More stories

  • in

    Co-activation of Akt, Nrf2, and NF-κB signals under UPRER in torpid Myotis ricketti bats for survival

    1.
    Carey, H. V., Andrews, M. T. & Martin, S. L. Mammalian hibernation: cellular and molecular responses to depressed metabolism and low temperature. Physiological Rev. 83, 1153–1181 (2003).
    CAS  Article  Google Scholar 
    2.
    Geiser, F. Metabolic rate and body temperature reduction during hibernation and daily torpor. Annu. Rev. Physiol. 66, 239–274 (2004).
    CAS  Article  Google Scholar 

    3.
    Lindell, S. L. et al. Natural resistance to liver cold ischemia-reperfusion injury associated with the hibernation phenotype. Am. J. Physiol.-Gastrointest. Liver Physiol. 288, G473–G480 (2005).
    CAS  Article  Google Scholar 

    4.
    Dave, K. R., Christian, S. L., Perez-Pinzon, M. A. & Drew, K. L. Neuroprotection: lessons from hibernators. Comp. Biochem. Physiol. Part B: Biochem. Mol. Biol. 162, 1–9 (2012).
    CAS  Article  Google Scholar 

    5.
    Hofmann, S., Cherkasova, V., Bankhead, P., Bukau, B. & Stoecklin, G. Translation suppression promotes stress granule formation and cell survival in response to cold shock. Mol. Biol. Cell 23, 3786–3800 (2012).
    CAS  Article  Google Scholar 

    6.
    Pluquet, O., Pourtier, A. & Abbadie, C. The unfolded protein response and cellular senescence. A review in the theme: cellular mechanisms of endoplasmic reticulum stress signaling in health and disease. Am. J. Physiol.-Cell Physiol. 308, C415–C425 (2015).
    CAS  Article  Google Scholar 

    7.
    Hetz, C. The unfolded protein response: controlling cell fate decisions under ER stress and beyond. Nat. Rev. Mol. Cell Biol. 13, 89 (2012).
    CAS  Article  Google Scholar 

    8.
    Guo, F.-J. et al. XBP1S protects cells from ER stress-induced apoptosis through Erk1/2 signaling pathway involving CHOP. Histochemistry Cell Biol. 138, 447–460 (2012).
    CAS  Article  Google Scholar 

    9.
    Walter, P. & Ron, D. The unfolded protein response: from stress pathway to homeostatic regulation. Science 334, 1081–1086 (2011).
    CAS  Article  Google Scholar 

    10.
    Kramer, G. Two phosphorylation sites on eIF-2α. FEBS Lett. 267, 181–182 (1990).
    CAS  Article  Google Scholar 

    11.
    Wek, R., Jiang, H.-Y. & Anthony, T. Coping with stress: eIF2 kinases and translational control. (Portland Press Limited, 2006).

    12.
    Palam, L. R., Baird, T. D. & Wek, R. C. Phosphorylation of eIF2 facilitates ribosomal bypass of an inhibitory upstream ORF to enhance CHOP translation. J. Biol. Chem. 286, 10939–10949 (2011).

    13.
    Roller, C. & Maddalo, D. The molecular chaperone GRP78/BiP in the development of chemoresistance: mechanism and possible treatment. Front. Pharmacol. 4, 10 (2013).
    Article  CAS  Google Scholar 

    14.
    Lee, A. S. The ER chaperone and signaling regulator GRP78/BiP as a monitor of endoplasmic reticulum stress. Methods 35, 373–381 (2005).
    CAS  Article  Google Scholar 

    15.
    Jiang, H.-Y. et al. Phosphorylation of the α subunit of eukaryotic initiation factor 2 is required for activation of NF-κB in response to diverse cellular stresses. Mol. Cell. Biol. 23, 5651–5663 (2003).
    CAS  Article  Google Scholar 

    16.
    Prell, T. et al. Endoplasmic reticulum stress is accompanied by activation of NF-κB in amyotrophic lateral sclerosis. J. Neuroimmunol. 270, 29–36 (2014).
    CAS  Article  Google Scholar 

    17.
    Rajesh, K. et al. Phosphorylation of the translation initiation factor eIF2α at serine 51 determines the cell fate decisions of Akt in response to oxidative stress. Cell Death Dis. 6, e1591 (2015).
    CAS  Article  Google Scholar 

    18.
    Nivon, M. et al. NFκB is a central regulator of protein quality control in response to protein aggregation stresses via autophagy modulation. Mol. Biol. Cell 27, 1712–1727 (2016).
    CAS  Article  Google Scholar 

    19.
    Lemasters, J. J. In Molecular Pathology (second edn) 1–24 (Elsevier, 2018).

    20.
    Karin, M. & Lin, A. NF-κB at the crossroads of life and death. Nat. Immunol. 3, 221 (2002).
    CAS  Article  Google Scholar 

    21.
    Schmidlin, C. J., Dodson, M. B., Madhavan, L. & Zhang, D. D. Redox regulation by NRF2 in aging and disease. Free Rad. Biol. Med. 134, 702–707 (2019).

    22.
    Cullinan, S. B. & Diehl, J. A. Coordination of ER and oxidative stress signaling: the PERK/Nrf2 signaling pathway. Int. J. Biochem. Cell Biol. 38, 317–332 (2006).
    CAS  Article  Google Scholar 

    23.
    Wiersma, M. et al. Torpor-arousal cycles in Syrian hamster heart are associated with transient activation of the protein quality control system. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 223, 23–28 (2018).

    24.
    Mamady, H. & Storey, K. B. Up-regulation of the endoplasmic reticulum molecular chaperone GRP78 during hibernation in thirteen-lined ground squirrels. Mol. Cell. Biochem. 292, 89–98 (2006).
    CAS  Article  Google Scholar 

    25.
    Mamady, H. & Storey, K. B. Coping with the stress: expression of ATF4, ATF6, and downstream targets in organs of hibernating ground squirrels. Arch. Biochem. Biophys. 477, 77–85 (2008).
    CAS  Article  Google Scholar 

    26.
    Zhang, J. et al. Prosurvival roles mediated by the PERK signaling pathway effectively prevent excessive endoplasmic reticulum stress-induced skeletal muscle loss during high-stress conditions of hibernation. J. Cell. Physiol. 234, 19728–19739 (2019).

    27.
    Carey, H., Frank, C. & Seifert, J. Hibernation induces oxidative stress and activation of NF-κB in ground squirrel intestine. J. Comp. Physiol. B 170, 551–559 (2000).
    CAS  Article  Google Scholar 

    28.
    Fleck, C. C. & Carey, H. V. Modulation of apoptotic pathways in intestinal mucosa during hibernation. Am. J. Physiol.-Regulatory, Integr. Comp. Physiol. 289, R586–R595 (2005).
    CAS  Article  Google Scholar 

    29.
    Zhang, Y. et al. Critical roles of mitochondria in brain activities of torpid Myotis ricketti bats revealed by a proteomic approach. J. Proteom. 105, 266–284 (2014).
    CAS  Article  Google Scholar 

    30.
    Cui, X. A., Zhang, H. & Palazzo, A. F. p180 promotes the ribosome-independent localization of a subset of mRNA to the endoplasmic reticulum. PLoS Biol. 10, e1001336 (2012).
    CAS  Article  Google Scholar 

    31.
    Wen, W.-L. et al. Vgl1, a multi-KH domain protein, is a novel component of the fission yeast stress granules required for cell survival under thermal stress. Nucleic Acids Res. 38, 6555–6566 (2010).
    CAS  Article  Google Scholar 

    32.
    Tsuchiya, N. et al. SND1, a component of RNA-induced silencing complex, is up-regulated in human colon cancers and implicated in early stage colon carcinogenesis. Cancer Res. 67, 9568–9576 (2007).
    CAS  Article  Google Scholar 

    33.
    Yoo, B. K. et al. Increased RNA-induced silencing complex (RISC) activity contributes to hepatocellular carcinoma. Hepatology 53, 1538–1548 (2011).
    CAS  Article  Google Scholar 

    34.
    Halperin, L., Jung, J. & Michalak, M. The many functions of the endoplasmic reticulum chaperones and folding enzymes. IUBMB Life 66, 318–326 (2014).
    CAS  Article  Google Scholar 

    35.
    Obchoei, S. et al. Cyclophilin A: potential functions and therapeutic target for human cancer. Med. Sci. Monit. 15, RA221–RA232 (2009).
    CAS  Google Scholar 

    36.
    Wei, Y. et al. Antiapoptotic and proapoptotic signaling of cyclophilin A in endothelial cells. Inflammation 36, 567–572 (2013).
    Article  CAS  Google Scholar 

    37.
    Kelleher, D. J. & Gilmore, R. DAD1, the defender against apoptotic cell death, is a subunit of the mammalian oligosaccharyltransferase. Proc. Natl Acad. Sci. USA 94, 4994–4999 (1997).
    CAS  Article  Google Scholar 

    38.
    Zhou, L. et al. DsbA-L alleviates endoplasmic reticulum stress-induced adiponectin downregulation. Diabetes 59, 2809–2816 (2010).
    CAS  Article  Google Scholar 

    39.
    Liu, M. et al. Endoplasmic reticulum (ER) localization is critical for DsbA-L protein to suppress ER stress and adiponectin down-regulation in adipocytes. J. Biol. Chem. 290, 10143–10148 (2015).
    CAS  Article  Google Scholar 

    40.
    Santhekadur, P. K. et al. Multifunction protein staphylococcal nuclease domain containing 1 (SND1) promotes tumor angiogenesis in human hepatocellular carcinoma through novel pathway that involves nuclear factor κB and miR-221. J. Biol. Chem. 287, 13952–13958 (2012).
    CAS  Article  Google Scholar 

    41.
    Pan, Y. H. et al. Adaptation of phenylalanine and tyrosine catabolic pathway to hibernation in bats. PLoS ONE 8, e62039 (2013).
    CAS  Article  Google Scholar 

    42.
    Vattem, K. M. & Wek, R. C. Reinitiation involving upstream ORFs regulates ATF4 mRNA translation in mammalian cells. Proc. Natl Acad. Sci. USA 101, 11269–11274 (2004).
    CAS  Article  Google Scholar 

    43.
    Lu, Z. & Xu, S. ERK1/2 MAP kinases in cell survival and apoptosis. IUBMB Life 58, 621–631 (2006).
    CAS  Article  Google Scholar 

    44.
    Sun, F.-C. et al. Localization of GRP78 to mitochondria under the unfolded protein response. Biochem. J. 396, 31–39 (2006).
    CAS  Article  Google Scholar 

    45.
    van Breukelen, F. & Martin, S. L. Translational initiation is uncoupled from elongation at 18 C during mammalian hibernation. Am. J. Physiol.-Regulatory, Integr. Comp. Physiol. 281, R1374–R1379 (2001).
    Article  Google Scholar 

    46.
    van Breukelen, F., Sonenberg, N. & Martin, S. L. Seasonal and state-dependent changes of eIF4E and 4E-BP1 during mammalian hibernation: implications for the control of translation during torpor. Am. J. Physiol.-Regulatory, Integr. Comp. Physiol. 287, R349–R353 (2004).
    Article  Google Scholar 

    47.
    Pan, P. & van Breukelen, F. Preference of IRES-mediated initiation of translation during hibernation in golden-mantled ground squirrels, Spermophilus lateralis. Am. J. Physiol. Regul. Integr. Comp. Physiol. 301, R370–R377 (2011).
    CAS  Article  Google Scholar 

    48.
    Liu, L. et al. Hypoxia-induced energy stress regulates mRNA translation and cell growth. Mol. Cell 21, 521–531 (2006).
    Article  CAS  Google Scholar 

    49.
    Harvey, R., Dezi, V., Pizzinga, M. & Willis, A. E. Post-transcriptional control of gene expression following stress: the role of RNA-binding proteins. Biochem. Soc. Trans. 45, 1007–1014 (2017).
    CAS  Article  Google Scholar 

    50.
    Mobin, M. B. et al. The RNA-binding protein vigilin regulates VLDL secretion through modulation of Apob mRNA translation. Nat. Commun. 7, 12848 (2016).
    CAS  Article  Google Scholar 

    51.
    Srere, H. K., Wang, L. & Martin, S. L. Central role for differential gene expression in mammalian hibernation. Proc. Natl Acad. Sci. USA 89, 7119–7123 (1992).
    CAS  Article  Google Scholar 

    52.
    Han, Y. et al. Adaptation of peroxisome proliferator-activated receptor alpha to hibernation in bats. BMC Evolut. Biol. 15, 88 (2015).
    Article  CAS  Google Scholar 

    53.
    Lee, M., Choi, I. & Park, K. Activation of stress signaling molecules in bat brain during arousal from hibernation. J. Neurochem. 82, 867–873 (2002).
    CAS  Article  Google Scholar 

    54.
    Storey, K. B. Out cold: biochemical regulation of mammalian hibernation-a mini-review. Gerontology 56, 220–230 (2010).
    Article  Google Scholar 

    55.
    Lei, M., Dong, D., Mu, S., Pan, Y.-H. & Zhang, S. Comparison of brain transcriptome of the greater horseshoe bats (Rhinolophus ferrumequinum) in active and torpid episodes. PLoS ONE 9, e107746 (2014).
    Article  CAS  Google Scholar 

    56.
    Wortel, I. M., van der Meer, L. T., Kilberg, M. S. & van Leeuwen, F. N. Surviving stress: modulation of ATF4-mediated stress responses in normal and malignant cells. Trends Endocrinol. Metab. 28, 794–806 (2017).
    CAS  Article  Google Scholar 

    57.
    Cullinan, S. B. et al. Nrf2 is a direct PERK substrate and effector of PERK-dependent cell survival. Mol. Cell. Biol. 23, 7198–7209 (2003).
    CAS  Article  Google Scholar 

    58.
    Ni, M., Zhang, Y. & Lee, A. S. Beyond the endoplasmic reticulum: atypical GRP78 in cell viability, signalling and therapeutic targeting. Biochem. J. 434, 181–188 (2011).
    CAS  Article  Google Scholar 

    59.
    Yin, Q. et al. Antioxidant defenses in the brains of bats during hibernation. PLoS ONE 11, e0152135 (2016).
    Article  CAS  Google Scholar 

    60.
    Allan, M. E. & Storey, K. B. Expression of NF-κB and downstream antioxidant genes in skeletal muscle of hibernating ground squirrels, Spermophilus tridecemlineatus. Cell Biochem. Funct. 30, 166–174 (2012).
    CAS  Article  Google Scholar 

    61.
    Ni, Z., McMullen, D. C. & Storey, K. B. Expression of Nrf2 and its downstream gene targets in hibernating 13-lined ground squirrels, Spermophilus tridecemlineatus. Mol. Cell. Biochem. 312, 121–129 (2008).
    Article  CAS  Google Scholar 

    62.
    Németh, J. et al. S100A8 and S100A9 are novel nuclear factor kappa B target genes during malignant progression of murine and human liver carcinogenesis. Hepatology 50, 1251–1262 (2009).
    Article  CAS  Google Scholar 

    63.
    Sun, S. et al. Cyclophilin A (CypA) interacts with NF-κB subunit, p65/RelA, and contributes to NF-κB activation signaling. PLoS ONE 9, e96211 (2014).
    Article  CAS  Google Scholar 

    64.
    Bolignano, D. et al. Neutrophil gelatinase-associated lipocalin (NGAL) in human neoplasias: a new protein enters the scene. Cancer Lett. 288, 10–16 (2010).
    CAS  Article  Google Scholar 

    65.
    Drew, K. L., Rice, M. E., Kuhn, T. B. & Smith, M. A. Neuroprotective adaptations in hibernation: therapeutic implications for ischemia-reperfusion, traumatic brain injury and neurodegenerative diseases. Free Radic. Biol. Med. 31, 563–573 (2001).
    CAS  Article  Google Scholar 

    66.
    Bouma, H. R. et al. Induction of torpor: mimicking natural metabolic suppression for biomedical applications. J. Cell Physiol. 227, 1285–1290 (2012).
    CAS  Article  Google Scholar 

    67.
    Cerri, M. et al. Hibernation for space travel: Impact on radioprotection. Life Sci. Space Res. 11, 1–9 (2016).
    Article  Google Scholar 

    68.
    Uchida, Y., Tokizawa, K. & Nagashima, K. Characteristics of activated neurons in the suprachiasmatic nucleus when mice become hypothermic during fasting and cold exposure. Neurosci. Lett. 579, 177–182 (2014).
    CAS  Article  Google Scholar 

    69.
    Sato, N., Marui, S., Ozaki, M. & Nagashima, K. Cold exposure and/or fasting modulate the relationship between sleep and body temperature rhythms in mice. Physiol. Behav. 149, 69–75 (2015).
    CAS  Article  Google Scholar 

    70.
    Tokizawa, K., Uchida, Y. & Nagashima, K. Thermoregulation in the cold changes depending on the time of day and feeding condition: physiological and anatomical analyses of involved circadian mechanisms. Neuroscience 164, 1377–1386 (2009).
    CAS  Article  Google Scholar 

    71.
    Van Breukelen, F. & Martin, S. L. Invited review: molecular adaptations in mammalian hibernators: unique adaptations or generalized responses? J. Appl. Physiol. 92, 2640–2647 (2002).
    Article  Google Scholar 

    72.
    Piersma, S. R. et al. Workflow comparison for label-free, quantitative secretome proteomics for cancer biomarker discovery: method evaluation, differential analysis, and verification in serum. J. Proteome Res. 9, 1913–1922 (2010).
    CAS  Article  Google Scholar 

    73.
    Perez-Riverol, Y. et al. The PRIDE database and related tools and resources in 2019: improving support for quantification data. Nucleic Acids Res. 47, D442–D450 (2018).
    Article  CAS  Google Scholar 

    74.
    Yu, D. et al. EGPS 1.0: Comprehensive software for multi-omic and evolutionary analyses. Natl Sci. Rev. 6, 867–869 (2019).

    75.
    Yin, Q. et al. Maintenance of neural activities in torpid Rhinolophus ferrumequinum bats revealed by 2D gel-based proteome analysis. Biochim. Biophys. Acta Proteins Proteom. 1865, 1004–1019 (2017).
    CAS  Article  Google Scholar 

    76.
    Romero-Calvo, I. et al. Reversible Ponceau staining as a loading control alternative to actin in Western blots. Anal. Biochem. 401, 318–320 (2010).
    CAS  Article  Google Scholar 

    77.
    Deutsch, E. W. et al. The ProteomeXchange consortium in 2020: enabling ‘big data’approaches in proteomics. Nucleic Acids Res. 48, D1145–D1152 (2020).
    CAS  PubMed  Google Scholar  More

  • in

    Toxicity of the herbicides diuron, propazine, tebuthiuron, and haloxyfop to the diatom Chaetoceros muelleri

    1.
    Carbery, K., Owen, R., Frickers, T., Otero, E. & Readman, J. Contamination of Caribbean coastal waters by the antifouling herbicide Irgarol 1051. Mar. Pollut. Bull. 52, 635–644. https://doi.org/10.1016/j.marpolbul.2005.10.013 (2006).
    CAS  Article  Google Scholar 
    2.
    Hernández-Romero, A. H., Tovilla-Hernández, C., Malo, E. A. & Bello-Mendoza, R. Water quality and presence of pesticides in a tropical coastal wetland in southern Mexico. Mar. Pollut. Bull. 48, 1130–1141. https://doi.org/10.1016/j.marpolbul.2004.01.003 (2004).
    CAS  Article  Google Scholar 

    3.
    Castillo, L. E., de la Cruz, E. & Ruepert, C. Ecotoxicology and pesticides in tropical aquatic ecosystems of Central America. Environ. Toxicol. Chem. 16, 41–51. https://doi.org/10.1002/etc.5620160104 (1997).
    CAS  Article  Google Scholar 

    4.
    Basheer, C., Obbard, J. P. & Lee, H. K. Persistent organic pollutants in Singapore’s coastal marine environment: part I, seawater. Water Air Soil Pollut. 149, 295–313. https://doi.org/10.1023/A:1025689600993 (2003).
    ADS  CAS  Article  Google Scholar 

    5.
    Ali, H. R. et al. Contamination of diuron in coastal waters around Malaysian Peninsular. Mar. Pollut. Bull. 85, 287–291. https://doi.org/10.1016/j.marpolbul.2014.05.049 (2014).
    CAS  Article  Google Scholar 

    6.
    Okamura, H., Aoyama, I., Ono, Y. & Nishida, T. Antifouling herbicides in the coastal waters of western Japan. Mar. Pollut. Bull. 47, 59–67. https://doi.org/10.1016/S0025-326X(02)00418-6 (2003).
    CAS  Article  Google Scholar 

    7.
    Roche, H., Salvat, B. & Ramade, F. Assessment of the pesticides pollution of coral reefs communities from French Polynesia. Rev. Ecol. https://hdl.handle.net/2042/55860 (2011).

    8.
    Sarkar, S. K. et al. Occurrence, distribution and possible sources of organochlorine pesticide residues in tropical coastal environment of India: an overview. Environ. Int. 34, 1062–1071. https://doi.org/10.1016/j.envint.2008.02.010 (2008).
    CAS  Article  Google Scholar 

    9.
    Devlin, M. M. et al. Advancing our understanding of the source, management, transport and impacts of pesticides on the Great Barrier Reef 2011–2015. Report for the Queensland Department of Environment and Heritage Protection. Tropical Water & Aquatic Ecosystem Research (TropWATER) Publication, James Cook University, Cairns, Australia (2015).

    10.
    GBR. Great Barrier Reef Marine Park Authority 2019, Great Barrier Reef Outlook Report 2019, GBRMPA, Townsville. https://www.gbrmpa.gov.au/our-work/outlook-report-2019 (2019).

    11.
    Brodie, J. et al. Terrestrial pollutant runoff to the Great Barrier Reef: an update of issues, priorities and management responses. Mar. Pollut. Bull. 65, 81–100. https://doi.org/10.1016/j.marpolbul.2011.12.012 (2012).
    CAS  Article  Google Scholar 

    12.
    Lewis, S. E. et al. Herbicides: a new threat to the Great Barrier Reef. Environ. Pollut. 157, 2470–2484. https://doi.org/10.1016/j.envpol.2009.03.006 (2009).
    CAS  Article  Google Scholar 

    13.
    RWQIP. Reef 2050 Water Quality Improvement Plan 2017–2022. Australian and Queensland Government. https://www.reefplan.qld.gov.au/__data/assets/pdf_file/0017/46115/reef-2050-water-quality-improvement-plan-2017-22.pdf (2018).

    14.
    Shaw, M. et al. Monitoring pesticides in the Great Barrier Reef. Mar. Pollut. Bull. 60, 113–122. https://doi.org/10.1016/j.marpolbul.2009.08.026 (2010).
    CAS  Article  Google Scholar 

    15.
    Grant, S. et al. Marine Monitoring Program: Annual report for inshore pesticide monitoring 2015–2016 (Report for the Great Barrier Reef Marine Park Authority, Great Barrier Reef Marine Park Authority, Townsville, Australia, 2017).
    Google Scholar 

    16.
    O’Brien, D. et al. Spatial and temporal variability in pesticide exposure downstream of a heavily irrigated cropping area: application of different monitoring techniques. J. Agric. Food Chem. 64, 3975–3989. https://doi.org/10.1021/acs.jafc.5b04710 (2016).
    CAS  Article  Google Scholar 

    17.
    Radcliffe, J. Pesticide use in Australia. A review undertaken by the Australian Academy of Technological Sciences, Victoria, Australia. https://www.atse.org.au/ (2002).

    18.
    Oettmeier, W. Herbicides of photosystems II. in Structure, Function and Molecular Biology (Barber, J., ed) (Elsevier, Amsterdam, 349–408). https://doi.org/10.1016/B978-0-444-89440-3.50018-7 (1992).

    19.
    Lewis, S. E. et al. Using monitoring data to model herbicides exported to the Great Barrier Reef, Australia. in The 19th International Congress on Modelling and Simulation, Modelling and Simulation Society of Australia and New Zealand. MODSIM2011, 2051–2056 (2011).

    20.
    Kennedy, K. et al. The influence of a season of extreme wet weather events on exposure of the World Heritage Area Great Barrier Reef to pesticides. Mar. Pollut. Bull. 64, 1495–1507. https://doi.org/10.1016/j.marpolbul.2012.05.014 (2012).
    CAS  Article  Google Scholar 

    21.
    Kennedy, K. et al. Long term monitoring of photosystem II herbicides: correlation with remotely sensed freshwater extent to monitor changes in the quality of water entering the Great Barrier Reef, Australia. Mar. Pollut. Bull. 65, 292–305. https://doi.org/10.1016/j.marpolbul.2011.10.029 (2012).
    CAS  Article  Google Scholar 

    22.
    Mercurio, P., Mueller, J. F., Eaglesham, G., Flores, F. & Negri, A. P. Herbicide persistence in seawater simulation experiments. PLoS ONE 10, e0136391. https://doi.org/10.1371/journal.pone.0136391 (2015).
    CAS  Article  Google Scholar 

    23.
    Mercurio, P. et al. Degradation of herbicides in the tropical marine environment: influence of light and sediment. PLoS ONE 11, e0165890. https://doi.org/10.1371/journal.pone.0165890 (2016).
    CAS  Article  Google Scholar 

    24.
    Gallen, C. et al. Marine Monitoring Program: Annual report for inshore pesticide monitoring 2017–18. Report for the Great Barrier Reef Marine Park Authority, Great Barrier Reef Marine Park Authority, Townsville, Australia. https://elibrary.gbrmpa.gov.au/jspui/handle/11017/3489 (2019).

    25.
    Davis, A., Lewis, S., Brodie, J. & Benson, A. The potential benefits of herbicide regulation: a cautionary note for the Great Barrier Reef catchment area. Sci. Total Environ. 490, 81–92. https://doi.org/10.1016/j.scitotenv.2014.04.005 (2014).
    ADS  CAS  Article  Google Scholar 

    26.
    Thomas, M. C., Flores, F., Kaserzon, S., Fisher, R. & Negri, A. P. Toxicity of ten herbicides to the tropical marine microalgae Rhodomonas salina. Sci. Rep. 10, 7612. https://doi.org/10.1038/s41598-020-64116-y (2020).
    ADS  CAS  Article  Google Scholar 

    27.
    ANZG. Revised Australian and New Zealand guidelines for fresh and marine water quality. Australian and New Zealand Environment and Conservation Council and Agriculture and Resource Management Council of Australia and New Zealand, Canberra, Australia. https://www.waterquality.gov.au/anz-guidelines/guideline-values/default/water-quality-toxicants/toxicants (2018).

    28.
    Warne, M. St. J. et al. Revised method for deriving Australian and New Zealand water quality guideline values for toxicants: update of 2015 version. Prepared for the revision of the Australian and New Zealand guidelines for fresh and marine water quality. Australian and New Zealand Governments and Australian state and territory governments, Canberra, Australia. 48 pp, https://www.waterquality.gov.au/anz-guidelines/guideline-values/derive/warne-method-derive. https://doi.org/10.13140/RG.2.2.36577.35686 (2018).

    29.
    Warne, M. St. J., Smith, R. & Turner, R. Analysis of pesticide mixtures discharged to the lagoon of the Great Barrier Reef, Australia. Environ. Pollut. 265, 114088. https://doi.org/10.1016/j.envpol.2020.114088 (2020).
    CAS  Article  Google Scholar 

    30.
    Magnusson, M., Heimann, K., Quayle, P. & Negri, A. P. Additive toxicity of herbicide mixtures and comparative sensitivity of tropical benthic microalgae. Mar. Pollut. Bull. 60, 1978–1987. https://doi.org/10.1016/j.marpolbul.2010.07.031 (2010).
    CAS  Article  Google Scholar 

    31.
    Faust, M. et al. Predicting the joint algal toxicity of multi-component s-triazine mixtures at low-effect concentrations of individual toxicants. Aquat. Toxicol. 56, 13–32. https://doi.org/10.1016/S0166-445X(01)00187-4 (2001).
    CAS  Article  Google Scholar 

    32.
    Wilkinson, A. D., Collier, C. J., Flores, F. & Negri, A. P. Acute and additive toxicity of ten photosystem-II herbicides to seagrass. Sci. Rep. 5, 17443. https://doi.org/10.1038/srep17443 (2015).
    ADS  CAS  Article  Google Scholar 

    33.
    Traas, T. P. et al. The potentially affected fraction as a measure of ecological risk. in Species sensitivity distributions in ecotoxicology (L. Posthuma, & G. W. Suter, Eds.) (pp. 315–344). https://doi.org/10.1201/9781420032314-20 (2002).

    34.
    Negri, A. P. et al. Adjusting tropical marine water quality guideline values for elevated ocean temperatures. Environ. Sci. Technol. 54, 1102–1110. https://doi.org/10.1021/acs.est.9b05961 (2019).
    ADS  CAS  Article  Google Scholar 

    35.
    King, O., Smith, R., Mann, R. & Warne, M. St. J. Proposed aquatic ecosystem protection guideline values for pesticides commonly used in the Great Barrier Reef catchment area: Part 1 (amended): 2,4-D, Ametryn, Diuron, Glyphosate, Hexazinone, Imazapic, Imidacloprid, Isoxaflutole, Metolachlor, Metribuzin, Metsulfuron-methyl, Simazine, Tebuthiuron. Department of Environment and Science, Brisbane, Australia. https://www.publications.qld.gov.au/dataset/proposed-guideline-values-27-pesticides-used-in-the-gbr-catchment (2017).

    36.
    King, O., Smith, R., Warne, M. St. J. & Mann, R. Proposed aquatic ecosystem protection guideline values for pesticides commonly used in the Great Barrier Reef catchment area: Part 2: Bromacil, Chlorothalonil, Fipronil, Fluometuron, Fluroxypyr, Haloxyfop, MCPA, Pendimethalin, Prometryn, Propazine, Propiconazole, Terbutryn, Triclopyr and Terbuthylazine. Department of Science, Information Technology and Innovation, Brisbane, Australia. https://www.publications.qld.gov.au/dataset/proposed-guideline-values-27-pesticides-used-in-the-gbr-catchment (2017).

    37.
    Fleeger, J. W., Carman, K. R. & Nisbet, R. M. Indirect effects of contaminants in aquatic ecosystems. Sci. Total Environ. 317, 207–233. https://doi.org/10.1016/S0048-9697(03)00141-4 (2003).
    ADS  CAS  Article  Google Scholar 

    38.
    Ralph, P. J. & Gademann, R. Rapid light curves: a powerful tool to assess photosynthetic activity. Aquat. Bot. 82, 222–237. https://doi.org/10.1016/j.aquabot.2005.02.006 (2005).
    CAS  Article  Google Scholar 

    39.
    Schreiber, U. Pulse-amplitude-modulation (PAM) fluorometry and saturation pulse method: an overview. in Chlorophyll a fluorescence (Springer, Dordrecht, 2004) 279–319.

    40.
    Magnusson, M., Heimann, K. & Negri, A. P. Comparative effects of herbicides on photosynthesis and growth of tropical estuarine microalgae. Mar. Pollut. Bull. 56, 1545–1552. https://doi.org/10.1016/j.marpolbul.2008.05.023 (2008).
    CAS  Article  Google Scholar 

    41.
    Sjollema, S. B. et al. Hazard and risk of herbicides for marine microalgae. Environ. Pollut. 187, 106–111. https://doi.org/10.1016/j.envpol.2013.12.019 (2014).
    CAS  Article  Google Scholar 

    42.
    Muller, R. et al. Rapid exposure assessment of PSII herbicides in surface water using a novel chlorophyll a fluorescence imaging assay. Sci. Total Environ. 401, 51–59. https://doi.org/10.1016/j.scitotenv.2008.02.062 (2008).
    ADS  CAS  Article  Google Scholar 

    43.
    Bengston-Nash, S. M., Quayle, P. A., Schreiber, U. & Muller, J. F. The selection of a model microalgal species as biomaterial for a novel aquatic phytotoxicity assay. Aquat. Toxicol. 72, 315–326. https://doi.org/10.1016/j.aquatox.2005.02.004 (2005).
    CAS  Article  Google Scholar 

    44.
    Duggleby, R. G., McCourt, J. A. & Guddat, L. W. Structure and mechanism of inhibition of plant acetohydroxyacid synthase. Plant Physiol. Biochem. 46, 309–324. https://doi.org/10.1016/j.plaphy.2007.12.004 (2008).
    CAS  Article  Google Scholar 

    45.
    Grossmann, K. Auxin herbicides: current status of mechanism and mode of action. Pest Manage. Sci. 66, 113–120. https://doi.org/10.1002/ps.1860 (2010).
    CAS  Article  Google Scholar 

    46.
    OECD. Organisation for Economic Cooperation and Development (OECD) guidelines for the testing of chemicals: freshwater alga and cyanobacteria, growth inhibition test. Test No. 201, https://search.oecd.org/env/test-no-201-alga-growth-inhibition-test-9789264069923-en.htm (2011).

    47.
    Lewis, K. A., Tzilivakis, J., Warner, D. J. & Green, A. An international database for pesticide risk assessments and management. Hum. Ecol. Risk Assess. Int. J. 22, 1050–1064. https://doi.org/10.1080/10807039.2015.1133242 (2016).
    CAS  Article  Google Scholar 

    48.
    Rutherford, A. W. & Krieger-Liszkay, A. Herbicide-induced oxidative stress in photosystem II. Trends Biochem. Sci. 26, 648–653. https://doi.org/10.1016/S0968-0004(01)01953-3 (2001).
    CAS  Article  Google Scholar 

    49.
    Chen, S., Yin, C., Strasser, R. J., Yang, C. & Qiang, S. Reactive oxygen species from chloroplasts contribute to 3-acetyl-5-isopropyltetramic acid-induced leaf necrosis of Arabidopsis thaliana. Plant Physiol. Biochem. 52, 38–51. https://doi.org/10.1016/j.plaphy.2011.11.004 (2012).
    CAS  Article  Google Scholar 

    50.
    Chesworth, J., Donkin, M. & Brown, M. The interactive effects of the antifouling herbicides Irgarol 1051 and Diuron on the seagrass Zostera marina (L.). Aquat. Toxicol. 66, 293–305. https://doi.org/10.1016/j.aquatox.2003.10.002 (2004).
    CAS  Article  Google Scholar 

    51.
    Jones, R. J. & Kerswell, A. P. Phytotoxicity of photosystem II (PSII) herbicides to coral. Mar. Ecol. Prog. Ser. 261, 149–159. https://doi.org/10.3354/meps261149 (2003).
    ADS  CAS  Article  Google Scholar 

    52.
    van Dam, J. W., Negri, A. P., Mueller, J. F. & Uthicke, S. Symbiont-specific responses in foraminifera to the herbicide diuron. Mar. Pollut. Bull. 65, 373–383. https://doi.org/10.1016/j.marpolbul.2011.08.008 (2012).
    CAS  Article  Google Scholar 

    53.
    Negri, A. P., Flores, F., Röthig, T. & Uthicke, S. Herbicides increase the vulnerability of corals to rising sea surface temperature. Limnol. Oceanogr. 56, 471–485. https://doi.org/10.4319/lo.2011.56.2.0471 (2011).
    ADS  CAS  Article  Google Scholar 

    54.
    USEPA. ECOTOX User Guide: ECOTOXicology Database System. Version 5.0. United States Environmental Protection Agency. https://cfpub.epa.gov/ecotox/ (2019).

    55.
    Bao, V. W., Leung, K. M., Qiu, J.-W. & Lam, M. H. Acute toxicities of five commonly used antifouling booster biocides to selected subtropical and cosmopolitan marine species. Mar. Pollut. Bull. 62, 1147–1151. https://doi.org/10.1016/j.marpolbul.2011.02.041 (2011).
    CAS  Article  Google Scholar 

    56.
    Gatidou, G., Thomaidis, N. S. & Zhou, J. L. Fate of Irgarol 1051, diuron and their main metabolites in two UK marine systems after restrictions in antifouling paints. Environ. Int. 33, 70–77. https://doi.org/10.1016/j.envint.2006.07.002 (2007).
    CAS  Article  Google Scholar 

    57.
    Jung, S. et al. Acute toxicity of organic antifouling biocides to phytoplankton Nitzschia pungens and zooplankton Artemia larvae. Mar. Pollut. Bull. 124, 811–818. https://doi.org/10.1016/j.marpolbul.2016.11.047 (2017).
    CAS  Article  Google Scholar 

    58.
    Koutsaftis, A. & Aoyama, I. The interactive effects of binary mixtures of three antifouling biocides and three heavy metals against the marine algae Chaetoceros gracilis. Environ. Toxicol. Int. J. 21, 432–439. https://doi.org/10.1002/tox.20202 (2006).
    ADS  CAS  Article  Google Scholar 

    59.
    Booij, P. et al. Identification of photosynthesis inhibitors of pelagic marine algae using 96-well plate microfractionation for enhanced throughput in effect-directed analysis. Environ. Sci. Technol. 48, 8003–8011. https://doi.org/10.1021/es405428t (2014).
    ADS  CAS  Article  Google Scholar 

    60.
    DeLorenzo, M. E., Danese, L. E. & Baird, T. D. Influence of increasing temperature and salinity on herbicide toxicity in estuarine phytoplankton. Environ. Toxicol. 28, 359–371. https://doi.org/10.1002/tox.20726 (2013).
    ADS  CAS  Article  Google Scholar 

    61.
    Devilla, R. A. et al. Impact of antifouling booster biocides on single microalgal species and on a natural marine phytoplankton community. Mar. Ecol. Prog. Ser. 286, 1–12. https://doi.org/10.3354/MEPS286001 (2005).
    ADS  CAS  Article  Google Scholar 

    62.
    Mercurio, P. et al. Contribution of transformation products towards the total herbicide toxicity to tropical marine organisms. Sci. Rep. 8, 4808. https://doi.org/10.1038/s41598-018-23153-4 (2018).
    ADS  CAS  Article  Google Scholar 

    63.
    Jones, R. The ecotoxicological effects of Photosystem II herbicides on corals. Mar. Pollut. Bull. 51, 495–506. https://doi.org/10.1016/j.marpolbul.2005.06.027 (2005).
    CAS  Article  Google Scholar 

    64.
    Guasch, H. & Sabater, S. Light history influences the sensitivity to atrazine in periphytic algae. J. Phycol. 34, 233–241. https://doi.org/10.1046/j.1529-8817.1998.340233.x (1998).
    CAS  Article  Google Scholar 

    65.
    Millie, D. F., Hersh, C. M. & Dionigi, C. P. Simazine-induced inhibition in photoacclimated populations of Anabaena circinalis (Cyanophyta). J. Phycol. 28, 19–26. https://doi.org/10.1111/j.0022-3646.1992.00019.x (1992).
    CAS  Article  Google Scholar 

    66.
    Bérard, A. et al. Comparison of the ecotoxicological impact of the triazines Irgarol 1051 and atrazine on microalgal cultures and natural microalgal communities in Lake Geneva. Chemosphere 53, 935–944. https://doi.org/10.1016/S0045-6535(03)00674-X (2003).
    ADS  CAS  Article  Google Scholar 

    67.
    Descolas-Gros, C. & Oriol, L. Variations in carboxylase activity in marine phytoplankton cultures. ß-carboxylation in carbon flux studies. Mar. Ecol. Prog. Ser. 85, 163–169 (1992).
    ADS  CAS  Article  Google Scholar 

    68.
    Tang, J., Hoagland, K. D. & Siegfried, B. D. Uptake and bioconcentration of atrazine by selected freshwater algae. Environ. Toxicol. Chem. 17, 1085–1090. https://doi.org/10.1002/etc.5620170614 (1998).
    CAS  Article  Google Scholar 

    69.
    Magnusson, M., Heimann, K., Ridd, M. & Negri, A. P. Chronic herbicide exposures affect the sensitivity and community structure of tropical benthic microalgae. Mar. Pollut. Bull. 65, 363–372. https://doi.org/10.1016/j.marpolbul.2011.09.029 (2012).
    CAS  Article  Google Scholar 

    70.
    Tuchman, N. C., Schollett, M. A., Rier, S. T. & Geddes, P. Differential heterotrophic utilization of organic compounds by diatoms and bacteria under light and dark conditions. Hydrobiologia 561, 167–177. https://doi.org/10.1007/s10750-005-1612-4 (2006).
    CAS  Article  Google Scholar 

    71.
    APVMA. Australian Pesticides and Veterinary Medicines Authority. https://apvma.gov.au/ (2019).

    72.
    EPA. U.S. Environmental Protection Agency. https://www.epa.gov/pesticides (2020).

    73.
    EC. European Commission. EU Pesticides database. https://ec.europa.eu/food/plant/pesticides/eu-pesticides-database/ (2020).

    74.
    Novic, A. J. et al. Monitoring herbicide concentrations and loads during a flood event: a comparison of grab sampling with passive sampling. Environ. Sci. Technol. 51, 3880–3891. https://doi.org/10.1021/acs.est.6b02858 (2017).
    ADS  CAS  Article  Google Scholar 

    75.
    Mercurio, P. Herbicide persistence and toxicity in the tropical marine environment. PhD University of Queensland. 148 p. https://doi.org/10.14264/uql.2016.722 (2016).

    76.
    MacBean, C. The pesticide manual: a world compendium, 6th Edition 598–601 (British Crop Production Council (BCPC), Alton, 2012).

    77.
    Huerlimann, R. & Heimann, K. Comprehensive guide to acetyl-carboxylases in algae. Crit. Rev. Biotechnol. 33, 49–65. https://doi.org/10.3109/07388551.2012.668671 (2013).
    CAS  Article  Google Scholar 

    78.
    Kukorelli, G., Reisinger, P. & Pinke, G. ACCase inhibitor herbicides – selectivity, weed resistance and fitness cost: a review. Int. J. Pest Manage. 59, 165–173. https://doi.org/10.1080/09670874.2013.821212 (2013).
    CAS  Article  Google Scholar 

    79.
    Huerlimann, R., Zenger, K. R., Jerry, D. R. & Heimann, K. Phylogenetic analysis of nucleus-encoded acetyl-CoA carboxylases targeted at the cytosol and plastid of algae. PLoS ONE https://doi.org/10.1371/journal.pone.0131099 (2015).
    Article  Google Scholar 

    80.
    Tang, C. Y., Huang, Z. & Allen, H. C. Interfacial water structure and effects of Mg2+ and Ca2+ binding to the COOH headgroup of a palmitic acid monolayer studied by sum frequency spectroscopy. J. Phys. Chem. B 115, 34–40. https://doi.org/10.1021/jp1062447 (2010).
    CAS  Article  Google Scholar 

    81.
    Brzozowska, A., Duits, M. H. & Mugele, F. Stability of stearic acid monolayers on Artificial Sea Water. Colloids Surf. Physicochem. Eng. Aspects 407, 38–48. https://doi.org/10.1016/j.colsurfa.2012.04.055 (2012).
    CAS  Article  Google Scholar 

    82.
    Bengston-Nash, S. M., Schreiber, U., Ralph, P. J. & Muller, J. F. The combined SPE : ToxY-PAM phytotoxicity assay; application and appraisal of a novel biomonitoring tool for the aquatic environment. Biosens. Bioelectron. 20, 1443–1451. https://doi.org/10.1016/j.bios.2004.09.019 (2005).
    CAS  Article  Google Scholar 

    83.
    Schreiber, U., Quayle, P., Schmidt, S., Escher, B. I. & Mueller, J. F. Methodology and evaluation of a highly sensitive algae toxicity test based on multiwell chlorophyll fluorescence imaging. Biosens. Bioelectron. 22, 2554–2563. https://doi.org/10.1016/j.bios.2006.10.018 (2007).
    CAS  Article  Google Scholar 

    84.
    Haynes, D., Muller, J. & Carter, S. Pesticide and herbicide residues in sediments and seagrasses from the Great Barrier Reef World Heritage Area and Queensland coast. Mar. Pollut. Bull. 41, 279–287. https://doi.org/10.1016/s0025-326x(00)00097-7 (2000).
    CAS  Article  Google Scholar 

    85.
    Ralph, P., Smith, R., Macinnis-Ng, C. & Seery, C. Use of fluorescence-based ecotoxicological bioassays in monitoring toxicants and pollution in aquatic systems. Toxicol. Environ. Chem. 89, 589–607. https://doi.org/10.1080/02772240701561593 (2007).
    CAS  Article  Google Scholar 

    86.
    Lemmermann, E. D. grosse Waterneverstorfer Binnensee: Eine biologische Studie. Forsch. Biol. Station Plön 6, 166–205 (1896).
    Google Scholar 

    87.
    Li, Y. et al. Diversity in the globally distributed diatom genus Chaetoceros (Bacillariophyceae): three new species from warm-temperate waters. PLoS ONE https://doi.org/10.1371/journal.pone.0168887 (2017).
    Article  Google Scholar 

    88.
    Helm, M. M., and Neil Bourne. Hatchery culture of bivalves: a practical manual. Ed. Alessandro Lovatelli. Fisheries Technical Paper 471. Food and Agriculture Organization of the United (FAO), 177 pp (2004).

    89.
    Guillard, R. R. & Ryther, J. H. Studies of marine planktonic diatoms: I Cyclotellanana Hustedt, and Detonulaconfervacea (Cleve) Gran. Can. J. Microbiol. 8, 229–239. https://doi.org/10.1139/m62-029 (1962).
    CAS  Article  Google Scholar 

    90.
    Schreiber, U., Müller, J. F., Haugg, A. & Gademann, R. New type of dual-channel PAM chlorophyll fluorometer for highly sensitive water toxicity biotests. Photosynth. Res. 74, 317–330. https://doi.org/10.1023/A:1021276003145 (2002).
    CAS  Article  Google Scholar 

    91.
    Fisher, R., Ricardo, G., and Fox, D. jags NEC: A Bayesian No Effect Concentration (NEC) package. https://github.com/AIMS/NEC-estimation (2019).

    92.
    Fox, D. R. A Bayesian approach for determining the no effect concentration and hazardous concentration in ecotoxicology. Ecotoxicol. Environ. Saf. 73, 123–131. https://doi.org/10.1016/j.ecoenv.2009.09.012 (2010).
    CAS  Article  Google Scholar  More

  • in

    Plant resistome profiling in evolutionary old bog vegetation provides new clues to understand emergence of multi-resistance

    1.
    World Health Organization. Antimicrobial resistance: global report on surveillance. 2014.
    2.
    Kåhrström CT. Entering a post-antibiotic era? Nat Rev Microbiol. 2013;11:146.
    PubMed  PubMed Central  Google Scholar 

    3.
    Martínez JL. Antibiotics and antibiotic resistance genes in natural environments. Science. 2008;321:365–7.
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    4.
    Pal C, Bengtsson-Palme J, Kristiansson E, Larsson DGJ. The structure and diversity of human, animal and environmental resistomes. Microbiome. 2016;4:1–15.
    CAS  Article  Google Scholar 

    5.
    Chen Q-L, Cui H-L, Su J-Q, Penuelas J, Zhu Y-G. Antibiotic resistomes in plant microbiomes. Trends Plant Sci. 2019;24:530–41.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    6.
    Cernava T, Erlacher A, Soh J, Sensen CW, Grube M, Berg G. Enterobacteriaceae dominate the core microbiome and contribute to the resistome of arugula (Eruca sativa Mill.). Microbiome. 2019;7:13.
    PubMed  PubMed Central  Article  Google Scholar 

    7.
    Blau K, Jacquiod S, Sørensen SJ, Su J-Q, Zhu Y-G, Smalla K, et al. Manure and doxycycline affect the bacterial community and its resistome in lettuce rhizosphere and bulk soil. Front Microbiol. 2019;10:725.
    PubMed  PubMed Central  Article  Google Scholar 

    8.
    Berendonk TU, Manaia CM, Merlin C, Fatta-Kassinos D, Cytryn E, Walsh F, et al. Tackling antibiotic resistance: the environmental framework. Nat Rev Microbiol. 2015;13:310–7.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    9.
    Wright GD. The antibiotic resistome: the nexus of chemical and genetic diversity. Nat Rev Microbiol. 2007;5:175–86.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    10.
    Page SE, Baird AJ. Peatlands and global change: response and resilience. Annu Rev Environ Resour. 2016;41:35–57.
    Article  Google Scholar 

    11.
    Kostka JE, Weston DJ, Glass JB, Lilleskov EA, Shaw AJ, Turetsky MR. The Sphagnum microbiome: new insights from an ancient plant lineage. N Phytologist. 2016;211:57–64.
    CAS  Article  Google Scholar 

    12.
    Bragina A, Oberauner-Wappis L, Zachow C, Halwachs B, Thallinger GG, Müller H, et al. The Sphagnum microbiome supports bog ecosystem functioning under extreme conditions. Mol Ecol. 2014;23:4498–510.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    13.
    Opelt K, Chobot V, Hadacek F, Schönmann S, Eberl L, Berg G. Investigations of the structure and function of bacterial communities associated with Sphagnum mosses. Environ Microbiol. 2007;9:2795–809.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    14.
    Bragina A, Berg C, Cardinale M, Shcherbakov A, Chebotar V, Berg G. Sphagnum mosses harbour highly specific bacterial diversity during their whole lifecycle. ISME J. 2012;6:802–13.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    15.
    Opelt K, Berg C, Schönmann S, Eberl L, Berg G. High specificity but contrasting biodiversity of Sphagnum-associated bacterial and plant communities in bog ecosystems independent of the geographical region. ISME J. 2007;1:502–16.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    16.
    Bragina A, Berg C, Berg G. The core microbiome bonds the Alpine bog vegetation to a transkingdom metacommunity. Mol Ecol. 2015;24:4795–807.
    PubMed  Article  PubMed Central  Google Scholar 

    17.
    Opelt K, Berg C, Berg G. The bryophyte genus Sphagnum is a reservoir for powerful and extraordinary antagonists and potentially facultative human pathogens. FEMS Microbiol Ecol. 2007;61:38–53.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    18.
    Allen HK, Donato J, Wang HH, Cloud-Hansen KA, Davies J, Handelsman J. Call of the wild: antibiotic resistance genes in natural environments. Nat Rev Microbiol. 2010;8:251–9.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    19.
    Woodcroft BJ, Singleton CM, Boyd JA, Evans PN, Emerson JB, Zayed AAF, et al. Genome-centric view of carbon processing in thawing permafrost. Nature. 2018;560:49–54.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    20.
    Mondav R, Woodcroft BJ, Kim EH, Mccalley CK, Hodgkins SB, Crill PM, et al. Discovery of a novel methanogen prevalent in thawing permafrost. Nat Commun. 2014;5:1–7.
    Article  CAS  Google Scholar 

    21.
    Müller CA, Oberauner-Wappis L, Peyman A, Amos GCA, EMH Wellington, Berg G. Mining for NRPS and PKS genes revealed a high diversity in the Sphagnum bog metagenome. Appl Environ Microbiol. 2015;81:5064–72.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    22.
    Bragina A, Maier S, Berg C, Müller H, Chobot V, Hadacek F, et al. Similar diversity of Alphaproteobacteria and nitrogenase gene amplicons on two related Sphagnum mosses. Front Microbiol. 2011;2:275.
    PubMed  PubMed Central  Google Scholar 

    23.
    Wardwell LH, Jude BA, Moody JP, Olcerst AI, Gyure RA, Nelson RE, et al. Co-selection of mercury and antibiotic resistance in sphagnum core samples dating back 2000 years. Geomicrobiol J. 2009;26:351–60.
    CAS  Article  Google Scholar 

    24.
    McInnes L, Healy J, Melville J. UMAP: Uniform manifold approximation and projection for dimension reduction. arXiv:1802.03426. 2020.

    25.
    Becht E, McInnes L, Healy J, Dutertre CA, Kwok IWH, Ng LG, et al. Dimensionality reduction for visualizing single-cell data using UMAP. Nat Biotechnol. 2019;37:38–47.
    CAS  Article  Google Scholar 

    26.
    Breiman L. Random forests. Mach Learn. 2001;45:5–32.

    27.
    Altmann A, Toloşi L, Sander O, Lengauer T. Permutation importance: a corrected feature importance measure. Bioinformatics. 2010;26:1340–7.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    28.
    Streit M, Lex A, Gratzl S, Partl C, Schmalstieg D, Pfister H, et al. Guided visual exploration of genomic stratifications in cancer. Nat Methods. 2014;11:884–5.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    29.
    McArthur AG, Waglechner N, Nizam F, Yan A, Azad MA, Baylay AJ, et al. The comprehensive antibiotic resistance database. Antimicrobial agents Chemother. 2013;57:3348–57.
    CAS  Article  Google Scholar 

    30.
    Buchfink B, Xie C, Huson DH. Fast and sensitive protein alignment using DIAMOND. Nat Methods. 2015;12:59–60.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    31.
    Elbehery AHA, Aziz RK, Siam R. Antibiotic resistome: improving detection and quantification accuracy for comparative metagenomics. OMICS: A J Integr Biol. 2016;20:229–38.
    CAS  Article  Google Scholar 

    32.
    Boyd JA, Woodcroft BJ, Tyson GW. GraftM: a tool for scalable, phylogenetically informed classification of genes within metagenomes. Nucleic Acids Res. 2018;46:e59.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    33.
    Allaire J. RStudio: Integrated development environment for R. Boston, MA: RStudio Inc.; 2012. https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6194118/#!po=0.819672.

    34.
    Core Team R. R: a language and environment for statistical computing. R Foundation for statistical computing, Vienna. 2013.

    35.
    McMurdie PJ, Holmes S. phyloseq: an R package for reproducible interactive analysis and graphics of microbiome census data. PloS ONE. 2013;8:e61217.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    36.
    Oksanen J, Kindt R, Legendre P, O’Hara B, Stevens MHH, Oksanen MJ, et al. The vegan package. Commun Ecol Package. 2007;10:631–7.
    Google Scholar 

    37.
    Mantel N. The detection of disease clustering and a generalized regression approach. Cancer Res. 1967;27:209–20.
    Google Scholar 

    38.
    Li D, Liu C-M, Luo R, Sadakane K, Lam T-W. MEGAHIT: an ultra-fast single-node solution for large and complex metagenomics assembly via succinct de Bruijn graph. Bioinformatics. 2015;31:1674–6.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    39.
    Alneberg J, Bjarnason BS, De Bruijn I, Schirmer M, Quick J, Ijaz UZ, et al. Binning metagenomic contigs by coverage and composition. Nat Methods. 2014;11:1144–6.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    40.
    Kang DD, Li F, Kirton E, Thomas A, Egan R, An H, et al. MetaBAT 2: an adaptive binning algorithm for robust and efficient genome reconstruction from metagenome assemblies. PeerJ. 2019;7:e7359.
    PubMed  PubMed Central  Article  Google Scholar 

    41.
    Wu Y-W, Simmons BA, Singer SW. MaxBin 2.0: an automated binning algorithm to recover genomes from multiple metagenomic datasets. Bioinformatics. 2016;32:605–7.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    42.
    Sieber CM, Probst AJ, Sharrar A, Thomas BC, Hess M, Tringe SG, et al. Recovery of genomes from metagenomes via a dereplication, aggregation and scoring strategy. Nat Microbiol. 2018;3:836–43.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    43.
    Parks DH, Imelfort M, Skennerton CT, Hugenholtz P, Tyson GW. CheckM: assessing the quality of microbial genomes recovered from isolates, single cells, and metagenomes. Genome Res. 2015;25:1043–55.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    44.
    Bowers RM, Kyrpides NC, Stepanauskas R, Harmon-Smith M, Doud D, Reddy T, et al. Minimum information about a single amplified genome (MISAG) and a metagenome-assembled genome (MIMAG) of bacteria and archaea. Nat Biotechnol. 2017;35:725–31.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    45.
    Buchfink B, Xie C, Huson DH. Fast and sensitive protein alignment using DIAMOND. Nat Methods. 2015;12:59.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    46.
    Arango-Argoty G, Garner E, Pruden A, Heath LS, Vikesland P, Zhang L. DeepARG: a deep learning approach for predicting antibiotic resistance genes from metagenomic data. Microbiome. 2018;6:1–15.
    Article  Google Scholar 

    47.
    Huerta-Cepas J, Serra F, Bork P. ETE 3: reconstruction, analysis, and visualization of phylogenomic data. Mol Biol Evol. 2016;33:1635–8.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    48.
    Szklarczyk D, Gable AL, Lyon D, Junge A, Wyder S, Huerta-Cepas J, et al. STRING v11: protein–protein association networks with increased coverage, supporting functional discovery in genome-wide experimental datasets. Nucleic Acids Res. 2019;47:D607–13.

    49.
    Mauri M, Elli T, Caviglia G, Uboldi G, Azzi M. RAWGraphs: a visualisation platform to create open outputs. In: Proceedings of the 12th Biannual Conference on Italian SIGCHI Chapter. New York, New York, USA: ACM Press; 2017. p. 28:1–28:5.

    50.
    Shannon P, Markiel A, Ozier O, Baliga NS, Wang JT, Ramage D, et al. Cytoscape: a software environment for integrated models of biomolecular interaction networks. Genome Res. 2003;13:2498–504.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    51.
    Vercammen K, Garcia-Armisen T, Goeders N, Van Melderen L, Bodilis J, Cornelis P. Identification of a metagenomic gene cluster containing a new class A beta-lactamase and toxin-antitoxin systems. MicrobiologyOpen. 2013;2:674–83.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    52.
    Allen HK, Moe LA, Rodbumrer J, Gaarder A, Handelsman J.Functional metagenomics reveals diverse β-lactamases in a remote Alaskan soil. ISME J. 2009;3:243–51.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    53.
    European Committee for Antimicrobial Susceptibility Testing (EUCAST) of the European Society of Clinical Microbiology and Infectious Diseases (ESCMID). Determination of minimum inhibitory concentrations (MICs) of antibacterial agents by broth dilution. Clin Microbiol Infect. 2003;9:1–7.
    Google Scholar 

    54.
    Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics. 2014;30:2114–20.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    55.
    Bankevich A, Nurk S, Antipov D, Gurevich AA, Dvorkin M, Kulikov AS, et al. SPAdes: a new genome assembly algorithm and its applications to single-cell sequencing. J Comput Biol. 2012;19:455–77.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    56.
    García-Alcalde F, Okonechnikov K, Carbonell J, Cruz LM, Götz S, Tarazona S, et al. Qualimap: evaluating next-generation sequencing alignment data. Bioinformatics. 2012;28:2678–9.
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    57.
    Seemann T. Prokka: rapid prokaryotic genome annotation. Bioinformatics. 2014;30:2068–9.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    58.
    Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol. 2018;35:1547–9.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    59.
    Edgar RC. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids. Res 2004;32:1792–7.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    60.
    Saitou N, Nei M. The neighbor-joining method: a new method for reconstructing phylogenetic trees. Mol Biol Evol. 1987;4:406–25.
    CAS  PubMed  PubMed Central  Google Scholar 

    61.
    Nei M, Kumar S. Molecular Evolution and Phylogenetics. New York: Oxford University Press; 2000.

    62.
    Collignon PC, Conly JM, Andremont A, McEwen SA, Aidara-Kane A, Agerso Y, et al. World Health Organization ranking of antimicrobials according to their importance in human medicine: a critical step for developing risk management strategies to control antimicrobial resistance from food animal production. Clin Infect Dis. 2016;63:1087–93.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    63.
    Philippon A, Slama P, Dény P, Labia R. A structure-based classification of class A β-lactamases, a broadly diverse family of enzymes. Clin Microbiol Rev. 2016;29:29–57.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    64.
    Ambler RP. The structure of β-lactamases. Philos Trans R Soc B: Biol Sci. 1980;289:321–31.
    CAS  Google Scholar 

    65.
    D’Costa VM, McGrann KM, Hughes DW, Wright GD. Sampling the antibiotic resistome. Science. 2006;311:374–7.
    PubMed  Article  PubMed Central  Google Scholar 

    66.
    Forsberg KJ, Patel S, Wencewicz TA, Dantas G. Bacterial phylogeny structures soil resistome across habitats. Nature. 2014;509:612–6.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    67.
    Van Goethem MW, Pierneef R, Bezuidt OKI, Van De Peer Y, Cowan DA, Makhalanyane TP. A reservoir of ´historical´antibiotic resistance genes in remote pristine Antarctic soils. Microbiome. 2018;6:40.
    PubMed  PubMed Central  Article  Google Scholar 

    68.
    Mahnert A, Moissl-Eichinger C, Zojer M, Bogumil D, Mizrahi I, Rattei T, et al. Man-made microbial resistances in built environments. Nat Commun. 2019;10:968.
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    69.
    Carr VR, Witherden EA, Lee S, Shoaie S, Mullany P, Proctor GB, et al. Abundance and diversity of resistomes differ between healthy human oral cavities and gut. Nat Commun. 2020;11:693.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    70.
    Bragina A, Cardinale M, Berg C, Berg G, Schmid M, Zentrum H. Vertical transmission explains the specific Burkholderia pattern in Sphagnum mosses at multi-geographic scale. Front Microbiol. 2013;4:394.
    PubMed  PubMed Central  Article  Google Scholar 

    71.
    Crofts TS, Gasparrini AJ, Dantas G. Next-generation approaches to understand and combat the antibiotic resistome. Nat Rev Microbiol. 2017;15:422.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    72.
    Tyson GH, McDermott PF, Li C, Chen Y, Tadesse DA, Mukherjee S, et al. WGS accurately predicts antimicrobial resistance in Escherichia coli. J Antimicrobial Chemother. 2015;70:2763–9.
    CAS  Article  Google Scholar 

    73.
    Schmidt K, Mwaigwisya S, Crossman L, Doumith M, Munroe D, Pires C, et al. Identification of bacterial pathogens and antimicrobial resistance directly from clinical urines by nanopore-based metagenomic sequencing. J Antimicrobial Chemother. 2016;72:104–14.
    Article  CAS  Google Scholar 

    74.
    Andersen H, Connolly N, Bangar H, Staat M, Mortensen J, Deburger B, et al. Use of shotgun metagenome sequencing to detect fecal colonization with multidrug-resistant bacteria in children. J Clin Microbiol. 2016;54:1804–13.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    75.
    Martinez JL, Sánchez MB, Martínez-Solano L, Hernandez A, Garmendia L, Fajardo A, et al. Functional role of bacterial multidrug efflux pumps in microbial natural ecosystems. FEMS Microbiol Rev. 2009;33:430–49.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    76.
    Pehrsson EC, Forsberg KJ, Gibson MK, Ahmadi S, Dantas G. Novel resistance functions uncovered using functional metagenomic investigations of resistance reservoirs. Front Microbiol. 2013;4:1–11.
    Article  Google Scholar 

    77.
    Leclercq R. Mechanisms of resistance to macrolides and lincosamides: nature of the resistance elements and their clinical implications. Clin Infect Dis. 2002;34:482–92.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    78.
    Belova SE, Pankratov TA, Dedysh SN. Bacteria of the genus Burkholderia as a typical component of the microbial community of Sphagnum peat bogs. Microbiology. 2006;75:90–96.
    CAS  Article  Google Scholar 

    79.
    Le Fléche-MatéosA, Kügler JH, Hansen SH, Syldatk C, Hausmann R, Lomprez F, et al. Rouxiella badensis sp. nov. and Rouxiella silvae sp. nov. isolated from peat bog soil and emendation description of the genus Rouxiella. Int J Syst Evolut Microbiol. 2017;67:1255–9.
    Article  CAS  Google Scholar 

    80.
    Coenye T, Falsen E, Hoste B, Ohle M, Goris J, Govan JRW, et al. Description of Pandoraea gen. nov. with Pandoraea apista sp. nov., Pandoraea pulmonicola sp. nov., Pandoraea pnomenusa sp. nov., Pandoraea sputorum sp. nov. and Pandoraea norimbergensis comb. nov 2000;50:887–99.
    CAS  Google Scholar 

    81.
    Green H, Jones AM. Emerging Gram-negative bacteria: pathogenic or innocent bystanders. Curr Opin Pulm Med. 2018;24:592–8.
    PubMed  Article  PubMed Central  Google Scholar 

    82.
    Schneider I, Queenan AM, Bauernfeind A. Novel carbapenem-hydrolyzing oxacillinase OXA-62 from Pandoraea pnomenusa. Antimicrobial Agents Chemother. 2006;50:1330–5.
    CAS  Article  Google Scholar 

    83.
    Mahlen SD. Serratia infections: from military experiments to current practice. Clin Microbiol Rev. 2011;24:755–91.
    PubMed  PubMed Central  Article  Google Scholar 

    84.
    Silby MW, Winstanley C, Godfrey SAC, Levy SB, Jackson RW. Pseudomonas genomes: Diverse and adaptable. FEMS Microbiol Rev. 2011;35:652–80.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    85.
    Mitter B, Pfaffenbichler N, Flavell R, Compant S, Antonielli L, Petric A, et al. A new approach to modify plant microbiomes and traits by introducing beneficial bacteria at flowering into progeny seeds. Front Microbiol. 2017;8:11.
    PubMed  PubMed Central  Google Scholar 

    86.
    Vandamme P, Opelt K, Knochel N, Berg C, Schonmann S, De Brandt E, et al. Burkholderia bryophila sp. nov. and Burkholderia megapolitana sp. nov., moss-associated species with antifungal and plant-growth-promoting properties. Int J Syst Evolut Microbiol. 2007;57:2228–35.
    CAS  Article  Google Scholar 

    87.
    Berg G, Eberl L, Hartmann A. The rhizosphere as a reservoir for opportunistic human pathogenic bacteria. Environ Microbiol. 2005;7:1673–85.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    88.
    Lycus P, Lovise Bøthun K, Bergaust L, Peele Shapleigh J, Reier Bakken L, Frostegård Å. Phenotypic and genotypic richness of denitrifiers revealed by a novel isolation strategy. ISME J. 2017;11:2219–32.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    89.
    Bai Y, Müller DB, Srinivas G, Garrido-Oter R, Potthoff E, Rott M, et al. Functional overlap of the Arabidopsis leaf and root microbiota. Nature. 2015;528:364–9.
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    90.
    Papanicolaou GA, Medeiros AA, Jacoby GA. Novel plasmid-mediated β-lactamase (MIR-1) conferring resistance to oxyimino- and α-methoxy β-lactams in clinical isolates of Klebsiella pneumoniae. Antimicrobial Agents Chemother. 1990;34:2200–9.
    CAS  Article  Google Scholar 

    91.
    Bauernfeind A, Stemplinger I, Jungwirth R, Ernst S, Casellas JM. Sequences of β-lactamase genes encoding CTX-M-1 (MEN-1) and CTX-M-2 and relationship of their amino acid sequences with those of other β-lactamases. Antimicrobial Agents Chemother. 1996;40:509–13.
    CAS  Article  Google Scholar 

    92.
    Dahmen S, Mansour W, Charfi K, Boujaafar N, Arlet G, Bouallègue O. Imipenem resistance in Klebsiella pneumoniae is associated to the combination of plasmid-mediated CMY-4 AmpC β-lactamase and loss of an outer membrane protein. Microb Drug Resistance. 2012;18:479–83.
    CAS  Article  Google Scholar 

    93.
    Forsberg KJ, Patel S, Gibson MK, Lauber CL, Knight R, Fierer N, et al. Bacterial phylogeny structures soil resistomes across habitats. Nature. 2014;509:612–6.
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    94.
    Kurm V, van der Putten WH, de Boer W, Naus-Wiezer S, Hol WHG. Low abundant soil bacteria can be metabolically versatile and fast growing. Ecology. 2017;98:555–64.
    PubMed  Article  PubMed Central  Google Scholar 

    95.
    Vadstein O, Attramadal KJK, Bakke I, Olsen Y. K-Selection as microbial community management strategy: a method for improved viability of larvae in aquaculture. Front Microbiol. 2018;9:2730.
    PubMed  PubMed Central  Article  Google Scholar  More

  • in

    Assessing the climate suitability and potential economic impacts of Oak wilt in Canada

    Species distribution modelling
    A total of 1548 occurrence locations were obtained for B. fagacearum from the United States Forest Service (provided to us by Erin Bullas-Appleton of the Canadian Food Inspection Agency in July 2018) and the Global Biodiversity Information Facility17. These records were filtered at a 300 arcsecond (approximately 10-km) resolution to remove duplicates and reduce spatial clustering, leaving 1401 unique location records (Fig. 1a). Occurrence locations for two key insect vectors of oak wilt, C. truncatus (Fig. 2a) and C. sayi (Fig. 3a), were obtained from GBIF, publications10,18, and from specimens in the following collections: Atlantic Forestry Centre, Fredericton, NB, Canada; Canadian National Collection of Insects, Arachnids, and Nematodes, Agriculture and Agri-Food Canada Research Centre, Ottawa, ON, Canada; Ontario Forest Research Institute, Sault Ste. Marie, ON, Canada; Gareth S. Powell Collection, Nephi, UT, USA; Reginald Webster Collection, Charters Settlement, New Brunswick, Canada; and the Florida State Collection of Arthropods, Gainesville, FL, USA . After filtering at a 10-km resolution, there were 82 and 58 unique occurrence records for C. truncatus and C. sayi respectively.
    Figure 1

    Occurrence data (a) used for generating climate suitability models for Bretziella fagacearum. Maps with colour gradients indicate Maxent-derived climate suitability for B. fagacearum for the: 1981–2010 period (b); 2011–2040 period (c); and 2041–2070 period (d). Stippling delineates the ANUCLIM-derived climate envelope for B. fagacearum in each time period. Hatching delineates the current distribution of Quercus in Canada. Climate projections are based on a composite of four climate models and the RCP 4.5 emissions scenario (see text for further details). Maps were generated using ARCGIS v.9.3 (ESRI, Redlands, CA, USA; https://www.esri.com/arcgis/about-arcgis).

    Full size image

    Figure 2

    Occurrence data (a) used for generating climate suitability models for Colopterus truncatus. Maps with colour gradients indicate Maxent-derived climate suitability for C. truncatus for the: 1981–2010 period (b); 2011–2040 period (c); and 2041–2070 period (d). Stippling delineates the ANUCLIM-derived climate envelope for C. truncatus in each time period. Hatching delineates the current distribution of Quercus in Canada. Climate projections are based on a composite of four climate models and the RCP 4.5 emissions scenario (see text for further details). Maps were generated using ARCGIS v.9.3 (ESRI, Redlands, CA, USA; https://www.esri.com/arcgis/about-arcgis).

    Full size image

    Figure 3

    Occurrence data (a) used for generating climate suitability models for Carpophilus sayi. Maps with colour gradients indicate Maxent-derived climate suitability for C. sayi for the: 1981–2010 period (b); 2011–2040 period (c); and 2041–2070 period (d). Stippling delineates the ANUCLIM-derived climate envelope for C. sayi in each time period. Hatching delineates the current distribution of Quercus in Canada. Climate projections are based on a composite of four climate models and the RCP 4.5 emissions scenario (see text for further details). Maps were generated using ARCGIS v.9.3 (ESRI, Redlands, CA, USA; https://www.esri.com/arcgis/about-arcgis).

    Full size image

    Climate estimates were obtained at each occurrence location by interrogating North American climate models (described in McKenney et al.19) of the 1981–2010 normal period for the following four variables: (1) fall (i.e., September–December) precipitation (FALLPCP); (2) average spring (i.e., March–June) temperature (SPRINGTMP); (3) annual climate moisture index (CMI; a climate-based moisture balance variable; see Hogg20 for details); and (4) Average Minimum Temperature of the Coldest Month (MINTCM). These climate variables were selected based on their reported influences on B. fagacearum (FALLPCP and CMI21,22) C. sayi and C. truncatus (SPRINGTMP10), and insect distributions in general (MINTCM23). Further, none of the selected variables were highly correlated (i.e., r  60 years) for each province in our study area. Focusing on current/near-term oak timber stocks (since we are not estimating B. fagacearum spread), we multiplied merchantable volume over 40 years old—roughly the age at which oak becomes harvestable in Ontario42—by average provincial stumpage values.
    Estimating stumpage values was somewhat challenging due to inter-provincial variation in stumpage systems and reporting of stumpage fees. For the province of Québec, we obtained oak-specific stumpage fees for 191 harvest zones for the period April 1, 2019 to March 31, 2020 (Bureau de mise en marché43). Since stumpage fees in Québec vary by wood quality class (i.e., A, B, and C), we further obtained information on the proportion of wood harvested in each class over the same period (unpublished dataset, Bureau de mise en marché des Bois). We then calculated the average stumpage fee for the province by averaging across harvest zones and quality classes, while weighting by the proportion of wood in each quality class. For Ontario, we obtained stumpage fees for two quality classes of hardwoods (i.e., Class 1 and Class 2) and four oak-related product types (Veneer, Sawlogs, Composite, and Firewood) for January 1 to December 31, 2019 (Ontario Ministry of Natural Resources and Forestry44). Given that oak is typically considered a higher value hardwood, and in lieu of information on how oak is partitioned across product types in Ontario, we calculated oak stumpage fees as an average across product types for the Class 1 hardwood category. Note that the stumpage rates employed here include the Renewal and Forest Futures fees that are part of the Ontario stumpage system. Finally, stumpage values for the province of Nova Scotia were obtained for a single hardwood quality class for the period April 1, 2017 to March 31, 2018 (Province of Nova Scotia45). As in Ontario, stumpage fees were averaged across product types. The stumpage values for Nova Scotia were applied to the relatively small amount of oak in the neighbouring Maritime Provinces of New Brunswick and Prince Edward Island (PEI).
    Alternatively, gross domestic product (GDP) can provide an estimate of the total economic activity associated with a given industry. Annual GDP estimates for broad categories (e.g., forestry and logging industry, and wood product manufacturing industry) are available for each province at Natural Resources Canada’s forestry statistics website (https://cfs.nrcan.gc.ca/statsprofile/overview/ca). In order to estimate GDP specifically for oak-related timber products, we first multiplied these provincial broad category GDP values by the proportion of the total provincial harvest that was composed of hardwoods (multipliers obtained from published provincial data sources as detailed in the Results section below). This value was then further refined by multiplying by the proportion of hardwoods in the province that was composed of oak species. These estimates were obtained from forest attribute grids13, by summing merchantable volume of (1) oak and (2) all broadleaf species within the industrial forestry limits of each province. Spatial summaries were carried out using the raster and rgdal packages in r. Though admittedly coarse, we felt this approach was the best available given the dearth of readily available economic data for individual tree species/genera; similar approaches have been used previously to estimate economic impacts of invasive species46,47.
    These two approaches (i.e., stumpage-based and GDP-based) provide different perspectives on oak-related timber values at risk. The stumpage approach attaches a basic price to standing timber resources, but does not consider downstream economic activities associated with harvest, such as wages, equipment purchases, and capital expenditures. This approach implicitly assumes that substitution possibilities (e.g., other tree species) can fully replace oak-related contributions to the economy with minimal adjustment costs and, as such, is a conservative estimate of potential timber value losses. Alternatively, the GDP approach attempts to include all downstream economic contributions and assumes little or no opportunity for substitution, such that oak timber losses would be accompanied by a proportional reduction in economic activity. We present both estimates here to provide policy-makers with a range of possible impacts. The value of costs through time is generally arrived at using economic discounting; however, here we have no estimates of the timeline associated with oak wilt spread and hence have chosen to report gross, undiscounted values. See Aukema et al.48 for further discussion. More

  • in

    A global population assessment of the Chinstrap penguin (Pygoscelis antarctica)

    1.
    BirdLife International. Pygoscelis antarcticus. The IUCN Red List of Threatened Species 2018:e.T22697761A132601557. https://doi.org/10.2305/IUCN.UK.2018-2.RLTS.T22697761A132601557.en (2018).
    2.
    Sander, M., Balbao, T. C., Polito, M. J., Costa, E. S. & Carneiro, A. P. B. Recent decrease in Chinstrap penguin (Pygoscelis antarctica) populations at two of Admiralty Bay’s islets on King George Island, South Shetland Islands, Antarctica. Polar Biol. 30, 659–661 (2007).
    Article  Google Scholar 

    3.
    Naveen, R., Lynch, H. J., Forrest, S., Mueller, T. & Polito, M. First direct, site-wide penguin survey at Deception Island, Antarctica, suggests significant declines in breeding Chinstrap penguins. Polar Biol. 35(12), 1879–1888 (2012).
    Google Scholar 

    4.
    Lynch, H. J. et al. In stark contrast to widespread declines along the Scotia Arc, a survey of the South Sandwich Islands finds a robust seabird community. Polar Biol. 39, 1–11 (2016).
    ADS  Article  Google Scholar 

    5.
    Dunn, M. et al. Population size and decadal trends of three penguin species nesting at Signy Island South Orkney Islands. PLoS ONE 11, e0164025 (2016).
    Article  Google Scholar 

    6.
    Strycker, N. et al. Fifty-year change in penguin abundance on Elephant Island, South Shetland Islands, Antarctica: Results of the 2019–20 census (2020) (in review).

    7.
    Croxall, J. P. & Furse, J. R. Food of Chinstrap penguins (Pygoscelis antarctica) and Macaroni penguins (Eudyptes chrysolophus) at Elephant Island Group, South Shetland Islands. Ibis 122, 237–245 (1980).
    Article  Google Scholar 

    8.
    Trivelpiece, W. Z. et al. Variability in krill biomass links harvesting and climate warming to penguin population changes in Antarctica. PNAS 108(18), 7625–7628 (2011).
    ADS  CAS  Article  Google Scholar 

    9.
    Lynch, H. J., Naveen, R., Trathan, P. N. & Fagan, W. F. Spatially integrated assessment reveals widespread changes in penguin populations on the Antarctic Peninsula. Ecology 93(6), 1367–1377 (2012).
    Article  Google Scholar 

    10.
    Casanovas, P., Naveen, R., Forrest, S., Poncet, J. & Lynch, H. J. A comprehensive coastal seabird survey maps out the front lines of ecological change on the Western Antarctic Peninsula. Polar Biol. 38, 927–940 (2015).
    Article  Google Scholar 

    11.
    Croxall, J. P., Trathan, P. & Murphy, E. Environmental change and Antarctic seabird populations. Science 297, 1510–1514 (2002).
    ADS  CAS  Article  Google Scholar 

    12.
    CCAMLR. Standard methods for monitoring parameters of predator species. CCAMLR Ecosystem Monitoring Programme, Hobart (2004).

    13.
    Hinke, J., Salwicka, K., Trivelpiece, S., Watters, G. & Trivelpiece, W. Divergent responses of Pygoscelis penguins reveal common environmental driver. Oecologia 153, 845–855 (2007).
    ADS  Article  Google Scholar 

    14.
    Lynch, H. J., Fagan, W. F. & Naveen, R. Population trends and reproductive success at a frequently visited penguin colony on the western Antarctic Peninsula. Polar Biol. 33, 493–503 (2010).
    Article  Google Scholar 

    15.
    Fraser, W. R., Trivelpiece, W. Z., Ainley, D. G. & Trivelpiece, S. G. Increases in Antarctic penguin populations: Reduced competition with whales or a loss of sea ice due to environmental warming?. Polar Biol. 11(8), 525–531 (1992).
    Article  Google Scholar 

    16.
    Hinke, J. et al. Identifying risk: concurrent overlap of the Antarctic krill fishery with krill-dependent predators in the Scotia Sea. PLoS ONE 12, e0170132 (2017).
    Article  Google Scholar 

    17.
    Hill, S. L. et al. Reference points for predators will progress ecosystem-based management of fisheries. Fish Fish. 21(2), 368–378 (2020).
    Article  Google Scholar 

    18.
    Juáres, M. A. et al. Adélie penguin population changes at Stranger Point: 19 years of monitoring. Antarct. Sci. 27(5), 455–461 (2015).
    ADS  Article  Google Scholar 

    19.
    Lynch, M., Youngflesh, C., Agha, N. M., Ottinger, M. A. & Lynch, H. Tourism and stress hormone measures in Gentoo Penguins on the Antarctic Peninsula. Polar Biol. 42, 1299–1306 (2019).
    Article  Google Scholar 

    20.
    Croxall, J. P. & Kirkwood, E. D. The Distribution of Penguins on the Antarctic Peninsula and Islands of the Scotia Sea (British Antarctic Survey, Cambridge, 1979).
    Google Scholar 

    21.
    Woehler, E. J. The Distribution and Abundance of Antarctic and Subantarctic Penguins (Scientific Committee on Antarctic Research, Cambridge, 1993).
    Google Scholar 

    22.
    Conroy, J. W. H., White, M. G., Furse, J. R. & Bruce, G. Observations on the breeding biology of the Chinstrap penguin, Pygoscelis Antarctica, at Elephant Island, South Shetland Islands. Br. Antarct. Surv. Bull. 40, 23–32 (1975).
    Google Scholar 

    23.
    Lowther, A. D. Antarctic marine mammals. Encyclopedia of Marine Mammals 3rd edn, 27–32 (Academic Press, London, 2018).
    Google Scholar 

    24.
    Hinke, J. et al. Spatial and isotopic niche partitioning during winter in Chinstrap and Adélie penguins from the South Shetland Islands. Ecosphere 6, 125 (2015).
    Article  Google Scholar 

    25.
    Dimitrijević, D. et al. Isotopic niches of sympatric Gentoo and Chinstrap penguins: evidence of competition for Antarctic krill?. Polar Biol. 41, 1655–1669 (2018).
    Article  Google Scholar 

    26.
    Trathan, P. N., Croxall, J. P. & Murphy, E. J. Dynamics of Antarctic penguin populations in relation to inter-annual variability in sea ice distribution. Polar Biol. 16, 321–330 (1996).
    Article  Google Scholar 

    27.
    He, H. et al. Aerial photography based census of Adélie penguin and its application in CH4 and N20 budget estimation in Victoria Land, Antarctic. Sci. Rep. 7, 12942 (2017).
    ADS  Article  Google Scholar 

    28.
    Bromwich, D. H. et al. Central West Antarctica among the most rapidly warming regions on Earth. Nat. Geosci. 6, 139–145 (2013).
    ADS  CAS  Article  Google Scholar 

    29.
    World Meteorological Organization. 2020.

    30.
    Flores, H. Impact of climate change on Antarctic krill. Mar. Ecol. Prog. Ser. 458, 1–19 (2012).
    ADS  Article  Google Scholar 

    31.
    Reid, K., Sims, M., White, R. W. & Gillon, K. W. Spatial distribution of predator/prey interactions in the Scotia Sea: Implications for measuring predator/fisheries overlap. Deep Sea Res. Part II 51, 1383–1396 (2004).
    ADS  Article  Google Scholar 

    32.
    Watters, G. M., Hinke, J. & Reiss, C. S. Long-term observations from Antarctica demonstrate that mismatched scales of fisheries management and predator–prey interaction lead to erroneous conclusions about precaution. Sci. Rep. 10, 2314 (2020).
    ADS  CAS  Article  Google Scholar 

    33.
    Polito, M. J. et al. Stable isotope analyses of feather amino acids identify penguin migration strategies at ocean basin scales. Biol. Lett. 13, 20170241 (2017).
    Article  Google Scholar 

    34.
    Hinke, J., Santos, M., Malgorzata Korczak-Abshire, G. M. & Watters, G. Individual variation in migratory movements of Chinstrap penguins leads to widespread occupancy of ice-free winter habitats over the continental shelf and deep ocean basins of the Southern Ocean. PLoS ONE 14(12), e0226207 (2019).
    CAS  Article  Google Scholar 

    35.
    Hurlbert, A. H. & Liang, Z. Spatiotemporal variation in avian migration phenology: citizen science reveals effects of climate change. PLoS ONE 7, e31662 (2012).
    ADS  CAS  Article  Google Scholar 

    36.
    Cohen, J. M., Lajeunesse, J. M. & Rohr, J. R. A global synthesis of animal phenological responses to climate change. Nat. Clim. Change 8, 224–228 (2018).
    ADS  Article  Google Scholar 

    37.
    Samplonius, J. M. et al. Strengthening the evidence base for temperature-mediated phenological asynchrony and its impacts. EcoEvoRxiv Preprints. https://doi.org/10.32942/osf.io/jmy67. Published online ahead of print May 5, 2020.

    38.
    Robinson, R. A. et al. Travelling through a warming world: climate change and migratory species. Endanger. Species Res. 7, 87–99 (2009).
    ADS  Article  Google Scholar 

    39.
    Atkinson, A. et al. Krill (Euphausia superba) distribution contracts southward during rapid regional warming. Nat. Clim. Change 9, 142–147 (2019).
    ADS  Article  Google Scholar 

    40.
    Freer, J. et al. Limited genetic differentiation among Chinstrap penguin (Pygoscelis antarctica) colonies in the Scotia Arc and Western Antarctic Peninsula. Polar Biol. 38(9), 1493–1502 (2015).
    Article  Google Scholar 

    41.
    Clucas, G. V. et al. Comparative population genomics reveals key barriers to dispersal in Southern Ocean penguins. Mol. Ecol. 27, 4680–4697 (2018).
    CAS  Article  Google Scholar 

    42.
    Humphries, G. et al. Mapping Application for Penguin Populations and Projected Dynamics (MAPPPD): data and tools for dynamic management and decision support. Polar Rec. 53, 1–7 (2017).
    Article  Google Scholar 

    43.
    Lynch, H. J. & LaRue, M. A. First global census of the Adélie penguin. The Auk Ornithol. Adv. 131, 457–466 (2014).
    Google Scholar 

    44.
    Che-Castaldo, C. et al. Pan-Antarctic analysis aggregating spatial estimates of Adélie penguin abundance reveals robust dynamics despite stochastic noise. Nat. Commun. 8, 832 (2017).
    ADS  Article  Google Scholar  More

  • in

    Redox-informed models of global biogeochemical cycles

    Microorganisms drive biogeochemical cycling in the earth system1 (Fig. 1). Photoautotrophic microorganisms are responsible for about half of CO2 fixation and O2 production on earth, and heterotrophic microorganisms are responsible for much of the return reaction: the oxidation of organic matter back into CO2. The temporal and spatial separation of photoautotrophy and heterotrophy in the global environment drives the biological sequestration of carbon, the reduction of atmospheric CO2, and the maintenance of elevated atmospheric and oceanic O22,3,4,5. Chemoautotrophic microorganisms also fix CO2 and, together with anaerobic heterotrophic metabolisms, carry out diverse chemical transformations including the fluxes of nitrogen to and from biologically available states and the formation of the potent greenhouse gas nitrous oxide (N2O)6,7. Since these transformations respond to, and feedback on, changes in climate (Fig. 1), estimating microbial activity accurately at global scales is important for climate science.
    Fig. 1: Key microbially driven redox transformations that mediate the atmospheric fluxes of climatically relevant gases.

    Radiatively active gases are notated with red type. The processes in black type are represented in some way (though not necessarily with electron balancing) in both the marine and terrestrial biospheres in earth system models within the Coupled Model Intercomparison Project (land: NCAR Community Earth System Model103; ocean: GFDL COBALTv2104), which are used for projections of climate change in reports by the Intergovernmental Panel on Climate Change. Processes in green type are represented in only the terrestrial model. Current models do not yet include other relevant reactions, some of which are represented in gray type, such as anaerobic ammonia oxidation (anammox), the marine production and consumption of methane, the redox cycling of iron, manganese, and other metals, and the methane-relevant redox chemistry of phosphorus105. COBALTv2 does account for sulfate reduction in marine sediments, but sulfate is not represented. Image courtesy of NASA.

    Full size image

    However, understanding and projecting the impacts of microbial processes are limited in part due to oversimplified representation in earth system models. For example, in marine biogeochemical models, much attention is given to the complex impacts of phytoplankton—the photoautotrophic microorganisms responsible for primary production—and their small zooplankton predators8,9,10. The bacterial and archaeal activities responsible for other critical aspects of biogeochemical cycling in the land and ocean—remineralization, denitrification, nitrogen fixation, methanogenesis, etc.—are often crudely parameterized10. Such models have limited prognostic capability. For example, models typically prescribe the ecological niche of a given metabolism with imposed, empirically determined parameters that are site- or organism-specific. These parameterizations may or may not apply to other environments, including past and future ecosystems.
    These simplistic approaches have been largely necessary due to the difficulties of characterizing the taxonomy and metabolic capabilities of natural microbial communities. However, the rapid expansion of genetic sequencing capabilities has enabled a clearer view of microbial biogeography and activity in the environment. In consequence, computational biogeochemistry is opening up the black box of remineralization and other microbially mediated processes in marine and terrestrial environments11,12,13,14,15,16,17,18.
    As we expand models to include the full metabolic potential of microorganisms, how can we organize and reduce the complexity of the descriptions of metabolic diversity? Non-photosynthetic organisms oxidize chemical species for energy, and thus their respiration is biogeochemically significant19. Here, we explain how the key reduction-oxidation (redox) reactions that supply energy for metabolisms can provide an additional organizing principle for explicit descriptions of microbial populations in ecosystem models. This redox basis can be exploited to quantitatively resolve chemical transformations in terms of assimilatory and respiratory fluxes. While not yet incorporated into earth system models, this view has been advocated for such applications20, and has been embraced and employed in the field of environmental biotechnology, such as in the interpretation and modeling of wastewater bioreactors21. Just as models of ocean and atmospheric circulation are constrained by conservation of energy and potential vorticity, complementing mass balance with powerful redox and energetic constraints enables self-consistent descriptions of diverse microbial metabolisms.
    This approach aims to advance ecological modeling beyond species-specific descriptions to those that matter for biogeochemical function, in line with trait-based modeling approaches9. In analogy to the use of redox chemistry, trait-based functional type models of phytoplankton have used cell size as an organizing principle for understanding phytoplankton biogeography, biodiversity, and impact on biogeochemistry9,22,23. These types of theoretical constraints allow for the inclusion of more functional types without introducing as many degrees of freedom as would be necessary if each were empirically described. The guiding perspective is that organizing complex biological behavior by its underlying chemical and physical constraints gives more universally applicable descriptions of large-scale biogeochemistry.
    When incorporating a redox-balanced approach into ecosystem models, microbial function emerges from underlying chemistry as a consequence of interactions between populations modeled as metabolic functional types and their environment. Resulting theoretically grounded ecosystem models independently simulate microbial growth, respiration, and abundances in ways that we can compare with observations such as sequencing datasets. Thus, sequencing datasets are used as critical tests for the models, as external constraints rather than as input to the models, allowing for an iterative relationship between theory, observations, and models.
    In contrast with empirically informed models, this approach involves constructing a model of microbial activity theoretically, and then comparing the results with the observations in order to gain an understanding of the system. The goal is to understand why biology functions as it does, in addition to anticipating global impacts. From a first-principles biogeochemical perspective with respect to physical and chemical forcing, genes are an intermediate step between forcing and function, with the detailed complexity of biological reality following the underlying chemical and physical constraints (analogous to the form follows function principle of architect Louis Sullivan). This does not equate to thinking that biology (or genetic information) does not matter or can be replaced entirely in models by physics and chemistry. Rather, we want to fundamentally understand biological activity as an integrated part of an ecosystem, and physics and chemistry become tools for doing so.
    Here, we outline the basis for using redox chemistry as an organizing principle and its translation into quantitative descriptions of microbial activity that are simple enough for global earth system models. We then discuss the benefits of this approach in the context of their implications for improved understanding and projections of global change impacts. Finally, we discuss limitations and possible future developments.
    Predicting microbial activity
    From one perspective, microbial communities are characterized by interactions at the micro-scale: gene expression, enzymatic capabilities, metabolites, species-specific interdependencies, etc., as well as the physical and chemical environment surrounding small cells24,25,26,27,28. The information from sequencing in particular has allowed for a huge expansion of insight into the detailed in situ activity of uncultivated species. When investigating global-scale impacts, how do we decide which of these details may be bypassed for simplicity? Or, if this simplification is impossible, must we incrementally construct a microbial ecosystem model that incorporates all known micro-scale detail?
    Another way forward arises from a macro-scale perspective, which examines how ecosystem function relates to the chemical potential utilized by organisms for energy20,29,30. For example, it is well known that microbial communities in sediments and anoxic zones organize according to the redox tower – the ranking of half-reactions by electrochemical potential14,31,32. Furthermore, respiration by living organisms increases the entropy of the environment by dissipating concentrated sources of chemical energy in accordance with the Second Law of Thermodynamics30,33,34.
    This perspective suggests that chemical potential can be used to predict the activity of microbial communities and their biogeochemical impact. However, given the notorious complexity of microbial cells and systems, which is many steps away from governing chemical or physical equations, how can we be sure that this activity is indeed predictable? Frentz et al.35 demonstrated that external conditions cause the seemingly random fluctuations observed in microbial growth, rather than stochastic variation in gene expression. This provides direct evidence of deterministic behavior, and so the authors conclude that microbial systems can in principle be determined by macroscopic laws.
    How is this determinism manifested? If microbial communities can respond relatively quickly to changes to their local environment, they may predictably optimize the exploitation of locally available resources. In the ocean, dispersal in microbes is thought to be a highly efficient process such that microbial communities can in effect draw from an extensive seed bank36,37, as captured in the phrase “everything is everywhere, the environment selects”38. Furthermore, recent evidence also shows that gene acquisitions and deletions happen quickly enough to allow for horizontal gene transfer to dominate bacterial adaptation39,40,41,42,43, implying that evolution can occur within few generations and thus on timescales similar to ecological interactions. Perhaps consequentially, similar geochemical environments have been demonstrated to have high-microbial functional redundancy despite different taxonomic compositions17,44. This may be interpreted with the hypothesis that physics and chemistry selects for metabolic traits, and that these traits can be housed in different organisms with taxonomic composition shaped by micro-scale or biotic interactions17,44,45.
    The prediction of microbial activity from environmental chemical potential has a long history in microbiology20,21,46,47,48,49,50,51,52, and is conceptually similar to other redox-balanced approaches to understanding microbial activity in sediments, soils, subsurfaces, and aquatic systems13,31,53,54,55,56,57,58. Illustrating the power of these approaches, anticipating metabolism from chemical potential resulted in a prediction that anaerobic ammonia oxidation (anammox) should exist decades before it was observed59,60. Quantitatively understanding microbially mediated rates of conversion of substrates has practical implications for wastewater treatment, and thus the field of biotechnology has established methodologies for an approach in textbook form21. Flux balance analysis (FBA) models can be considered as much more highly detailed analogs of this approach that resolve the mass and electron balances among a multitude of chemical reactions within a cell61,62.
    Redox-balanced metabolic functional types
    We can resolve microbial activity in global ecosystem models using the underlying redox chemistry of diverse metabolisms as a constraint. One specific way forward is to model distinct metabolisms as populations of metabolic functional types. This systematically quantifies relative rates of substrate consumption, biomass synthesis, and excretions of transformed products associated with each metabolism. Coupled with estimates of substrate uptake, this replaces implicit parameterizations of processes such as organic matter consumption, oxygen depletion, and denitrification with electron-balanced respiratory fluxes of dynamic microbial populations. Box 1 provides a detailed description of this methodology for multi-dimensional models.
    A particular set of redox reactions may distinguish a functional type, such as the oxidation of organic matter using oxygen (aerobic heterotrophy), or the oxidation of ammonia or nitrite using oxygen (chemoautotrophic nitrification) as exemplified in Table 1. For each metabolism, an electron-balanced description consists of multiple half-reactions: biomass synthesis, oxidation of an electron donor, and reduction of an electron acceptor21,48. The ratio of anabolism and catabolism can then be represented by the fraction f of electrons fueling cell synthesis vs. respiration for energy, following ref. 21. This provides a yield y (moles biomass synthesized per mole substrate utilized) of each required substrate that reflects two inputs: electron fraction f and the coefficients of the half-reactions (Fig. 2). The interlinked yields reflect the energy supplied by the redox reaction, the energy required for synthesis and other cellular demands, and the inefficiencies of energy conversion. Either f or y for any one of the substrates may be estimated theoretically with Gibbs free energies of reaction21 or with a combination of theoretical and empirical strategies63.
    Table 1 Simplified equations describing two exemplary metabolic functional types.
    Full size table

    Fig. 2: Schematic of a single cell represented as a metabolic functional type carrying out the aerobic oxidation of ammonia.

    The redox balance informs the elemental ratios of substrates utilized, biomass synthesized, and waste products excreted (Table 1).

    Full size image

    The result is a stoichiometric budget of the metabolism of the whole organism (Table 1). These descriptions quantify the elemental ratios of utilized substrates, biomass, and the excretion of waste products. For example, the descriptions account for the CO2 produced by heterotrophic metabolisms as well as the CO2 fixed by chemoautotrophic metabolisms (Table 1, Fig. 2, and Supplementary Fig. 2), linking microbial activity directly to global carbon cycling.
    To estimate the growth rate of each functional type, the yields from the metabolic budgets are combined with the uptake rates of the required substrates (Box 1). Limiting uptake rates may rely on empirically derived uptake kinetic parameters, or they can be estimated theoretically from diffusive supply, cell size, membrane physiology, and other physical constraints64,65,66. If theoretical models of uptake are used, the physical constraints on substrate acquisition and the redox chemical constraints on energy acquisition can provide an entirely theoretical estimate of the growth of each metabolic functional type.
    One strategy is to represent the populations carrying out each of these discrete metabolisms as one functional type population, which aggregates the diverse community of many species that are fueled by the same (or a similar) redox reaction (Fig. 3). Such aggregation has been deemed a useful strategy for representing the biogeochemical impacts of microbial communities for certain research questions67,68. However, for other questions this wipes out critical diversity among the aggregated populations. For example, diverse aerobic heterotrophic populations consume organic matter over a wide range of rates, and these rates dictate the amount of biologically sequestered carbon in the ocean. Redox chemistry and physical limitations alone may not inform the heterogeneity among similar metabolisms. One additional constraint is the limited capacity of the cell and thus its allocation of proteome towards different functions69. While the electrons supplied to the cell must be conserved following the redox balance, the electrons may be partitioned differently into machinery for substrate uptake vs. biomass synthesis, for instance, for different phenotypes. This partitioning can be quantitatively related to ecological fitness and biogeochemical impact via uptake kinetics, effective yields, and other traits9 (Supplementary Note 1 and Supplementary Fig. 1).
    Fig. 3: Solutions from a global simulation resolving multiple metabolic functional types.

    Net primary productivity (NPP), the biomasses of the metabolic functional types, and the sinking particulate organic carbon (POC) flux are resolved along a transect of a global microbial ecosystem model coupled with an estimate of the ocean circulation (Darwin-MITgcm18).

    Full size image

    Box 1 Incorporating metabolic functional types into ecosystem models

    A metabolic functional type can be represented as a population with a growth rate that is limited or co-limited by multiple required substrates. If Liebig’s Law of the Minimum is employed, the limiting growth rate μ is described as

    $$mu = min (V_{mathrm{i}}y_{mathrm{i}})$$
    (1)

    where Vi is the specific uptake rate of substrate i, and yield yi is the biomass yield with respect to that substrate. Yields for the different substrates and elements are interlinked in the metabolic budget derived from the underlying redox chemistry. Yields reflect Gibbs free energies of reaction among other factors. In the simplest model, non-limiting substrates are consumed in proportion to the limiting resource according to the metabolic budget, although in reality they may accumulate in the form of storage molecules.
    Each metabolic functional type population can be incorporated into a multi-dimensional environmental model (e.g., an ocean simulation) with physical transport as

    $$frac{{dB}}{{dt}} = mu B – Lleft( B right)B – underbrace {nabla cdot left( {bf{u}}{B} right)}_{{mathrm{advection}}} + underbrace {nabla cdot left( {{mathbf{kappa }}nabla B} right)}_{{mathrm{diffusion}}}$$
    (2)

    for biomass concentration B, loss rate L, velocity u, and diffusion coefficient κ. The loss rate function varies with biomass and represents a combination of processes, including predation, viral lysis, maintenance, and senescence. These processes remain largely unconstrained, although efforts have been made to relate losses to ecological dynamics107,108.
    The yield partitions the amount of substrate taken up by the population into that used for growth, Viyi, versus that exiting the cell in modified form as a waste product, Vi(1 − yi) (Fig. 2 and Table 1). Equation 1 suggests a correlation between μ and y, but yields may be further modified by other factors. For example, accounting for maintenance energy decreases the ratio of growth to respiration, contributing to a decoupling between growth rate and yield particularly at low growth rates109. Furthermore, a trade-off between uptake rate and yield at the cellular level reflects the allocation of enzyme towards machinery for substrate uptake vs. biomass synthesis, among other factors. Considering a proteome constraint can incorporate this trade-off (Supplementary Note 1).

    Benefits and implications for anticipating global change
    Redox chemistry aids in reducing the number of degrees of freedom in descriptions of diverse microbial metabolisms. We next discuss the benefits of this electron-balanced approach, each contextualized by specific projected impacts of global change due to microbial activity and broad challenges in the fields of microbial ecology and biogeochemistry.
    Flexible and broadly applicable metabolic thresholds: A key question for microbial biogeochemical studies, for which biogeochemical models are primed to answer, is how the biogeographies of diverse, active metabolisms vary with changes in the physical and chemical environment. What threshold determines the viability of a given metabolism?
    Redox-balanced metabolic budgets obviate the need to impose critical concentrations or other thresholds that determine the presence of any given metabolism. Rather than being imposed following empirical relationships, metabolic biogeography emerges dynamically from ecological interactions and reflects environmental chemical potential. This flexibility aids in understanding metabolic thresholds more fundamentally, and it expands model applicability to diverse and unobserved environments. This is of particular importance for understanding global change, since past and future worlds may include very different ecosystems that do not reflect current empirical trends.
    For example, the oceans are currently losing oxygen due to global warming5,70. If anoxic zones continue to expand, this will increase the habitat of anaerobic microorganisms, whose respiration results in emissions of N2 and N2O to the atmosphere7. Many biogeochemical models prescribe O2 concentrations that inhibit anaerobic activity in accordance with observations of specific organisms or communities in experimental conditions. This assumes that the same O2 concentrations limit metabolism similarly in all environments, and often trades mechanistic understanding of oxygen limitation for empirical correlations that may reflect a variety of natural and introduced biases, such as micro-scale heterogeneity, physical mixing in the ocean, and experimental bottle effects.
    In contrast, a metabolic functional type model does not require imposed oxygen threshold concentrations (Supplementary Fig. 3). When oxygen supply is abundant, anaerobic types are competitively excluded because growth using alternative electron acceptors is lower than with oxygen. When oxygen supply is low, aerobic populations may persist and continue to deplete any available oxygen even as their growth is limited by oxygen, allowing for a steady state stable coexistence of aerobic and anaerobic metabolisms, which is consistent with a variety of observations71.
    Descriptions of microbial growth that reflect underlying chemical potential can enable predictions of many other metabolic transitions, such as nitrogen fixation, nitrification, and the transition to sulfur oxidation and reduction13,18,72,73,74. As another example, this approach predicts the restriction of nitrification from the sunlit surface ocean as a consequence of competitive exclusion by phytoplankton in many environments (Fig. 4 and Supplementary Fig. 2), as well as active nitrification in some surface locations where phytoplankton are limited by another factor not affecting the chemoautotrophs, such as at high latitudes where phytoplankton are limited by light18. The emergent exclusion from most of the surface ocean anticipates that many clades of nitrifying microorganisms have adapted to long-term exclusion from the surface and consequentially lost (or did not develop) photoprotective cellular machinery.
    Fig. 4: Model simulation and observations of the marine nitrification system.

    Biogeochemistry is driven by microbial metabolic functional types in a vertical water column model18. Lines are model solutions, and marked points are observations from two stations in the Pacific Ocean75,106 (see Supplementary Fig. 2 for more detail) (a). Chlorophyll a concentrations and abundances of ammonia-oxidizing organisms (AOO) and nitrite-oxidizing organisms (NOO). Observed abundances are of the 16S rRNA abundances of archaeal Marine Group I and Nitrospina-like bacteria75,106. Model abundances are converted from biomass with 0.1 fmol N cell-1 for AOO, 0.2 fmol N cell-1 for NOO76, and one gene copy per cell. (b). Light (solar irradiance) and bulk nitrification rates.

    Full size image

    Replacing implicit descriptions of organic matter remineralization: The fate of organic matter dictates the amount of carbon sequestered in the marine and terrestrial biospheres. Microbial consumption mediates the carbon stored in soils, the carbon stored in the ocean as dissolved organic matter (DOM), and the sinking flux of organic carbon that constitutes the marine biological carbon pump4, without which atmospheric CO2 would be 100–200 ppm higher than current levels. We want to understand how these carbon reservoirs respond to changes in climate, such as increased temperatures and changes in precipitation patterns. However, in biogeochemical models, simple rate constants often dictate the remineralization of elements from organic back into inorganic constituents.
    Replacing simplistic parameterizations with dynamic metabolic functional types means that electron-balanced descriptions of growth and respiration instead drive the fate of organic matter in earth system models (Fig. 3). In addition to a more sophisticated and responsive description of carbon sequestration, non-living organic matter is fully integrated into ecosystem frameworks, enabling theoretical studies of phytoplankton-bacteria interactions to complement observational and experimental approaches.
    Much work remains in the development of these descriptions. As we discuss below, accurate estimates of organic matter turnover rates require more accurate descriptions of the complex processes governing microbial uptake rates of organic matter. However, the redox-informed yields are still useful for quantifying the relative amount of CO2 excreted and the absolute amount of biomass sustained on a given substrate, independent of uptake kinetics (Supplementary Note 2).
    Relationships between abundances, rates, nutrient concentrations, and elemental ratios: An overarching puzzle challenging microbial ecology is to understand how chemical transformations in the environment are set by the ecological interactions at the organism level, among individual microscopic cells. It is clear that abundances of populations are not simply and directly correlated with biogeochemical impact (i.e., higher abundance does not necessarily imply an associated higher rate of chemical transformation). Untangling the relationship between abundances and biogeochemical function is also necessary for interpretation of genetic evidence that provide insight into this complex ecosystem structure.
    Redox-balanced metabolic functional type modeling links rates of biomass synthesis associated with a particular metabolism to its rate of respiration as well as the standing stock of limiting nutrients. As functional type modeling is coupled with estimates of population loss rates due to grazing, viral lysis, or other mortality, simulations also resolve the standing stocks of functional biomass. This quantifies the relationship between biomass concentrations and volumetric rates of chemical transformations, emphasizing how relatively low biomass may be associated with relatively high bulk rates71.
    For example, the approach has revealed a clear example of the signature of chemical potential in the ecology of marine nitrification18 (Fig. 4). In this model, the two steps of nitrification are represented by two functional type populations. This predicts about a three-fold difference in the abundances of the organisms responsible for each of the two steps of nitrification, despite the fact that the two populations carry out the same rate of subsurface N-cycling at steady state6,18. A three-fold or greater difference in abundance and associated ammonium (NH4+) and nitrite (NO2−) concentrations is consistent with observed differences18,75, and it reflects that the oxidation of one mole of NH4+ generates three times more electrons than the oxidation of one mole of NO2−, with differences in cell size further contributing to differences in abundances (Fig. 4 and Supplementary Fig. 2). Recent observations confirm the redox-based difference in NH4+ and NO2− biomass yield76,77, although measured rates from a nonsteady environment suggest that NO2−-oxidizing bacteria can partition electrons more efficiently than NH4+-oxidizing archaea76 (i.e., higher fraction f despite lower yield y; see Supplementary Note 3).
    As redox-based descriptions resolve the stoichiometry of whole organism metabolism, they also link together elemental cycles. Explicit description of relative elemental flow through the ecosystem, and specifically their variation from average values, is critical for understanding climate-biogeochemical feedbacks78,79,80. For example, the nitrification model also estimates the CO2 fixation rates associated with nitrification rates (Supplementary Fig. 2), enabling global-scale, electron-balanced projections of the amount of carbon converted to organic form by chemoautotrophic nitrifying microorganisms.
    Connections with sequencing datasets: How do we relate metabolic functional type models to sequencing datasets measuring genetic, transcriptomic, and proteomic diversity? Connecting biogeochemical models with sequencing data is critical because this data provides an enormous amount of information about ecosystem structure and function. Genes (or transcripts) themselves are not necessarily the most concise or useful currency given functional redundancies, unattributed function, and variation in gene dosage from horizontal gene transfer as well as growth rate40. Recent gene-centric models aim to resolve the abundances of key genes as proxies for a predetermined set of metabolic pathways13,14,17. However, the parameters used to describe metabolic pathways in these models are estimated similarly to the redox-balanced yields and efficiencies described here.
    The innovation of gene-centric models is the sophisticated conversion of estimates of biogeochemical activity and biomass to genes. For example, the model of Coles et al.17 resolves biomass and nutrient concentrations prognostically, and then uses a three-part formula—representing constitutive, regulated, and steady state transcription—to diagnostically calculate transcription rates from modeled biomass and growth rates17. Thus, the two types of modeling are complimentary, with redox chemistry providing estimates of metabolic activity from fundamental principles, and the careful calibrations between activity and sequencing providing a comparative metric.
    The examples here externalize the conversion between modeled activity and sequencing information as a transparent process. In Fig. 4, the predicted functional biomass of ammonia-oxidizing population is related to archaeal Marine Group I (MGI) and Nitrospina-like 16S rRNA genes with two conversion factors: the cell elemental quota (fmol N cell−1) and the number of cellular gene copies. Conversion error arises since cell mass and size vary with growth rate81,82. Maintaining transparency of the conversion from predicted microbial activity to genes and transcripts allows interdisciplinary audiences to understand and critique the models.
    Limitations and possible extensions
    Using chemical potential as a theoretically grounding organizing principle for the resolution of diverse metabolisms can greatly improve microbial descriptions in global biogeochemical models. However, the approach does have its limitations, which generally increase in significance with increased temporal or spatial resolution.
    Modeling metabolic diversity with functional type populations requires choosing how metabolisms are distributed among the populations. This has consequences when interpreting time-varying states: model solutions become dependent on the partitioning of metabolism among the functional types as the timescales of physical change approach the timescales of microbial growth (see Supplementary Note 4, Supplementary Fig. 3, and Supplementary Fig. 4 for a detailed example). Other species-specific time-varying phenomena such as the lag response of organisms to substrate availability also become relevant83. On one hand, this is beneficial for resolution of microbial processes in fine-grained ocean circulation models where flow can vary on the order of days. However, incorporating another constraint, such as proteome allocation69, is necessary to inform these choices. For example, considering enzymatic allocation in combination with energetics allowed for the prediction of both the division of nitrification into a two-step process in mixed environments and the combined, complete pathway in one organism (comammox) in biofilms, which preceded observations of the latter84,85,86.
    Uncertainty in distributions of metabolism lies not only in the length of a metabolic pathway, but also in the degree of metabolic versatility (metabolic mixotrophy). Such versatility characterizes key players in large-scale biogeochemistry, such as nitrite-oxidizing bacteria and photoheterotrophs87,88,89. Mixotrophic lifestyles can increase the fitness of populations in their environments, impacting overall ecosystem function90. In one sense, the approach here provides a prediction of where we might expect such mixotrophy by resolving stable coexistences of diverse metabolisms. In Fig. 3, for example, syntrophic coexistence occurs at depth among heterotrophs, ammonia oxidizers, and nitrite oxidizers, and future work could investigate what determines which combinations of these coexistences remain as passive interactions, which develop into mutualistic dependencies as active interactions91, and which evolve into mixotrophic phenotypes or endosymbionts. Additionally, by considering the potential to carry out a metabolism as a trait, we can use the current framework along with an additional constraint to investigate implications of metabolic mixotrophy. For example, Coles et al.17 impose a trade-off between the degree of metabolic diversity of a single functional type and growth rate, enabling the exploration the consequences of distribution of metabolism on the biogeochemical state.
    Also, the metabolic functional type approach resolves only active functional biomass, while evidence suggests that less than 10% to more than 75% of the microbial community may be inactive92. Some seemingly inactive populations may slowly metabolize over long timescales, requiring longer model integration times and careful attention to their loss rates for resolution, while some populations are periodically active as revealed by high-resolution observations in time37.
    The proposed modeling approach relies on estimates of the limiting uptake rates of required substrates. In lieu of suitable theoretical descriptions, the use of empirically derived uptake kinetic parameters still employs the benefits of the redox-informed yields (Supplementary Note 2). However, underlying physical constraints to substrate acquisition can in principle be exploited to develop more universally applicable descriptions for a variety of substrates and contexts. Uptake kinetics are complex, but for many limiting resources, encounter effectively controls the uptake, and the physics of encounter has been relatively well described. For example, uptake rates estimated from diffusive supply of substrate, cellular geometry, and membrane physiology64,65,66 have been empirically supported93. For organic matter, future work is needed to develop suitable descriptions of consumption rates, whether empirical or theoretical. For example, descriptions require attention to the hydrolysis of organic compounds by extracellular enzymes and the ecology of sinking marine particles—the diffusion of monomer away from the particle, within-particle transport, and dynamic ecological interactions on particle surfaces, among other processes11,74,94,95.
    In Fig. 3, the electron-balanced description consists of an average stoichiometry and electron fraction for one sinking pool of organic matter in the ocean, which, as mentioned above, is not sufficient to accurately resolve the carbon storage that is shaped by a distribution of turnover rates. As one of the many factors impacting the rates, an energetics-based perspective can serve as a tool for further deciphering organic matter complexity. For example, organic matter may be partially organized by the nominal oxidation state of its carbon atoms, which relates to a measure of free energy and accessibility96,97. This could be used to improve the phenomenological description of organic matter in models as labile vs. non-labile, for example, with a more mechanistic underpinning.
    Descriptions of phytoplankton are currently much more sophisticated than of bacteria and archaea in models, reflecting a longer history of comprehensive sets of observations. However, further work could develop simple descriptions of photoautotrophic metabolisms from underlying energetics by connecting the supply of photons to available energy for biosynthesis within the cell. Many biogeochemical models account for an inefficiency of phytoplankton metabolism with a parameter that dictates their excretions of dissolved organic matter98. Incorporating this excretion into an energetic framework would enhance studies of phytoplankton ecology, such as studies of photoautotrophic-heterotrophic interactions in the ocean surface or photoautotrophic-chemoautotrophic interactions at the base of the euphotic zone where some phytoplankton excrete nitrite due to incomplete reduction of nitrate99.
    As a more radical extension, can we progress past population modeling and model microbial consortia as one aggregate community biomass34,100? This may improve resolution of time-varying metabolic versatility. However, if both steps of nitrification were a part of such a consortium, would the characteristic accumulation of nitrite be predicted (Supplementary Fig. 2)? We leave these questions for future research and conclude that the best choice for the degree of resolution of metabolism will depend on the specific research question and the available observations.
    We have described a useful approach for understanding and anticipating microbial control of biogeochemical cycling that is suitable for global applications. The approach aims to represent microbial growth and respiration explicitly and consistently from knowledge of chemical gradients in the environment, towards a goal of building an independently constructed theoretical ecosystem model that can then be compared to observations. Describing microbial communities with underlying energetic constraints connects metabolisms dynamically with global geochemical distributions, such as those of carbon dioxide, oxygen, and biologically available nitrogen. This deepens our understanding of microbial ecosystems and enables the incorporation of the feedbacks of microbial activity to changes in global biogeochemistry and the climate system. More

  • in

    Experimental tests of bivalve shell shape reveal potential tradeoffs between mechanical and behavioral defenses

    Experimental limitations and implications
    Many factors are known to contribute to mollusk shell strength including thickness, microstructure, previous shell damage, and ornamentation19,21,29,30,31. Thus, it is challenging to separate the influence of one feature from another when testing the strength of real shells. Fortunately, there are alternative methods by which to model shells22,32,33,34. 3D printing is an effective tool to create model shells that exhibit brittle behavior in compression, serving as an accurate first-order approximation for shell breakage (Fig. 2). 3D printed models are not created to replicate shell microstructure; rather they serve to normalize confounding factors encountered with real shells, like size, variations in thickness, and taphonomy/degradation of microstructure, to isolate variables of interest while maintaining the predominantly brittle behavior seen in real shells. As a result, the failure of different 3D printed shapes under bulk mechanical compression can be used for relative comparisons between morphologies without needing to reproduce the exact magnitude of load to failure of real shells22,33 (see “Methods”, Fig. 2). Natural shells ultimately break after cracking through microstructural layers. Analogously, in this study, 3D printed shells cracked predominantly through layers of printed materials, not along printing boundaries, demonstrating an appropriate proxy for shell failure (Figs. 1, 2).
    Resistance to compression by a predator can be tested directly using a variety of loading tests, the effectiveness of which has been demonstrated repeatedly by investigators studying topics ranging from predation to climate change22,33,34. Compression experiments of bivalve shells, specifically, have been used extensively for this purpose35,36,37,38,39,40,41,42,43,44,45,46. In this study, the mechanical strength of model bivalve shell shapes was analyzed to understand potential defensive value against durophagous (crushing) predators using hypothetical flat teeth and jaws to crush prey (see Crofts and Summers (2014)22 for examples of flat crushing morphologies) (Figs. 1, 2). These experiments are most analogous to small shell-crushers with flat dentitions, like fishes (e.g., guitar fishes, stingrays, etc.), rather than large predators like the modern walrus or extinct marine reptiles (e.g., placodonts, mosasaurs) which are so much larger than their prey that the differences in shell strength resulting from shape are likely insignificant. Bivalves are also preyed upon by many other predators including asteroids, gastropods, birds, and mammals47. The ability to escape is likely dependent upon predator capabilities, where not all escape mechanisms are equally effective against different predators. The results of these experiments are not intended to be general proxies of predation resistance—they are only applicable as a proxy for predators that use flat crushing dentitions. Additionally, it is important to note that compression of a shell by vertebrate predators is a different mechanical process from compression by an arthropod, as claws localize forces differently than teeth and jaws48. Therefore, the experiments used in this study are not meant to model predation by invertebrate durophages. This experimental setup represents an idealized case of shell compression assuming consistent shell thickness and a common predator using a quasistatic loading regime to isolate the influence of shell shape. Therefore, this study makes no conclusions as to the effects of shell shape on strength under conditions of point loading by claws or impact. Furthermore, while bivalve shape is undoubtedly influenced by many factors including fabrication (shell growth and construction)49, phylogenetics, location of soft tissue, etc., these experiments were designed to study an idealized case in which defense against vertebrate shell crushing predators is the most important functional constraint on shell shape.
    Advantageously, the use of mathematically generated theoretical bivalve shells, rather than real shells, enables testing a range of shapes—some of which can be or have been found in nature—and others that have yet to exist. For example, these methods enable physical testing of morphologies which are only found in the fossil record. Fossils cannot be used for mechanical experimentation due to changes in the integrity of the shell structure resulting from the fossilization process. Furthermore, testing shapes which do not exist due to biological constraints like fabrication, can provide valuable insights when combined with the study of shapes that have evolved naturally. For example, testing theoretical shapes can reveal morphologies that perform better than natural morphologies in specific functional settings. Thus, testing shapes which perform well, but do not exist, is one way to identify the potential influence of evolutionary constraints such as fabrication (shell growth patterns), phylogenetics, or other necessary biological functions. In this study, the extremely perpendicular-elongate shells represent theoretical shapes that are not seen in nature (likely due to space limitations for soft tissue). Developing theoretical physical models also allows for the elimination of many confounding variables that cannot be avoided when testing real shells. The effects of variations in shell thickness and ornamentation, which also play a role in shell strength, can thus be separated from the strength imparted by gross shell shape when theoretical models are generated. However, because overall shell strength is derived from a combination of these many factors (microstructure, thickness, etc.31.) the conclusions from these experiments represent a first-order approximation of the effects of shape on shell strength alone. Because the models for this study were generated mathematically, and are therefore not exact replicas of specific taxa, the discussion below aims to address potential tradeoffs based on general shell shapes that could be further studied for specific bivalve taxa in future experiments.
    Shell shape and strength
    Three parameters of shell shape were modified for this study. (1) Generating curve shape is both a modeling parameter and biological feature of mollusk shells27. In mathematical models, generating curve shape describes the shape that is rotated around an axis to create a surface that represents the overall shell shape. Biologically, the generating curve refers to the portion of the shell upon which new material is secreted by the mantle. In bivalves, this is the shape of the commissure. Thus, by changing the shape of the generating curve in a model bivalve, i.e., changing the shape of the ellipse that generates the surface of the shell, models can be perpendicular-elongated (a  > b) or parallel-elongated (a  More

  • in

    Ecological corridors for the amphibians and reptiles in the Natura 2000 sites of Romania

    1.
    Turner, M. G. Landscape ecology: The effects of pattern on process. Annu. Rev. Ecol. Syst. 20, 171–197 (1989).
    Article  Google Scholar 
    2.
    Noss, R. F. Wildlife corridors. in Ecology of Greenways (eds. Smith, D. & Hellmund, P.) 43–98 (University of Minesota Press, Minesota, 1993).

    3.
    Brooks, T. M. et al. Habitat loss and extinction in the hotspots of biodiversity. Conserv. Biol. 16, 909–923 (2002).
    Article  Google Scholar 

    4.
    Hanski, I. The Shrinking World: Ecological Consequences of Habitat Loss, Vol. 14 (International Ecology Institute, Philadelphia, 2005).
    Google Scholar 

    5.
    IUCN. The World Conservation Strategy. (IUCN, UNEP, 1980).

    6.
    IUCN. The IUCN Red List of Threatened Species. Version 2014.3. (2014).

    7.
    Dirnböck, T., Dullinger, S. & Grabherr, G. A regional impact assessment of climate and land-use change on alpine vegetation. J. Biogeogr. 30, 401–417 (2003).
    Article  Google Scholar 

    8.
    Gonçalves, J., Honrado, J. P., Vicente, J. R. & Civantos, E. A model-based framework for assessing the vulnerability of low dispersal vertebrates to landscape fragmentation under environmental change. Ecol. Complex. 28, 174–186 (2016).
    Article  Google Scholar 

    9.
    Saunders, D. A., Hobbs, R. J. & Margules, C. R. Biological consequences of ecosystem fragmentation: A review. Conserv. Biol. 5, 18–32 (1991).
    Article  Google Scholar 

    10.
    Fahrig, L. & Merriam, G. Conservation of fragmented populations. Conserv. Biol. 8, 50–59 (1994).
    Article  Google Scholar 

    11.
    Wiens, J. A. Habitat fragmentation: Island v landscape perspectives on bird conservation. Ibis 137, S97–S104 (1994).
    Article  Google Scholar 

    12.
    Diamond, J. M. ‘Normal’ extinctions of isolated populations. In extinctions (ed. Nitecki, M. H.) 191–246 (University of Chicago Press, Chicago, 1984).
    Google Scholar 

    13.
    Laurance, W. F. Comparative responses of five arboreal marsupials to tropical forest fragmentation. J. Mammal. 71, 641–653 (1990).
    Article  Google Scholar 

    14.
    Bennett, A. F. Linkages in the Landscape: The Role of Corridors and Connectivity in Wildlife Conservation. (IUCN, 2003).

    15.
    Opdam, P. Metapopulation theory and habitat fragmentation: a review of holarctic breeding bird studies. Landsc. Ecol. 5, 93–106 (1991).
    Article  Google Scholar 

    16.
    Thomas, C. D. & Jones, T. M. Partial recovery of a skipper butterfly (Hesperia comma) from population refuges: Lessons for conservation in a fragmented landscape. J. Anim. Ecol. 62, 472–481 (1993).
    Article  Google Scholar 

    17.
    Haddad, N. M. et al. Corridor use by diverse taxa. Ecology 84, 609–615 (2003).
    Article  Google Scholar 

    18.
    Grab, H. et al. Habitat enhancements rescue bee body size from the negative effects of landscape simplification. J. Appl. Ecol. 56, 2144–2154 (2019).
    Article  Google Scholar 

    19.
    Smeraldo, S. et al. Modelling risks posed by wind turbines and power lines to soaring birds: The black stork (Ciconia nigra) in Italy as a case study. Biodivers. Conserv. 29, 1959–1976 (2020).
    Article  Google Scholar 

    20.
    Noss, R. F. A regional landscape approach to maintain diversity. Bioscience 33, 700–706 (1983).
    Article  Google Scholar 

    21.
    Noss, R. F. & Harris, L. D. Nodes, networks and MUMS: Preserving diversity at all scales. Environ. Manag. 10, 299–309 (1986).
    ADS  Article  Google Scholar 

    22.
    Grumbine, R. E. What is ecosystem management?. Conserv. Biol. 8, 27–38 (1994).
    Article  Google Scholar 

    23.
    Forman, R. T. T. Land Mosaics. The Ecology of Landscapes and Regions ( Cambridge University Press, Cambridge, 1995).
    Google Scholar 

    24.
    Jongman, R. H. G. Nature conservation planning in Europe: Developing ecological networks. Landsc. Urban Plan. 32, 169–183 (1995).
    Article  Google Scholar 

    25.
    Kubeš, J. Biocentres and corridors in a cultural landscape. A critical assessment of the ‘territorial system of ecological stability’. Landsc. Urban Plan. 35, 231–240 (1996).
    Article  Google Scholar 

    26.
    Diamond, J. M. The island dilemma: lessons of modern biogeographic studies for the design of natural reserves. Biol. Cons. 7, 129–146 (1975).
    Article  Google Scholar 

    27.
    Wilson, E. O. & Willis, E. O. Applied biogeography. In Ecology and Evolution of Communities (eds Cody, M. L. & Diamond, J. M.) 522–534 (Belknap Press, New York, 1975).
    Google Scholar 

    28.
    Soulé, M. E. Land use planning and wildlife maintainance: Guidelines for conserving wildlife in an urban landscape. J. Am. Plan. Assoc. 3, 313–323 (1991).
    Article  Google Scholar 

    29.
    Opdam, P., Van Apeldoorn, R., Schotman, A. & Kalkhoven, J. Population responses to landscape fragmentation. In Landscape Ecology of A Stressed Environment (eds Vos, C. C. & Opdam, P.) 147–171 (Chapman and Hall, London, 1993).
    Google Scholar 

    30.
    Beier, P. & Noss, R. F. Do habitat corridors provide connectivity?. Conserv. Biol. 12, 1241–1252 (1998).
    Article  Google Scholar 

    31.
    Brown, J. H. & Kodric-Brown, A. Turnover rates in insular biogeography: Effect of immigration on extinction. Ecology 58, 445–449 (1977).
    Article  Google Scholar 

    32.
    Barrett, G. W. & Bohlen, P. J. Landscape Ecology Landscape Linkages and Biodiversity (Island Press, New York, 1991).
    Google Scholar 

    33.
    Forman, R. T. T. & Godron, M. Landscape Ecology (Wiley, New York, 1986).
    Google Scholar 

    34.
    Gilbert-Norton, L., Wilson, R., Stevens, J. R. & Beard, K. H. A meta-analytic review of corridor effectiveness. Conserv. Biol. 24, 660–668 (2010).
    Article  Google Scholar 

    35.
    Mech, S. G. & Hallett, J. G. Evaluating the effectiveness of corridors: A genetic approach. Conserv. Biol. 15, 467–474 (2001).
    Article  Google Scholar 

    36.
    Harris, L. D. & Scheck, J. From implications to applications: the dispersal corridor principle applied to the conservation of biological diversity. in Nature Conservation 2: The Role of Corridors (eds. Saunders, D. A. & Hobbs, R. J.) 189–220 (Surrey Beatty & Sons, 1991).

    37.
    Hobbs, R. J. & Hopkins, A. J. M. The role of conservation corridors in a changing climate. In The Role of Corridors (eds Saunders, D. A. & Hobbs, R. J.) 281–290 (Surrey Beaty & Sons, New York, 1991).
    Google Scholar 

    38.
    McLaughlin, J. F., Hellmann, J. J., Boggs, C. L. & Ehrlich, P. R. Climate change hastens population extinctions. Proc. Natl. Acad. Sci. 99, 6070–6074 (2002).
    ADS  CAS  Article  Google Scholar 

    39.
    Bennett, G. & Mulongoy, K. J. Review of Experience with Ecological Networks, Corridors and Buffer Zones (Secretariat of the Convention on Biological Diversity, 2006).

    40.
    MacArthur, R. H. & Wilson, E. O. An equilibrium theory of insular zoogeography. Evolution 17, 373–387 (1963).
    Article  Google Scholar 

    41.
    MacArthur, R. H. & Wilson, E. O. The Theory of Island Biogeography (Princeton University Press, Princeton, 1967).
    Google Scholar 

    42.
    Hanski, I. & Gilpin, M. Metapopulation dynamics: Brief history and conceptual domain. Biol. J. Lin. Soc. 42, 3–16 (1991).
    Article  Google Scholar 

    43.
    Bosso, L., Mucedda, M., Fichera, G., Kiefer, A. & Russo, D. A gap analysis for threatened bat populations on Sardinia. Hystrix Ital. J. Mammal. 27, 11788. https://doi.org/10.4404/hystrix-27.2-11788 (2016).
    Article  Google Scholar 

    44.
    Deus, E. et al. Current and future conflicts between eucalypt plantations and high biodiversity areas in the Iberian Peninsula. J. Nat. Conserv. 45, 107–117 (2018).
    Article  Google Scholar 

    45.
    Johovic, I., Gama, M., Banha, F., Tricarico, E. & Anastácio, P. M. A potential threat to amphibians in the European Natura 2000 network: Forecasting the distribution of the American bullfrog Lithobates catesbeianus. Biol. Conserv. 245, 108–551. https://doi.org/10.1016/j.biocon.2020.108551 (2020).
    Article  Google Scholar 

    46.
    van der Sluis, T. et al. How much Biodiversity is in Natura 2000? The “Umbrella Effect” of the European Natura 2000 protected area network. 147 (Alterra, Wageningen, 2016).

    47.
    Natura 2000 https://ec.europa.eu/environment/nature/natura2000/index_en.htm (2019).

    48.
    European Comission. Green Infrastructure (GI)—Enhancing Europe’s Natural Capital. 11 (Brussels, 2013).

    49.
    European Comission. Technical information on Green Infrastructure (GI). 24 (Brussels, 2013).

    50.
    Araújo, M. B., Thuiller, W. & Pearson, R. G. Climate warming and the decline of amphibians and reptiles in Europe. J. Biogeogr. 33, 1712–1728 (2006).
    Article  Google Scholar 

    51.
    Bennett, A. F. & Saunders, D. A. Habitat fragmentation and landscape change. In Conservation Biology for All (eds Sodhi, N. S. & Ehrlich, P. R.) 88–106 (Oxford University Press, Oxford, 2010).
    Google Scholar 

    52.
    Cushman, S. A. Effects of habitat loss and fragmentation on amphibians: A review and prospectus. Biol. Cons. 128, 231–240 (2006).
    Article  Google Scholar 

    53.
    Keinath, D. A. et al. A global analysis of traits predicting species sensitivity to habitat fragmentation. Glob. Ecol. Biogeogr. 26, 115–127 (2017).
    Article  Google Scholar 

    54.
    Blaustein, A. R. et al. Amphibian breeding and climate change. Conserv. Biol. 15, 1804–1809 (2001).
    Article  Google Scholar 

    55.
    Gibbons, W. J. et al. The Global Decline of Reptiles Déjà Vu Amphibians. Bioscience 50, 653–666 (2000).
    Article  Google Scholar 

    56.
    Rivera-Ortiz, F. A., Aguilar, R., Arizmendi, M. D. C., Quesada-Avendaño, M. & Oyama, K. Habitat fragmentation and genetic variability of tetrapod populations. Anim. Conserv. 18, 249–258 (2015).
    Article  Google Scholar 

    57.
    Andrews, K. M., Gibbons, J. W. & Jochimsen, D. M. Ecological effects of roads on amphibians and reptiles: A literature review. In Urban Herpetology (eds Mitchell, J. C. et al.) 121–143 (Society for the Study of Amphibians & Reptiles, London, 2008).
    Google Scholar 

    58.
    Hansen, N. A., Sato, C. F., Michael, D. R., Lindenmayer, D. B. & Driscoll, D. A. Predation risk for reptiles is highest at remnant edges in agricultural landscapes. J. Appl. Ecol. 56, 31–43 (2019).
    Article  Google Scholar 

    59.
    McCallum, M. L. Tropical Herpetology: A drop in the bucket. Trends Ecol. Evol. 20, 289–290 (2005).
    Article  Google Scholar 

    60.
    Bonnet, X., Shine, R. & Lourdais, O. Taxonomic chauvinism. Trends Ecol. Evol. 17, 1–3 (2002).
    Article  Google Scholar 

    61.
    Tingley, R., Meiri, S. & Chapple, D. G. Addressing knowledge gaps in reptile conservation. Biol. Cons. 204, 1–5 (2016).
    Article  Google Scholar 

    62.
    Beier, P., Majka, D. & Jenness, J. Conceptual Steps for Designing Wildlife Corridors. www.corridordesign.org (2007).

    63.
    Valencia-Aguilar, A., Cortés-Gómez, A. M. & Ruiz-Agudelo, C. A. Ecosystem services provided by amphibians and reptiles in Neotropical ecosystems. Int. J. Biodivers. Sci. Ecosyst. Serv. Manag. 9, 257–272 (2013).
    Article  Google Scholar 

    64.
    Hager, H. A. Area-sensitivity of reptiles and amphibians: Are there indicator species for habitat fragmentation?. Écoscience 66, 139–147 (1998).
    Article  Google Scholar 

    65.
    Almasieh, K., Mirghazanfari, S. M. & Mahmoodi, S. Biodiversity hotspots for modeled habitat patches and corridors of species richness and threatened species of reptiles in central Iran. Eur. J. Wildl. Res. 65, 92. https://doi.org/10.1007/s10344-019-1335-x (2019).
    Article  Google Scholar 

    66.
    Albert, C., Luque, G. M. & Courchamp, F. The twenty most charismatic species. PLoS ONE 13, e0199149. https://doi.org/10.1371/journal.pone.0199149 (2018).
    CAS  Article  Google Scholar 

    67.
    Brooke, Z. M., Bielby, J., Nambiar, K. & Carbone, C. correlates of research effort in carnivores: Body size, range size and diet matter. PLoS ONE 9, e93195. https://doi.org/10.1371/journal.pone.0093195 (2014).
    ADS  CAS  Article  Google Scholar 

    68.
    Rozylowicz, L., Popescu, V. D., Pătroescu, M. & Chișamera, G. The potential of large carnivores as conservation surrogates in the Romanian Carpathians. Biodivers. Conserv. 20, 561–579 (2011).
    Article  Google Scholar 

    69.
    Beier, P., Majka, D. R. & Spencer, W. D. Forks in the road choices in procedures for designing wildland linkages. Conserv. Biol. 22, 836–851 (2008).
    Article  Google Scholar 

    70.
    Burbrink, F. T., Phillips, C. A. & Heske, E. J. A riparian zone in southern Illinois as a potential dispersal corridor for reptiles and amphibians. Biol. Cons. 86, 107–115 (1998).
    Article  Google Scholar 

    71.
    Dixo, M. & Metzger, J. P. Are corridors, fragment size and forest structure important for the conservation of leaf-litter lizards in a fragmented landscape?. Oryx 43, 435–442 (2009).
    Article  Google Scholar 

    72.
    ArcGIS Release 10.4 (Redlands, CA, 2013).

    73.
    Hamer, A. J. & McDonnell, M. J. The response of herpetofauna to urbanization: Inferring patterns of persistence from wildlife databases. Austral Ecol. 35, 568–580 (2010).
    Article  Google Scholar 

    74.
    Vignoli, L., Mocaer, I., Luiselli, L. & Bologna, M. A. Can a large metropolis sustain complex herpetofauna communities? An analysis of the suitability of green space fragments in Rome. Anim. Conserv. 12, 456–466 (2009).
    Article  Google Scholar 

    75.
    Strugariu, A., Gherghel, I., Huțuleac-Volosciuc, M. V. & Pușcașu, C. M. Preliminary aspects concerning the herpetofauna from urban and peri-urban environments from North-Eastern Romania: A case study in the city of Suceava. Herpetol. Roman. 1, 53–61 (2007).
    Google Scholar 

    76.
    Gherghel, I., Strugariu, A., Sahlean, T. C. & Zamfirescu, O. Anthropogenic impact or anthropogenic accommodation? Distribution range expansion of the common wall lizard (Podarcis muralis) by means of artificial habitats in the north-eastern limits of its distribution range. Acta Herpetol. 4, 183–189 (2009).
    Google Scholar 

    77.
    Gherghel, I. & Tedrow, R. Manmade structures are used by an invasive species to colonize new territory across a fragmented landscape. Acta Oecol. 101, 103479. https://doi.org/10.1016/j.actao.2019.103479 (2019).
    Article  Google Scholar 

    78.
    Ward, M. et al. Just ten percent of the global terrestrial protected area network is structurally connected via intact land. Nat. Commun. 11, 4563. https://doi.org/10.1038/s41467-020-18457-x (2020).
    CAS  Article  Google Scholar 

    79.
    Watson, J. E. M. et al. The exceptional value of intact forest ecosystems. Nat. Ecol. Evol. 2, 599–610 (2018).
    Article  Google Scholar 

    80.
    Fitzgerald, L. A. et al. The future for reptiles: advances and challenges in the anthropocene. Encycl. Anthropocene 3, 163–174 (2018).
    Article  Google Scholar 

    81.
    Hof, C., Araújo, M. B., Jetz, W. & Rahbek, C. Additive threats from pathogens, climate and land-use change for global amphibian diversity. Nature 480, 516–519 (2011).
    ADS  CAS  Article  Google Scholar 

    82.
    Meta-Analysis, A. Rey Benayas, J. M., Newton, A. C., Diaz, A. & Bullock, J. M. Enhancement of biodiversity and ecosystem services by ecological restoration. Science 325, 1121–1124 (2009).
    Article  CAS  Google Scholar 

    83.
    Van Der Windt, H. J. & Swart, J. A. A. Ecological corridors, connecting science and politics: the case of the Green River in the Netherlands. J. Appl. Ecol. 45, 124–132 (2008).
    Article  Google Scholar 

    84.
    Hilty, J. et al. Guidelines for conserving connectivity through ecological networks and corridors (International Union for Conservation of Nature, 2020).

    85.
    Gregory, A. J. & Beier, P. Response variables for evaluation of the effectiveness of conservation corridors. Conserv. Biol. 28, 689–695 (2014).
    Article  Google Scholar 

    86.
    Mráz, P. & Ronikier, M. Biogeography of the Carpathians: Evolutionary and spatial facets of biodiversity. Biol. J. Lin. Soc. 119, 528–559 (2016).
    Article  Google Scholar 

    87.
    Deodatus, F. et al. Creation of ecological corridors in the Ukrainian Carpathians. In The Carpathians: Integrating Nature and Society Towards Sustainability Environmental Science and Engineering (eds Kozak, J. et al.) 701–717 (Springer, Berlin, 2013).
    Google Scholar 

    88.
    Favilli, F., Hoffmann, C., Elmi, M., Ravazzoli, E. & Streifeneder, T. The BioREGIO Carpathians project: Aims, methodology and results from the “Continuity and Connectivity” analysis. Nat. Conserv. 11, 95–111 (2015).
    Article  Google Scholar 

    89.
    Csagoly, P., Magnin, G. & Hulea, O. Lower Danube Green Corridor. in The Wetland Book: II: Distribution, Description and Conservation (eds. Finlayson, M. C., Milton, R. G., Prentice, C. R. & Davidson, N. C.) 1–6 (Springer, Netherlands, 2016).

    90.
    Belote, R. T. et al. Identifying corridors among large protected areas in the United States. PLoS ONE 11, e0154223. https://doi.org/10.1371/journal.pone.0154223 (2016).
    CAS  Article  Google Scholar 

    91.
    Breckheimer, I. et al. Defining and evaluating the umbrella species concept for conserving and restoring landscape connectivity. Conserv. Biol. 28, 1584–1593 (2014).
    Article  Google Scholar 

    92.
    Meurant, M., Gonzales, A., Doxa, A. & Albert, C. H. Selecting surrogate species for connectivity conservation. Biol. Cons. 227, 326–334 (2018).
    Article  Google Scholar 

    93.
    Dondina, O., Orioli, V., Chiatante, G. & Bani, L. Practical insights to select focal species and design priority areas for conservation. Ecol. Indic. 108, 105767. https://doi.org/10.1016/j.ecolind.2019.105767 (2020).
    Article  Google Scholar 

    94.
    Churko, G., Kienast, F. & Bolliger, J. A multispecies assessment to identify the functional connectivity of amphibians in a human-dominated landscape. Int. J. Geo-Inf. 9, 287. https://doi.org/10.3390/ijgi9050287 (2020).
    Article  Google Scholar 

    95.
    Cushman, S. A. & Landguth, E. L. Multi-taxa population connectivity in the Northern Rocky Mountains. Ecol. Model. 231, 101–112 (2012).
    Article  Google Scholar 

    96.
    Krosby, M. et al. Focal species and landscape “naturalness” corridor models offer complementary approaches for connectivity conservation planning. Landsc. Ecol. 30, 2121–2132 (2015).
    Article  Google Scholar 

    97.
    Wiens, J. A., Hayward, G. D., Holthausen, R. S. & Wisdom, M. J. Using surrogate species and groups for conservation planning and management. Bioscience 58, 241–252 (2008).
    Article  Google Scholar 

    98.
    Macdonald, E. A. et al. Identifying ambassador species for conservation marketing. Glob. Ecol. Conserv. 12, 204–214 (2017).
    Article  Google Scholar 

    99.
    Fleury, A. M. & Brown, R. D. A framework for the design of wildlife conservation corridors with specific application to southwestern Ontario. Landsc. Urban Plan. 37, 163–186 (1997).
    Article  Google Scholar 

    100.
    Cogălniceanu, D. et al. Diversity and distribution of amphibians in Romania. ZooKeys 296, 35–57 (2013).
    Article  Google Scholar 

    101.
    Cogălniceanu, D. et al. Diversity and distribution of reptiles in Romania. ZooKeys 341, 49–76 (2013).
    Article  Google Scholar 

    102.
    LaRue, M. A. & Nielsen, C. K. Modelling potential dispersal corridors for cougars in midwestern North America using least-cost path methods. Ecol. Model. 212, 372–381 (2008).
    Article  Google Scholar 

    103.
    Adriaensen, F. et al. The application of ‘least-cost’ modelling as a functional landscape model. Landsc. Urban Plan. 64, 233–247 (2003).
    Article  Google Scholar 

    104.
    Correa Ayram, C. A., Mendoza, M. E., Etter, A. & Salicrup, D. R. P. Habitat connectivity in biodiversity conservation: a review of recent studies and applications. Prog. Phys. Geogr. 1, 1–32 (2015).
    Google Scholar 

    105.
    Ribeiro, J. W. et al. LandScape Corridors (LSCORRIDORS): A new softwarepackage for modelling ecological corridors based onlandscape patterns and species requirements. Methods Ecol. Evol. 8, 1425–1432 (2017).
    Article  Google Scholar 

    106.
    Linkage Mapper Connectivity Analysis Software v. 2.0.0 (The Nature Conservancy, Seattle, 2011).

    107.
    Popescu, V. D., Rozylowicz, L., Cogălniceanu, D., Niculae, I. M. & Cucu, A. L. Moving into protected areas? Setting conservation priorities for Romanian Reptiles and Amphibians at risk from climate change. PLoS ONE 8, e79330. https://doi.org/10.1371/journal.pone.0079330 (2014).
    ADS  CAS  Article  Google Scholar 

    108.
    Lambeck, R. J. Focal species: A multispecies umbrella for nature conservation. Conserv. Biol. 11, 849–856 (1997).
    Article  Google Scholar  More