humanity on insect extinctions. Biol. Conserv. 242, 108426. https://doi.org/10.1016/j.biocon.2020.108426 (2020).Article 

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

    Google Scholar 
    Ollerton, J. Pollinator diversity: Distribution, ecological function, and conservation. Annu. Rev. Ecol. Evol. Syst. 48, 353–376. https://doi.org/10.1146/annurev-ecolsys-110316-022919 (2017).Article 

    Google Scholar 
    Klein, A.-M. et al. Importance of pollinators in changing landscapes for world crops. Proc. Biol. Sci. 274, 303–313. https://doi.org/10.1098/rspb.2006.3721 (2007).Article 
    PubMed 

    Google Scholar 
    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 
    Ollerton, J., Erenler, H., Edwards, M. & Crockett, R. Pollinator declines. Extinctions of aculeate pollinators in Britain and the role of large-scale agricultural changes. Science 346, 1360–1362. https://doi.org/10.1126/science.1257259 (2014).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Wenzel, A., Grass, I., Belavadi, V. V. & Tscharntke, T. How urbanization is driving pollinator diversity and pollination—A systematic review. Biol. Conserv. 241, 108321. https://doi.org/10.1016/j.biocon.2019.108321 (2020).Article 

    Google Scholar 
    Senapathi, D., Goddard, M. A., Kunin, W. E. & Baldock, K. C. R. Landscape impacts on pollinator communities in temperate systems: Evidence and knowledge gaps. Funct. Ecol. 31, 26–37. https://doi.org/10.1111/1365-2435.12809 (2017).Article 

    Google Scholar 
    Fenoglio, M. S., Rossetti, M. R. & Videla, M. Negative effects of urbanization on terrestrial arthropod communities: A meta-analysis. Glob. Ecol. Biogeogr. 29, 1412–1429. https://doi.org/10.1111/geb.13107 (2020).Article 

    Google Scholar 
    Ives, C. D. et al. Cities are hotspots for threatened species. Glob. Ecol. Biogeogr. 25, 117–126. https://doi.org/10.1111/geb.12404 (2016).Article 

    Google Scholar 
    Soanes, K. & Lentini, P. E. When cities are the last chance for saving species. Front. Ecol. Evol. 17, 225–231. https://doi.org/10.1002/fee.2032 (2019).Article 

    Google Scholar 
    Lynch, L. et al. Changes in land use and land cover along an urban-rural gradient influence floral resource availability. Curr. Landsc. Ecol. Rep. 6, 46–70. https://doi.org/10.1007/s40823-021-00064-1 (2021).Article 

    Google Scholar 
    Hall, D. M. et al. The city as a refuge for insect pollinators. Conserv. Biol. 31, 24–29. https://doi.org/10.1111/cobi.12840 (2017).Article 
    PubMed 

    Google Scholar 
    Buchholz, S. & Egerer, M. H. Functional ecology of wild bees in cities: Towards a better understanding of trait-urbanization relationships. Biodivers. Conserv. 29, 2779–2801. https://doi.org/10.1007/s10531-020-02003-8 (2020).Article 

    Google Scholar 
    Theodorou, P. et al. Urban areas as hotspots for bees and pollination but not a panacea for all insects. Nat. Commun. 11, 576. https://doi.org/10.1038/s41467-020-14496-6 (2020).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Khalifa, S. A. M. et al. Overview of bee pollination and its economic value for crop production. Insects https://doi.org/10.3390/insects12080688 (2021).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Doyle, T. et al. Pollination by hoverflies in the Anthropocene. Proc. Biol. Sci. 287, 20200508. https://doi.org/10.1098/rspb.2020.0508 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Rader, R. et al. Non-bee insects are important contributors to global crop pollination. Proc. Natl. Acad. Sci. USA. 113, 146–151. https://doi.org/10.1073/pnas.1517092112 (2016).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Persson, A. S., Ekroos, J., Olsson, P. & Smith, H. G. Wild bees and hoverflies respond differently to urbanisation, human population density and urban form. Landsc. Urban Plan. 204, 103901. https://doi.org/10.1016/j.landurbplan.2020.103901 (2020).Article 

    Google Scholar 
    Gathof, A. K., Grossmann, A. J., Herrmann, J. & Buchholz, S. Who can pass the urban filter? A multi-taxon approach to disentangle pollinator trait-environmental relationships. Oecologia 199, 165–179. https://doi.org/10.1007/s00442-022-05174-z (2022).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Baldock, K. C. R. et al. Where is the UK’s pollinator biodiversity? The importance of urban areas for flower-visiting insects. Proc. Biol. Sci. 282, 20142849. https://doi.org/10.1098/rspb.2014.2849 (2015).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Ramírez-Restrepo, L. & MacGregor-Fors, I. Butterflies in the city: A review of urban diurnal Lepidoptera. Urban Ecosyst. 20, 171–182. https://doi.org/10.1007/s11252-016-0579-4 (2017).Article 

    Google Scholar 
    Kuussaari, M. et al. Butterfly species’ responses to urbanization: Differing effects of human population density and built-up area. Urban Ecosyst. 24, 515–527. https://doi.org/10.1007/s11252-020-01055-6 (2020).Article 

    Google Scholar 
    Theodorou, P. The effects of urbanisation on ecological interactions. Curr. Opin. Insect. Sci. 52, 100922. https://doi.org/10.1016/j.cois.2022.100922 (2022).Article 
    PubMed 

    Google Scholar 
    Martins, K. T., Gonzalez, A. & Lechowicz, M. J. Patterns of pollinator turnover and increasing diversity associated with urban habitats. Urban Ecosyst. 20, 1359–1371. https://doi.org/10.1007/s11252-017-0688-8 (2017).Article 

    Google Scholar 
    Theodorou, P. et al. The structure of flower visitor networks in relation to pollination across an agricultural to urban gradient. Funct. Ecol. 31, 838–847. https://doi.org/10.1111/1365-2435.12803 (2017).Article 

    Google Scholar 
    Geslin, B., Gauzens, B., Thébault, E. & Dajoz, I. Plant pollinator networks along a gradient of urbanisation. PLoS ONE 8, e63421. https://doi.org/10.1371/journal.pone.0063421 (2013).Article 
    ADS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Udy, K. L., Reininghaus, H., Scherber, C. & Tscharntke, T. Plant–pollinator interactions along an urbanization gradient from cities and villages to farmland landscapes. Ecosphere https://doi.org/10.1002/ecs2.3020 (2020).Article 

    Google Scholar 
    Jędrzejewska-Szmek, K. & Zych, M. Flower-visitor and pollen transport networks in a large city: Structure and properties. Arthropod. Plant Interact. 7, 503–516. https://doi.org/10.1007/s11829-013-9274-z (2013).Article 

    Google Scholar 
    von der Lippe, M., Buchholz, S., Hiller, A., Seitz, B. & Kowarik, I. CityScapeLab Berlin: A research platform for untangling urbanization effects on biodiversity. Sustainability 12, 2565. https://doi.org/10.3390/su12062565 (2020).Article 

    Google Scholar 
    Dylewski, Ł, Maćkowiak, Ł & Banaszak-Cibicka, W. Are all urban green spaces a favourable habitat for pollinator communities? Bees, butterflies and hoverflies in different urban green areas. Ecol. Entomol. 44, 678–689. https://doi.org/10.1111/een.12744 (2019).Article 

    Google Scholar 
    Grossmann, A. J., Herrmann, J., Buchholz, S. & Gathof, A. K. Dry grassland within the urban matrix acts as favourable habitat for different pollinators including endangered species. Insect Conserv. Divers. https://doi.org/10.1111/icad.12607 (2022).Article 

    Google Scholar 
    Settele, J., Steiner, R., Feldmann, R. & Hermann, G. Schmetterlinge. Die Tagfalter Deutschlands: 720 Farbfotos. 3rd ed. (2015).Amiet, F. Hymenoptera Apidae, 1. Teil. Allgemeiner Teil, Gattungsschlüssel – Die Gattungen Apis, Bombus und Psithyrus (Centre Suisse de Cartographie de la Faune, 1996).
    Google Scholar 
    Amiet, F., Müller, A. & Neumeyer, R. Apidae 2. Colletes, Dufourea, Hylaeus, Nomia, Nomioides, Rhophitoides, Rophites, Sphecodes, Systropha (Fauna Helvetica, 1999).
    Google Scholar 
    Amiet, F., Herrmann, M., Müller, A. & Neumeyer, R. Apidae 3. Halictus, Lasioglossum (Centre Suisse de Cartographie de la Faune, 2001).
    Google Scholar 
    Amiet, F., Herrmann, M., Müller, A. & Neumeyer, R. Apidae 4. Anthidium, Chelostoma, Coelioxys, Dioxys, Heriades, Lithurgus, Megachile, Osmia, Stelis (Centre Suisse de Cartographie de la Faune, 2004).
    Google Scholar 
    Amiet, F., Herrmann, M., Müller, A. & Neumeyer, R. Apidae 5. Ammobates, Ammobatoides, Anthophora, Biastes, Ceratina, Dasypoda, Epeoloides, Epeolus, Eucera, Macropis, Melecta, Melitta, Nomada, Pasites, Tetralonia, Thyreus, Xylocopa (Centre Suisse de Cartographie de la Faune, 2007).
    Google Scholar 
    Amiet, F., Herrmann, M., Müller, A. & Neumeyer, R. Apidae 6. Andrena, Melliturga, Panurginus, Panurgus (Centre Suisse de Cartographie de la Faune, 2010).
    Google Scholar 
    Gokcezade, J. F., Gereben-Krenn, B.-A., Neumayer, J. & Krenn, H. W. Feldbestimmungsschlüssel für die Hummeln Österreichs, Deutschlands und der Schweiz (Hymenoptera, Apidae). Linzer biologische Beiträge 47, 5–42 (2015).
    Google Scholar 
    Bartsch, H. Tvåvingar: Blomflugor. Diptera: Syrphidae: Syrphinae: denna volym omfattar samtliga nordiska arter (ArtDatabanken Sveriges lantbruksuniversitet, 2009).
    Google Scholar 
    Bartsch, H. Tvåvingar: Blomflugor. Diptera: Syrphidae: Eristalinae & Microdontinae: denna volym omfattar samtliga nordiska arter (ArtDatabanken Sveriges lantbruksuniversitet, 2009).
    Google Scholar 
    Bot, S. & van de Meutter, F. Veldgids zweefvliegen (KNNV Uitgeverij, 2019).
    Google Scholar 
    Jäger, E. J. Rothmaler-Exkursionsflora von Deutschland. Gefäßpflanzen: Grundband 20th edn. (Springer Spektrum, 2011).
    Google Scholar 
    Senate Department for Urban Development and Housing. Berlin Environmental Atlas. 06.01 Actual Use of Built-up Areas/06.02 Inventory of Green and Open Spaces 2010 (2011).Holland, J. D., Bert, D. G. & Fahrig, L. Determining the spatial scale of species’ response to habitat. Bioscience 54, 227. https://doi.org/10.1641/0006-3568(2004)054[0227:DTSSOS]2.0.CO;2 (2004).Article 

    Google Scholar 
    Senate Department for Urban Development and Housing. Berlin Environmental Atlas. 05.08 Biotope Types (2014).Hanski, I. A practical model of metapopulation dynamics. J. Anim. Ecol. 63, 151. https://doi.org/10.2307/5591 (1994).Article 

    Google Scholar 
    Hanski, I. Habitat connectivity, habitat continuity, and metapopulations in dynamic landscapes. Oikos 87, 209. https://doi.org/10.2307/3546736 (1999).Article 

    Google Scholar 
    Senate Department for Urban Development and Housing. Berlin Environmental Atlas. 06.10 Building and Vegetation Heights (2014).Saura, S. & Torné, J. Conefor Sensinode 2.2: A software package for quantifying the importance of habitat patches for landscape connectivity. Environ. Model. Softw. 24, 135–139. https://doi.org/10.1016/j.envsoft.2008.05.005 (2009).Article 

    Google Scholar 
    Saure, C. Rote Liste und Gesamtartenliste der Bienen und Wespen (Hymenoptera part.) von Berlin mit Angaben zu den Ameisen. In Rote Listen der gefährdeten Pflanzen und Tiere von Berlin.Speight, M. C. D. Species Accounts of European Syrphidae (Diptera) (Syrph the Net Publications, 2014).
    Google Scholar 
    Middleton-Welling, J. et al. A new comprehensive trait database of European and Maghreb butterflies, Papilionoidea. Sci. Data 7, 351. https://doi.org/10.1038/s41597-020-00697-7 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Dormann, C. F., Fründ, J., Blüthgen, N. & Gruber, B. Indices, graphs and null models: Analyzing bipartite ecological networks. Open Ecol. J. 2, 7–24. https://doi.org/10.2174/1874213000902010007 (2009).Article 

    Google Scholar 
    Kaiser-Bunbury, C. N. & Blüthgen, N. Integrating network ecology with applied conservation: A synthesis and guide to implementation. AoB Plants https://doi.org/10.1093/aobpla/plv076 (2015).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Almeida-Neto, M., Guimarães, P., Guimarães, P. R., Loyola, R. D. & Ulrich, W. A consistent metric for nestedness analysis in ecological systems: Reconciling concept and measurement. Oikos 117, 1227–1239. https://doi.org/10.1111/J.0030-1299.2008.16644.X (2008).Article 

    Google Scholar 
    Dormann, C. F. & Strauss, R. A method for detecting modules in quantitative bipartite networks. Methods Ecol. Evol. 5, 90–98. https://doi.org/10.1111/2041-210X.12139 (2014).Article 

    Google Scholar 
    Blüthgen, N., Menzel, F. & Blüthgen, N. Measuring specialization in species interaction networks. BMC Ecol. 6, 9. https://doi.org/10.1186/1472-6785-6-9 (2006).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Patefield, W. M. Algorithm AS 159: An efficient method of generating random R × C tables with given row and column totals. J. Appl. Stat. 30, 91. https://doi.org/10.2307/2346669 (1981).Article 
    MATH 

    Google Scholar 
    Stein, K. et al. Plant–pollinator networks in Savannas of Burkina Faso, West Africa. Diversity 13, 1. https://doi.org/10.3390/d13010001 (2021).Article 
    ADS 

    Google Scholar 
    Escobedo-Kenefic, N. et al. Disentangling the effects of local resources, landscape heterogeneity and climatic seasonality on bee diversity and plant–pollinator networks in tropical highlands. Oecologia 194, 333–344. https://doi.org/10.1007/s00442-020-04715-8 (2020).Article 
    ADS 
    PubMed 

    Google Scholar 
    Renaud, E., Baudry, E. & Bessa-Gomes, C. Influence of taxonomic resolution on mutualistic network properties. Ecol. Evol. 10, 3248–3259. https://doi.org/10.1002/ece3.6060 (2020).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    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, e0222316. https://doi.org/10.1371/journal.pone.0222316 (2019).Article 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Egerer, M. & Kowarik, I. Confronting the modern gordian knot of urban beekeeping. Trends Ecol. Evol. 35, 956–959. https://doi.org/10.1016/j.tree.2020.07.012 (2020).Article 
    PubMed 

    Google Scholar 
    Zuur, A. F., Ieono, E. N., Walker, N., Saveliev, A. A. & Smith, G. M. Mixed Effects Models and Extensions in Ecology with R (Springer, 2009).Book 
    MATH 

    Google Scholar 
    Bartón, K. MuMIn. multi-model inference, R package version 1.42.1 (2018).Paradis, E., Claude, J. & Strimmer, K. APE: Analyses of phylogenetics and evolution in R language. Bioinformatics 20, 289–290. https://doi.org/10.1093/bioinformatics/btg412 (2004).Article 
    CAS 
    PubMed 

    Google Scholar 
    Wood, T. J., Kaplan, I. & Szendrei, Z. Wild bee pollen diets reveal patterns of seasonal foraging resources for honey bees. Front. Ecol. Evol. https://doi.org/10.3389/fevo.2018.00210 (2018).Article 

    Google Scholar 
    Proske, A., Lokatis, S. & Rolff, J. Impact of mowing frequency on arthropod abundance and diversity in urban habitats: A meta-analysis. Urban For Urban Green 76, 127714. https://doi.org/10.1016/j.ufug.2022.127714 (2022).Article 

    Google Scholar 
    Bates, A. J. et al. Changing bee and hoverfly pollinator assemblages along an urban-rural gradient. PLoS ONE 6, e23459. https://doi.org/10.1371/journal.pone.0023459 (2011).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Geslin, B. et al. The proportion of impervious surfaces at the landscape scale structures wild bee assemblages in a densely populated region. Ecol. Evol. 6, 6599–6615. https://doi.org/10.1002/ece3.2374 (2016).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Birdshire, K. R., Carper, A. L. & Briles, C. E. Bee community response to local and landscape factors along an urban-rural gradient. Urban Ecosyst. 23, 689–702. https://doi.org/10.1007/s11252-020-00956-w (2020).Article 

    Google Scholar 
    Goddard, M. A., Benton, T. G. & Dougill, A. J. Beyond the garden fence: Landscape ecology of cities. Trends Ecol. Evol. 25, 202–203. https://doi.org/10.1016/j.tree.2009.12.007 (2010).Article 

    Google Scholar 
    Theodorou, P. et al. Bumble bee colony health and performance vary widely across the urban ecosystem. J. Anim. Ecol. 91, 2135–2148. https://doi.org/10.1111/1365-2656.13797 (2022).Article 
    PubMed 

    Google Scholar 
    Potts, S. G., Vulliamy, B., Dafni, A., Ne’eman, G. & Willmer, P. Linking bees and flowers: How do floral communities structure pollinator communities?. Ecology 84, 2628–2642. https://doi.org/10.1890/02-0136 (2003).Article 

    Google Scholar 
    Ebeling, A., Klein, A.-M., Schumacher, J., Weisser, W. W. & Tscharntke, T. How does plant richness affect pollinator richness and temporal stability of flower visits?. Oikos 117, 1808–1815. https://doi.org/10.1111/j.1600-0706.2008.16819.x (2008).Article 

    Google Scholar 
    Theodorou, P. et al. Urban fragmentation leads to lower floral diversity, with knock-on impacts on bee biodiversity. Sci. Rep. 10, 21756. https://doi.org/10.1038/s41598-020-78736-x (2020).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Potts, S. G. et al. Role of nesting resources in organising diverse bee communities in a Mediterranean landscape. Ecol. Entomol. 30, 78–85. https://doi.org/10.1111/j.0307-6946.2005.00662.x (2005).Article 

    Google Scholar 
    Fründ, J., Linsenmair, K. E. & Blüthgen, N. Pollinator diversity and specialization in relation to flower diversity. Oikos 119, 1581–1590. https://doi.org/10.1111/j.1600-0706.2010.18450.x (2010).Article 

    Google Scholar 
    Fornoff, F. et al. Functional flower traits and their diversity drive pollinator visitation. Oikos 126, 1020–1030. https://doi.org/10.1111/oik.03869 (2017).Article 
    CAS 

    Google Scholar 
    Hofmann, M. M. & Renner, S. S. One-year-old flower strips already support a quarter of a city’s bee species. J. Hymenopt. Res. 75, 87–95. https://doi.org/10.3897/jhr.75.47507 (2020).Article 

    Google Scholar 
    Verboven, H. A., Uyttenbroeck, R., Brys, R. & Hermy, M. Different responses of bees and hoverflies to land use in an urban–rural gradient show the importance of the nature of the rural land use. Landsc. Urban Plan. 126, 31–41. https://doi.org/10.1016/j.landurbplan.2014.02.017 (2014).Article 

    Google Scholar 
    Luder, K., Knop, E. & Menz, M. H. M. Contrasting responses in community structure and phenology of migratory and non-migratory pollinators to urbanization. Divers. Distrib. 24, 919–927. https://doi.org/10.1111/ddi.12735 (2018).Article 

    Google Scholar 
    Merckx, T. & van Dyck, H. Urbanization-driven homogenization is more pronounced and happens at wider spatial scales in nocturnal and mobile flying insects. Glob. Ecol. Biogeogr. 28, 1440–1455. https://doi.org/10.1111/geb.12969 (2019).Article 

    Google Scholar 
    Tzortzakaki, O., Kati, V., Panitsa, M., Tzanatos, E. & Giokas, S. Butterfly diversity along the urbanization gradient in a densely-built Mediterranean city: Land cover is more decisive than resources in structuring communities. Landsc. Urban Plan. 183, 79–87. https://doi.org/10.1016/j.landurbplan.2018.11.007 (2019).Article 

    Google Scholar 
    Krauss, J., Steffan-Dewenter, I. & Tscharntke, T. How does landscape context contribute to effects of habitat fragmentation on diversity and population density of butterflies?. J. Biogeogr. 30, 889–900. https://doi.org/10.1046/j.1365-2699.2003.00878.x (2003).Article 

    Google Scholar 
    Cozzi, G., Müller, C. B. & Krauss, J. How do local habitat management and landscape structure at different spatial scales affect fritillary butterfly distribution on fragmented wetlands?. Landsc. Ecol. 23, 269–283. https://doi.org/10.1007/s10980-007-9178-3 (2008).Article 

    Google Scholar 
    He, M. et al. Effects of landscape and local factors on the diversity of flower-visitor groups under an urbanization gradient, a case study in Wuhan, China. Diversity 14, 208. https://doi.org/10.3390/d14030208 (2022).Article 

    Google Scholar 
    Buchholz, S., Gathof, A. K., Grossmann, A. J., Kowarik, I. & Fischer, L. K. Wild bees in urban grasslands: Urbanisation, functional diversity and species traits. Landsc. Urban Plan. 196, 103731. https://doi.org/10.1016/j.landurbplan.2019.103731 (2020).Article 

    Google Scholar 
    Chapman, R. E. & Bourke, A. F. G. The influence of sociality on the conservation biology of social insects. Ecol. Lett. 4, 650–662. https://doi.org/10.1046/j.1461-0248.2001.00253.x (2001).Article 

    Google Scholar 
    Gaertner, M. et al. Non-native species in urban environments: Patterns, processes, impacts and challenges. Biol. Invasions 19, 3461–3469. https://doi.org/10.1007/s10530-017-1598-7 (2017).Article 

    Google Scholar 
    Kowarik, I. On the role of alien species in urban flora and vegetation. In Urban Ecology. An International Perspective on the Interaction Between Humans and Nature (ed. Marzluff, J. M.) 321–338 (2008).Lorenz, S. & Stark, K. Saving the honeybees in Berlin? A case study of the urban beekeeping boom. Environ. Sociol. 1, 116–126. https://doi.org/10.1080/23251042.2015.1008383 (2015).Article 

    Google Scholar 
    Olesen, J. M., Bascompte, J., Dupont, Y. L. & Jordano, P. The modularity of pollination networks. Proc. Natl. Acad. Sci. USA 104, 19891–19896. https://doi.org/10.1073/pnas.0706375104 (2007).Article 
    ADS 
    PubMed 
    PubMed Central 
    MATH 

    Google Scholar 
    Thébault, E. & Fontaine, C. Stability of ecological communities and the architecture of mutualistic and trophic networks. Science 329, 853–856. https://doi.org/10.1126/science.1188321 (2010).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Dormann, C. F., Fründ, J. & Schaefer, H. M. Identifying causes of patterns in ecological networks: Opportunities and limitations. Annu. Rev. Ecol. Evol. Syst. 48, 559–584. https://doi.org/10.1146/annurev-ecolsys-110316-022928 (2017).Article 

    Google Scholar 
    Tylianakis, J. M., Laliberté, E., Nielsen, A. & Bascompte, J. Conservation of species interaction networks. Biol. Conserv. 143, 2270–2279. https://doi.org/10.1016/j.biocon.2009.12.004 (2010).Article 

    Google Scholar 
    Grilli, J., Rogers, T. & Allesina, S. Modularity and stability in ecological communities. Nat. Commun. 7, 12031. https://doi.org/10.1038/ncomms12031 (2016).Article 
    ADS 
    CAS 
    PubMed 
    PubMed Central 

    Google Scholar 
    Grass, I., Jauker, B., Steffan-Dewenter, I., Tscharntke, T. & Jauker, F. Past and potential future effects of habitat fragmentation on structure and stability of plant–pollinator and host-parasitoid networks. Nat. Ecol. Evol 2, 1408–1417. https://doi.org/10.1038/s41559-018-0631-2 (2018).Article 
    PubMed 

    Google Scholar 
    Kaiser-Bunbury, C. N. et al. Ecosystem restoration strengthens pollination network resilience and function. Nature 542, 223–227. https://doi.org/10.1038/nature21071 (2017).Article 
    ADS 
    CAS 
    PubMed 

    Google Scholar 
    Bommarco, R. et al. Dispersal capacity and diet breadth modify the response of wild bees to habitat loss. Proc. Biol. Sci. 277, 2075–2082. https://doi.org/10.1098/rspb.2009.2221 (2010).Article 
    PubMed 
    PubMed Central 

    Google Scholar 
    Alarcón, R., Waser, N. M. & Ollerton, J. Year-to-year variation in the topology of a plant–pollinator interaction network. Oikos 117, 1796–1807. https://doi.org/10.1111/j.0030-1299.2008.16987.x (2008).Article 

    Google Scholar 
    Dupont, Y. L., Padrón, B., Olesen, J. M. & Petanidou, T. Spatio-temporal variation in the structure of pollination networks. Oikos 118, 1261–1269. https://doi.org/10.1111/j.1600-0706.2009.17594.x (2009).Article 

    Google Scholar 
    Santamaría, S. et al. Landscape effects on pollination networks in Mediterranean gypsum islands. Plant Biol. 20(Suppl 1), 184–194. https://doi.org/10.1111/plb.12602 (2018).Article 
    PubMed 

    Google Scholar  More

  • in

    Balancing the bloom

    Algal blooms that form because of phytoplankton proliferation have key roles in marine ecology and carbon fixation. When the blooms die, most of the fixed carbon is transferred to higher trophic levels, and a small fraction sinks into the deep sea. Viral infection is one of the causes of bloom termination, but its effect on the fate and flow of carbon in the ocean is unknown. In this study, Vincent et al. perform a mesocosm experiment to analyse the bloom dynamics of the coccolithophore microalga Emiliania huxleyi and the impact of viral infection on surrounding bacterial communities and the carbon cycle. The authors observed that viral infection was not only the main cause of phytoplankton mortality, but it also shaped the composition of free-living bacterial and eukaryotic species in the blooms. On viral infection of E. huxleyi, the authors found a comparable biomass of eukaryotic and bacterial heterotrophic recyclers, as well as increased organic and inorganic carbon release that contributed to carbon sinking into the deep ocean. Altogether, these results highlight the impact of viruses on the microbial communities of blooms and the consequences on carbon cycling. More

  • in

    Fungi feed bacteria for biodegradation

    The pesticide hexachlorocyclohexane (HCH) is a toxic and persistent contaminant in the environment. Some bacteria and fungi can degrade HCH and its isomers under laboratory conditions. However, in heterogeneous environments, where many different factors are at play, the biodegradation capacity is challenged by the availability of nutrients to support degraders’ growth. As opposed to bacteria, fungi are more adapted to heterogeneous habitats, and in some cases mycelial fungi can facilitate the transport of organic substrates throughout the mycosphere, increasing their availability to promote bacterial contaminant biodegradation. However, how this occurs is not entirely understood. In this study, Khan et al. demonstrate that mycelial nutrients transferred from nutrient-rich to nutrient-deprived habitats promote co-metabolic degradation of HCH by bacteria. The authors incubated a non-HCH-degrading fungus (Fusarium equiseti K3) and a co-metabolically HCH-degrading bacterium (Sphingobium sp. S8) in a structured model ecosystem. Results from 13C isotope labelling and metaproteomics showed that fungal 13C was incorporated into bacterial proteins responsible for HCH degradation, thus illustrating the importance of synergistic fungal–bacterial interactions for contaminant biodegradation in nutrient-poor environments. More

  • in

    Impending anthropogenic threats and protected area prioritization for jaguars in the Brazilian Amazon

    Estes, J. et al. Trophic downgrading of planet Earth. Science 333, 301–306 (2011).Article 
    CAS 

    Google Scholar 
    Ripple, W. J. et al. Status and ecological effects of the World’s Largest Carnivores. Science 343, 151–162 (2014).Article 
    CAS 

    Google Scholar 
    Cardillo, M. et al. Multiple causes of high extinction risk in large mammal species. Science 309, 1239–1241 (2005).Article 
    CAS 

    Google Scholar 
    De La Torre, J. A., González-Maya, J. F., Zarza, H., Ceballos, G. & Medellín, R. A. The jaguar’s spots are darker than they appear: assessing the global conservation status of the jaguar (Panthera onca). Oryx 52, 300–315 (2018).Article 

    Google Scholar 
    Lindsey, P. A. et al. The performance of African protected areas for lions and their prey. Biol. Conserv. 209, 137–149 (2017).Article 

    Google Scholar 
    Carbone, C., Cowlishaw, G., Isaac, N. J. B. & Rowcliffe, J. M. How far do animals go? Determinants of day range in mammals. Am. Nat. 165, 290–297 (2005).Article 

    Google Scholar 
    Sanderson, E. W. et al. Planning to save a species: the jaguar as a model. Conserv. Biol. 16, 58–72 (2002).Article 

    Google Scholar 
    Rabinowitz, A. & Zeller, K. A. A range-wide model of landscape connectivity and conservation for the jaguar, Panthera onca. Biol. Conserv. 143, 939–945 (2010).Article 

    Google Scholar 
    Woodroffe, R. Predators and people: using human densities to interpret declines of large carnivores. Anim. Conserv. 3, 165–173 (2000).Article 

    Google Scholar 
    Crooks, K. R. Relative sensitivities of mammalian carnivores to habitat fragmentation. Conserv. Biol. 16, 488–502 (2002).Article 

    Google Scholar 
    Ferreira, A. S., Peres, C. A., Bogoni, J. A. & Cassano, C. G. Use of agroecosystem matrix habitats by mammalian carnivores (Carnivora): a global-scale analysis. Mammal Rev. https://doi.org/10.1111/mam12137 (2018).Thompson, J. J. et al. Range-wide factors shaping space use and movements by the Neotropic’s flagship predator: the jaguar. Curr. Biol. https://doi.org/10.1016/jcub202106029 (2021).Sunquist, M. & Sunquist, F. Wild Cats of the World. University of Chicago Press (2002).Leader-Williams, N. & Dublin, H. T. in Priorities for The Conservation Of Mammalian Diversity: Has The Panda Had Its Day? (eds. Entwistle, A., Dunstone, N.) 53−81 (Cambridge University Press, 2000).Thornton, D. et al. Assessing the umbrella value of a range-wide conservation network for jaguars (Panthera onca). Ecol. Appl. 26, 1112–1124 (2015).Article 

    Google Scholar 
    Olsoy, P. J. et al. Quantifying the effects of deforestation and fragmentation on a range-wide conservation plan for jaguars. Biol. Conserv. 203, 8–16 (2016).Article 

    Google Scholar 
    Morato, R. G., Beisiegel, B. M., Ramalho, E. E. & Boulhosa, R. L. P. Avaliação do risco de extinção da Onça-pintada Panthera onca (Linnaeus, 1758) no Brasil. Biodivers. Brasil. 3, 122–132 (2013).
    Google Scholar 
    Hunter, L. Carnivores of the World. Princeton Univ Press (2011).Morato, R. G. et al. Space use and movement of a neotropical top predator: The Endangered Jaguar. PLoS ONE 11, e0168176 (2016).Article 

    Google Scholar 
    Eriksson, C. E. et al. Extensive aquatic subsidies lead to territorial breakdown and high density of an apex predator. Ecology 103, e03543 (2022).Article 

    Google Scholar 
    Chapman, B. et al. in Animal Movement Across Scales 1st edn. (eds. Hansson, L-A, Akesson, S.) 11–30 (Oxford University Press, 2014).Quigley, H. et al. Panthera onca. (errata version published in 2018). The IUCN Red List of Threatened Species 2017:e.T15953A123791436. https://doi.org/10.2305/IUCN.UK.2017-3.RLTS.T15953A50658693.en (2017).Paviolo, A. et al. A biodiversity hotspot losing its top predator: the challenge of jaguar conservation in the Atlantic Forest of South America. Sci. Rep. 6, 37147 (2016).Article 
    CAS 

    Google Scholar 
    Tobler, M. W., Carillo-Perscastegui, S. E., Hartley, A. Z. & Powell, G. V. N. High jaguar densities and large population sizes in the core habitat of the southwestern Amazon. Biol. Conserv. 159, 375–381 (2013).Article 

    Google Scholar 
    Jędrzejewski, W. et al. Estimating large carnivore populations at global scale based on spatial predictions of density and distribution: application to the jaguar (Panthera onca). PLoS ONE 13, e0194719 (2018).Article 

    Google Scholar 
    Eva, H. D. et al. A proposal for defining the geographical boundaries of Amazonia; synthesis of the results from an expert consultation workshop organized by the European Commission in collaboration with the Amazon Cooperation Treaty Organization-JRC Ispra (No 21808-EN). https://core.ac.uk/download/pdf/38630683.pdf (2005).Nepstad, D. C., Stickler, C. M., Soares-Filho, B., Merry, F. & Nin, E. Interactions among Amazon land use, forests and climate: prospects for a near-term forest tipping point. Philos. Trans. R. Soc. B 363, 1737–1746 (2008).Article 

    Google Scholar 
    Marques, A. A. B., Schneider, M. & Peres, C. A. Human population and socioeconomic modulators of conservation performance in 788 Amazonian and Atlantic Forest reserves. PeerJ 4, pe2206 (2016).Article 

    Google Scholar 
    Jaguar 2030 Roadmap. Regional plan to save America’s largest cat and its ecosystems. https://www.internationaljaguarday.org/jaguar-conservation-roadmap (2018).Sanderson, E. W. et al. A systematic review of potential habitat suitability for the jaguar Panthera onca in central Arizona and New Mexico, USA. Oryx 2021, 1–12 (2021).
    Google Scholar 
    Simberloff, D. Flagships, umbrellas, and keystones: is single-species management passe’ in the landscape era. Biol. Conserv. 83, 247–57 (1998).Article 

    Google Scholar 
    Silvério, D. V. et al. Testing the Amazon savannization hypothesis: fire effects on invasion of a neotropical forest by native Cerrado and exotic pasture grasses. Philos. Trans. R. Soc. B 368, 20120427 (2013).Article 

    Google Scholar 
    Brazil’s National Institute for Space Research (INPE). Banco de dados de Queimadas INPE—Programa Queimadas. http://queimadasdgiinpebr/queimadas/bdqueimadas (2020b).Silva-Jr, C. H. L. et al. The Brazilian Amazon deforestation rate in 2020 is the greatest of the decade. Nat. Ecol. Evol. 5, 144–145 (2020).Article 

    Google Scholar 
    Walker, R. et al. Protecting the Amazon with protected areas. Proc. Natl Acad. Sci. USA 106, 10582–10586 (2009).Article 
    CAS 

    Google Scholar 
    Gray, C. L. et al. Local biodiversity is higher inside than outside terrestrial protected areas worldwide. Nat. Commun. 7, 12306 (2016).Article 
    CAS 

    Google Scholar 
    Begotti, R. A. & Peres, C. A. Rapidly escalating threats to the biodiversity and ethnocultural capital of Brazilian Indigenous Lands. Land Use Policy 96, 104694 (2020).Article 

    Google Scholar 
    Walker, W. S. et al. The role of forest conversion, degradation, and disturbance in the carbon dynamics of Amazon indigenous territories and protected areas. Proc. Natl Acad. Sci. USA 117, 3015–3025 (2020).Article 
    CAS 

    Google Scholar 
    Moilanen, A., Arponen, A., Stokland, J. N. & Cabeza, M. Assessing replacement cost of conservation areas: How does habitat loss influence priorities? Biol. Conserv. 142, 575–585 (2009).Article 

    Google Scholar 
    Almeida-Rocha, J. A. & Peres, C. A. Nominally protected buffer zones around tropical protected areas are as highly degraded as the wider unprotected countryside. Biol. Conserv. 256, 109068 (2021).Article 

    Google Scholar 
    Terborgh, J. The role of felid predators in Neotropical Forests. Vida Silv. Neotrop. 2, 3–5 (1990).
    Google Scholar 
    Woodroffe, R. & Ginsberg, J. R. Edge effects and the extinction of populations inside protected areas. Science 280, 2126–2128 (1998).Article 
    CAS 

    Google Scholar 
    Brando, P. M. et al. The gathering firestorm in southern Amazonia. Sci. Adv. 6, 1632 (2020).Convention on the Conservation of Migratory Species of Wild Animals (CMS). Proposal for the Inclusion of the Jaguar in Appendices I and II of the Convention. https://www.cms.int/en/document/proposal-inclusion-jaguar-appendices-i-and-ii-convention (2022).Ceddia, M. G., Bardsley, N. O., Gomez-y-Paloma, S. & Sedlacek, S. Governance, agricultural intensification, and land sparing in tropical South America. Proc. Natl Acad. Sci. USA 111, 7242–7247 (2014).Laurance, W. F. et al. Impacts of roads and hunting on central African rainforest mammals. Conserv. Biol. 20, 1251–1261 (2006).Article 

    Google Scholar 
    Brancalion, P. H. S. et al. Análise crítica da Lei de Proteção da Vegetação Nativa (2012), que substituiu o antigo Código Florestal: atualizações e ações em curso. Natureza Conservação 14, 1–16 (2016).Wilkie, D. S., Bennett, E. L., Peres, C. A. & Cunningham, A. A. The empty forest revisited. Ann. N. Y. Acad. Sci. 1223, 120–128 (2011).Article 

    Google Scholar 
    Bogoni, J. A., Peres, C. A. & Ferraz, K. M. P. M. B. Extent, intensity and drivers of mammal defaunation: a continental-scale analysis across the Neotropics. Sci. Rep. 10, 14750 (2020).Article 
    CAS 

    Google Scholar 
    Ferrante, L. & Fearnside, P. M. Brazil’s new president and ‘ruralists’ threaten Amazonia’s environment, traditional peoples and the global climate. Environ. Conserv. 46, 261–263 (2019).Article 

    Google Scholar 
    Aragão, L. E. O. C. & Shimabukuro, Y. E. The incidence of fire in Amazonian forests with implications for REDD. Science 328, 1275–1278 (2010).Article 

    Google Scholar 
    Barlow, J. & Peres, C. A. Fire-mediated dieback and compositional cascade in an Amazonian forest. Philos. Trans. R. Soc. B 363, 1787 (2008).Article 

    Google Scholar 
    Michalski, F., Boulhosa, R. L. P., Faria, A. & Peres, C. A. Human–wildlife conflicts in a fragmented Amazonian forest landscape: determinants of large felid depredation on livestock. Anim. Conserv. https://doi.org/10.1111/j1469-1795200600025x (2006).Article 

    Google Scholar 
    Jorge, M. L. S. P., Galetti, M., Ribeiro, M. C. & Ferraz, K. M. P. M. B. Mammal defaunation as surrogate of trophic cascades in a biodiversity hotspot. Biol. Conserv. 163, 49–57 (2013).Article 

    Google Scholar 
    Menezes, J. F. S., Tortato, F. R., Roque, F. O., Oliveira-Santos, L. G. & Morato, R. G. Deforestation, fires, and lack of governance are displacing thousands of jaguars in Brazilian Amazon. Conserv. Sci. Pract. 3, e477 (2021).Morato, R. G. et al. Resource selection in an apex predator and variation in response to local landscape characteristics. Biol. Conserv. 228, 233–240 (2018).Article 

    Google Scholar 
    Romero-Muñoz, A. et al. Habitat loss and overhunting synergistically drive the extirpation of jaguars from the Gran Chaco. Divers. Distrib. 25, 176–190 (2018).Romero-Muñoz, A., Morato, R. G., Tortato, F. & Kuemmerle, T. Beyond fangs: beef and soybean trade drive jaguar extinction. Front. Ecol. Environ. 18, 67–68 (2020).Article 

    Google Scholar 
    Vilela, T. et al. A better Amazon road network for people and the environment. Proc. Natl Acad. Sci. USA 117, 7095–7102 (2020).Article 
    CAS 

    Google Scholar 
    Benítez-López, A., Alkemade, R. & Verweij, P. A. The impacts of roads and other infrastructure on mammal and bird populations: a meta-analysis. Biol. Conserv. 143, 1307–1316 (2010).Article 

    Google Scholar 
    Carter, N., Killion, A., Easter, T., Brandt, J. & Ford, A. Road development in Asia: assessing the range-wide risks to tigers. Sci. Adv. 6, eaaz9619 (2020).Article 

    Google Scholar 
    Abra, F. D. et al. Pay or prevent? Human safety, costs to society and legal perspectives on animal-vehicle collisions in São Paulo state, Brazil. PLoS ONE 14, e0215152 (2019).Article 
    CAS 

    Google Scholar 
    Joshi, A. R. et al. Tracking changes and preventing loss in critical tiger habitat. Sci. Adv. 2, e1501675 (2016).Article 

    Google Scholar 
    Peres, C. A. & Terborgh, J. Amazonian nature reserves: an analysis of the defensibility status of existing conservation units and design criteria for the future. Conserv. Biol. 9, 34–46 (1995).Article 

    Google Scholar 
    Sistema Nacional de Unidades de Conservação (SNUC). Lei 9985 de 18 de julho de 2000; Ministério do Meio Ambiente. (2000).Stocks, A. Too much for too few: problems of indigenous land rights in Latin America Annual. Rev. Anthropol. 34, 85–104 (2005).Article 

    Google Scholar 
    Mooers, A. Ø., Faith, D. P. & Maddison, W. P. Converting endangered species categories to probabilities of extinction for phylogenetic conservation prioritization. PLoS ONE 3, e3700 (2008).Article 

    Google Scholar 
    Staal, A. et al. Hysteresis of tropical forests in the 21st century. Nat. Commun. 11, 4978 (2020).Article 
    CAS 

    Google Scholar 
    Miranda, E. B. P. et al. Tropical deforestation induces thresholds of reproductive viability and habitat suitability in Earth’s largest eagles. Sci. Rep. 11, 1–17 (2021).Article 

    Google Scholar 
    Bowman, K. W. et al. Environmental degradation of indigenous protected areas of the Amazon as a slow onset event. Curr. Opin. Environ. Sustain. 50, 260–271 (2021).Article 

    Google Scholar 
    Wilson, K. A., Carwardine, J. & Possingham, H. P. Setting conservation priorities. Ann. N. Y. Acad. Sci. 1162, 237–264 (2009).Article 

    Google Scholar 
    Venter, O. et al. Sixteen years of change in the global terrestrial human footprint and implications for biodiversity conservation. Nat. Commun. 7, 12558 (2016).Article 
    CAS 

    Google Scholar 
    Sales, L. P., Galetti, M. & Pires, M. M. Climate and land‐use change will lead to a faunal “savannization” on tropical rainforests. Glob. Change Biol. 26, 7036–7044 (2020).Article 

    Google Scholar 
    da Silva, J. M. C., Dias, T. C. A. C., da Cunha, A. C. & Cunha, H. F. A. Funding deficits of protected areas in Brazil. Land Use Policy 100, 104926 (2021).Article 

    Google Scholar 
    Nobre, C. A. et al. Land-use and climate change risks in the Amazon and the need of a novel sustainable development paradigm. Proc. Natl Acad. Sci. USA 113, 10759–10768 (2016).Article 
    CAS 

    Google Scholar 
    Kauano, E. E., Silva, J. M. C. & Michalski, F. Illegal use of natural resources in federal protected areas of the Brazilian Amazon. PeerJ 5, e3902 (2017).Article 

    Google Scholar 
    Olson, D. M. et al. Terrestrial ecoregions of the world: a new map of life on Earth. Bioscience 51, 933–938 (2011).Article 

    Google Scholar 
    Instituto Brasileiro de Geografia e Estatística (IBGE). Censo demográfico Rio de Janeiro. http://www.ibge.gov.br (2020).Instituto Brasileiro de Geografia e Estatística (IBGE). Censo demográfico Rio de Janeiro. http://www.ibge.gov.br (2010).Instituto Brasileiro de Geografia e Estatística (IBGE). BC250—Base Cartográfica Contínua do Brasil—1:250,000—2017 Diretoria de Geociências—DGC / Coordenação de Cartografia—CCAR. http://www.metadadosgeoibgegovbr/geonetwork_ibge/srv/por/metadatashow?uuid=5a47e9ea-e2cd-423b-8646-53f67ff4ed2d (2017).MapBiomas. Projeto MapBiomas Coleção 5 da Série Anual de Mapas de Cobertura e Uso de Solo do Brasil. https://mapbiomas.org/colecoes-mapbiomas-1 (2019).Brazil’s National Institute for Space Research (INPE). Monitoramento do Desmatamento da Floresta Amazônica Brasileira por Satélite. http://www.obtinpebr/OBT/assuntos/programas/amazonia/prodes (2020a).ESRI. ArcGIS Desktop: Release 10 Redlands. (Environmental Systems Research Institute, 2019).Ministério do Meio Ambiente (MMA). Cadastro Nacional de Unidades de Conservação (CNUC). https://antigo.mma.gov.br/areas-protegidas/cadastro-nacional-de-ucs/dados-georreferenciados.html (2019).Fundação Nacional dos Povos Indígenas (FUNAI). Modalidades de Terra Indígenas. http://www.funaigovbr/indexphp/indios-no-brasil/terras-indigenas (2019).Tobler, M. W. & Powell, G. V. N. Estimating jaguar densities with camera traps: Problems with current designs and recommendations for future studies. Biol. Conserv. 159, 109–118 (2013).Article 

    Google Scholar 
    de Oliveira, T. G. et al. Red list assessment of the jaguar in Brazilian Amazonia. CatNews 7, 8–13 (2012).
    Google Scholar 
    Ramalho, F. B. L. Jaguar (Panthera Onca) Population Dynamics, Feeding Ecology, Human Induced Mortality, and Conservation in the Várzea Floodplain Forests of Amazonia. PhD Thesis. (University of São Paulo, 2012).Duarte, H. O. B., Boron, V., Carvalho, W. D. & Toledo, J. J. Amazon islands as predator refugia: jaguar density and temporal activity in Maracá-Jipioca. J. Mammal. 103, 440–446 (2022).Article 

    Google Scholar 
    Zar, J. H. Biostatistical Analysis 4th edn., (Pretince-Hall, 1999).Medellín, R. A. et al. El jaguar en el nuevo milenio. Fondo de Cultura Económica (Universidad Nacional Autónoma de México, Wildlife Conservation Society, 2002).Quigley, H. et al. Observations and preliminary testing of Jaguar depredation reduction techniques in and between core Jaguar populations. Parks 21, 63–72 (2015).Article 

    Google Scholar 
    Bogoni, J. A., Ferraz, K. M. P. M. B. & Peres, C. A. Continental-scale local extinctions in mammal assemblages are synergistically induced by habitat loss and hunting pressure. Biol. Conserv. 272, 109635 (2022).Article 

    Google Scholar 
    Valsecchi, J., Monteiro, M. C., Alvarenga, G. C., Lemos, L. P. & Ramalho, E. E. Community-based monitoring of wild felid hunting in Central Amazonia. Animal Conser. https://zslpublications.onlinelibrary.wiley.com/doi/pdf/10.1111/acv.12811 (2022).WWF. WWF Jaguar Strategy 2020–2030. https://wwflac.awsassets.panda.org/downloads/estrategia_jaguar_2020_2030_wwf.pdf (2020).Chape, S., Harrison, J., Spalding, M. D. & Lysenko, I. Measuring the extent and effectiveness of protected areas as an indicator for meeting global biodiversity targets. Philos. Trans. R. Soc. Lond. B Biol. Sci. 360, 443–455 (2005).Article 
    CAS 

    Google Scholar 
    R Core Team. R: A language and environment for statistical computing. (R Foundation for Statistical Computing, 2020).Souza-Jr, C. M. et al. Reconstructing Three Decades of Land Use and Land Cover Changes in Brazilian Biomes with Landsat Archive and Earth Engine. Remote Sens. 12, 2735 (2020).Article 

    Google Scholar  More

  • in

    A molecular atlas reveals the tri-sectional spinning mechanism of spider dragline silk

    Chromosomal-scale genome assembly and full spidroin gene set of T. clavata
    To explore dragline silk production in T. clavata, we sought to assemble a high-quality genome of this species. Thus, we first performed a cytogenetic analysis of T. clavata captured from the wild in Dali City, Yunnan Province, China, and found a chromosomal complement of 2n = 26 in females and 2n = 24 in males, comprising eleven pairs of autosomal elements and unpaired sex chromosomes (X1X1X2X2 in females and X1X2 in males) (Fig. 1a). Then, DNA from adult T. clavata was used to generate long-read (Oxford Nanopore Technologies (ONT)), short-read (Illumina), and Hi-C data (Supplementary Data 1). A total of 349.95 Gb of Nanopore reads, 199.55 Gb of Illumina reads, and ~438.41 Gb of Hi-C raw data were generated. Our sequential assembly approach (Supplementary Fig. 1c) resulted in a 2.63 Gb genome with a scaffold N50 of 202.09 Mb and a Benchmarking Universal Single-Copy Ortholog (BUSCO) genome completeness score of 93.70% (Table 1; Supplementary Data 3). Finally, the genome was assembled into 13 pseudochromosomes. Sex-specific Pool-Seq analysis of spiders indicated that Chr12 and Chr13 were sex chromosomes (Fig. 1b; Supplementary Fig. 2). Based on the MAKER2 pipeline34 (Supplementary Fig. 1e), we annotated 37,607 protein-encoding gene models and predicted repetitive elements with a collective length of 1.42 Gb, accounting for 53.94% of the genome.Table 1 Characteristics of the T. clavata genome assemblyFull size tableTo identify T. clavata spidroin genes, we searched the annotated gene models for sequences similar to 443 published spidroins (Supplementary Data 6) and performed a phylogenetic analysis of the putative spidroin sequences for classification (Supplementary Fig. 12a). Based on the knowledge that a typical spidroin gene consists of a long repeat domain sandwiched between the nonrepetitive N/C-terminal domains16, 128 nonrepetitive hits were primarily identified. These candidates were further validated and reconstructed using full-length transcript isoform sequencing (Iso-seq) and transcriptome sequencing (RNA-seq) data. We thus identified 28 spidroin genes, among which 26 were full-length (Supplementary Fig. 11a), including 9 MaSps, 5 minor ampullate spidroins (MiSps), 2 flagelliform spidroins (FlSps), 1 tubuliform spidroin (TuSp), 2 aggregate spidroins (AgSp), 1 aciniform spidroin (AcSp), 1 pyriform spidroin (PySp), and 5 other spidroins. This full set of spidroin genes was located across nine of the 13 T. clavata chromosomes. Interestingly, we found that the MaSp1a–c & MaSp2e, MaSp2a–d, and MiSp-a–e genes were distributed in three independent groups on chromosomes 4, 7, and 6, respectively (Fig. 1c). Notably, using the genomic data of another orb-weaving spider species, Trichonephila antipodiana35, we identified homologous group distributions of spidroin genes on T. antipodiana chromosomes (Fig. 1d), which indicated the reliability of the grouping results of our study. When we compared the spidroin gene catalog of T. clavata and those of five other orb-web spider species with genomic data28,29,36,37, we found that T. clavata and Trichonephila clavipes possessed the largest number of spidroin genes (28 genes in both species; Fig. 1e).To further explore the expression of spidroin genes in different glands, all morphologically distinct glands (major and minor ampullate- (Ma and Mi), flagelliform- (Fl), tubuliform- (Tu), and aggregate (Ag) glands) were cleanly and separately dissected from adult female T. clavata spiders except for the aciniform and pyriform glands, which could not be cleanly separated because of their proximal anatomical locations and were therefore treated as a combined sample (aciniform & pyriform gland (Ac & Py)). After RNA sequencing of these silk glands, we performed expression clustering analysis of transcriptomic data and found that the Ma and Mi glands showed the closest relationship in terms of both morphological structure (Fig. 1g) and gene expression (Fig. 1f, h). We noted that the expression profiles of spidroin genes were largely consistent with their putative roles in the corresponding morphologically distinct silk glands; for example, MaSp expression was found in the Ma gland (Fig. 1h). However, some spidroin transcripts, such as MiSps and TuSp, were expressed in several silk glands (Fig. 1h). Unclassified spidroin genes, such as Sp-GP-rich, did not appear to show gland-specific expression (Fig. 1h).In summary, the chromosomal-scale genome of T. clavata allowed us to obtain detailed structural and location information for all spidroin genes of this species. We also found a relatively diverse set of spidroin genes and a grouped distribution of MaSps and MiSps in T. clavata.Dragline silk origin and the functional character of the Ma gland segmentsTo further evaluate the detailed molecular characteristics of the Ma gland-mediated secretion of dragline silk, we performed integrated analyses of the transcriptomes of the three T. clavata Ma gland segments and the proteome and metabolome of T. clavata dragline silk (Fig. 2a). Sodium dodecyl sulfate–polyacrylamide gel electrophoresis (SDS–PAGE) analysis of dragline silk mainly showed a thick band above 240 kDa, suggesting a relatively small variety of total proteins (Fig. 2b). Subsequent liquid chromatography–mass spectrometry (LC–MS) analysis identified 28 proteins, including ten spidroins (nine MaSps and one MiSp) and 18 nonspidroin proteins (one glucose dehydrogenase (GDH), one mucin-19, one venom protein, and 15 SpiCEs of dragline silk (SpiCE-DS)) (Fig. 2b; Supplementary Data 10). Among these proteins, we found that the core protein components of dragline silk in order of intensity-based absolute quantification (iBAQ) percentages were MaSp1c (37.7%), MaSp1b (12.2%), SpiCE-DS1 (11.9%, also referred to as SpiCE-NMa1 in a previous study28), MaSp1a (10.4%), and MaSp-like (7.2%), accounting for approximately 80% of the total protein abundance in dragline silk (Fig. 2b). These results revealed potential protein components that might be highly correlated with the excellent strength and toughness of dragline silk.Fig. 2: Dragline silk origin and the functional character of the Ma gland segments.a Schematic illustration of Ma gland segmentation. b Sodium dodecyl sulfate–polyacrylamide gel electrophoresis (SDS–PAGE) (left) and LC–MS (right) analyses of dragline silk protein. iBAQ, intensity-based absolute quantification. Similar results were obtained in three independent experiments and summarized in Source data. c Classification of the identified metabolites in dragline silk. d LC–MS analyses of the metabolites. e LC–MS analyses of the golden extract from T. clavata dragline silk. The golden pigment was extracted with 80% methanol. The extracted ion chromatograms (EICs) showed a peak at m/z 206 [M + H]+ for xanthurenic acid. f Pearson correlation of different Ma gland segments (Tail, Sac, and Duct). g Expression clustering of the Tail, Sac, and Duct. The transcriptomic data were clustered according to the hierarchical clustering (HC) method. h Combinational analysis of the transcriptome and proteome showing the expression profile of the dragline silk genes in the Tail, Sac, and Duct. i Concise biosynthetic pathway of xanthurenic acid (tryptophan metabolism) in the T. clavata Ma gland. Gene expression levels mapped to tryptophan metabolism are shown in three segments of the Ma gland. Enzymes involved in the pathway are indicated in red, and the genes encoding the enzymes are shown beside them. j Gene Ontology (GO) enrichment analysis of Ma gland segment-specific genes indicating the biological functions of the Tail, Sac, and Duct. The top 12 significantly enriched GO terms are shown for each segment of the Ma gland. A P-value  2) were identified in the 2 kb regions upstream and downstream of genes, and 10,501,151 (Tail), 11,356,55 (Sac), and 9,778,368 (Duct) significant ATAC peaks (RPKM  > 2) were identified at the whole-genome level. The Tail (mean RPKM: 1.78) and Sac (mean RPKM: 2.04) plots showed genes with more accessible chromatin than the Duct (mean RPKM: 1.59) plots (Fig. 3a). We then analyzed the genome-wide DNA methylation level in the Tail, Sac, and Duct. We found the highest levels of DNA methylation in the CG context (beta value: 0.12 in Tail, 0.13 in Sac, and 0.10 in Duct) and only a small amount in the CHH (beta value: 0.04 in Tail, 0.05 in Sac, and 0.03 in Duct) and CHG (beta value: 0.04 in Tail, 0.05 in Sac, and 0.04 in Duct) contexts (Fig. 3b). Overall, there was no significant difference in methylation levels among the Tail, Sac, and Duct. Taken together, our results suggest a potential regulatory role of CA rather than DNA methylation in the transcription of dragline silk genes.Fig. 3: Comprehensive epigenetic features and ceRNA network of the tri-sectional Ma gland.a Metagene plot of ATAC-seq signals and heatmap of the ATAC-seq read densities in the Tail, Sac, and Duct. The chromatin accessibility was indicated by the mean RPKM value (upper) and the blue region (bottom). b Metagene plot of DNA methylation levels in CG/CHG/CHH contexts in the Tail, Sac, and Duct. (c, d) Screenshots of the methylation and ATAC-seq tracks of the MaSp1b (c) and MaSp2b (d) genes within the Tail, Sac, and Duct. The potential TF motifs (E-value More