More stories

  • in

    Biogeography of the cosmopolitan terrestrial diatom Hantzschia amphioxys sensu lato based on molecular and morphological data

    In most of the forest soil samples used in this survey, specimens belonging to the genus Hantzschia are quite common. Based molecular as well as on light microscopy (LM) and scanning electron microscopy (SEM) observations of 25 strains, seven different taxa were recognized. Figure 1 contains the locations of the strain’s habitats. In anticipation of the nomenclatural consequences, we are using the new names already here but will describe them formally later.
    Figure 1

    Map with the habitat locations of the studied strains.

    Full size image

    Molecular data
    The obtained phylogenetic tree for representatives of the different strains Hantzschia contains several large clades, some of which are monophyletic, while others contain several different species names (Fig. 2). In the analyzed tree, the largest clade is represented by different strains of H. amphioxys, the structure of which is described in the corresponding molecular analysis section. At the same time, the most significant is that in the same clade there is strain H. amphioxys D27_008, which has been designated as epitype20. One of the largest is the clade with H. abundans, which, in addition to our strains, and some that have already been published, includes the group of strains referred to as “Hantzschia sp. 3” (Sterre6)e, (Sterre6)f from Souffreau et al.16. We propose to refer to all of these strains as H. abundans. The next clade consists of the new species of H. attractiva and three strains of Hantzschia sp. 2 (Mo1)a, (Mo1)e, (Mo1)m from Souffreau et al.16, the latter we propose to merge into the new species named H. pseudomongolica, which is sister to H. attractiva. Given the topology of the tree and the morphological features of the representatives, we can conclude that there is a close relation between H. abundans and H. attractiva plus H. pseudomongolica. A separate group consists of two clades with sufficient statistical support (likelihood bootstrap, LB 76; posterior probability, PP 100), one of which is represented by two strains of H. parva, and the other with strains of H. cf. amphioxys (Sterre1)f, (Sterre1)h. Another large clade represents a set of strains of Hantzschia sp. 1 and Hantzschia sp. 2 (Mo1)h, (Mo1)l from Souffreau et al.16, among which there are both large cells (86–89 µm length) and smaller ones (37–39 µm length); strains also differ by striation – from 18–20 striae in 10 μm (strain (Mo1)h) to 21–22 in 10 μm (strain (Ban1)h). It is possible that Hantzschia sp. 1 and Hantzschia sp. 2 (Mo1)h, (Mo1)l may be several closely related species. Besides the large clades, there are a number of separate branches in the tree, representing separate strains: Hantzschia sp. 1 (Ban1)d, and the new species H. belgica (H. cf. amphioxys (Sterre3)a from Souffreaua et al.16) and H. stepposa. Interesting is the position of the H. abundans (Tor3)c strain, which is very distant from other representatives of H. abundans and probably is a cryptic taxon, whose taxonomic status needs to be revised.
    Figure 2

    Bayesian tree for representatives of the different strains Hantzschia, from an alignment with 40 sequences and 1785 characters (partial rbcL gene and 28S rDNA fragments). Type strains indicated in bold. The epitype of Hantzschia amphioxys is underlined. Values above the horizontal lines (on the left of slash) are bootstrap support from RAxML analyses ( More

  • in

    Endocranial volume is variable and heritable, but not related to fitness, in a free-ranging primate

    1.
    Healy, S. D. & Rowe, C. A critique of comparative studies of brain size. Proc. R. Soc. B Biol. Sci. 274, 453–464 (2007).
    Article  Google Scholar 
    2.
    Roth, G. & Dicke, U. Evolution of the brain and intelligence. Trends Cogn. Sci. 9, 250–257 (2005).
    PubMed  Article  Google Scholar 

    3.
    Logan, C. J., Kruuk, L. E. B., Stanley, R., Thompson, A. M. & Clutton-Brock, T. H. Endocranial volume is heritable and is associated with longevity and fitness in a wild mammal. R. Soc. Open Sci. 3, 160622 (2016).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    4.
    Dunbar, R. I. M. Neocortex size as a constraint on group size in primates. J. Hum. Evol. 22, 469–493 (1992).
    Article  Google Scholar 

    5.
    Innocenti, G. M. & Kaas, J. H. The cortex. Trends Neurosci. 18, 371–372 (1995).
    CAS  Article  Google Scholar 

    6.
    Kaas, J. H. The evolution of isocortex. Brain. Behav. Evol. 46, 187–196 (1995).
    CAS  PubMed  Article  Google Scholar 

    7.
    Barton, R. A. & Harvey, P. H. Mosaic evolution of brain structure in mammals. Nature 405, 1055–1058 (2000).
    ADS  CAS  PubMed  Article  Google Scholar 

    8.
    Reader, S. M. & Laland, K. N. Social intelligence, innovation, and enhanced brain size in primates. Proc. Natl. Acad. Sci. 99, 4436–4441 (2002).
    ADS  CAS  PubMed  Article  Google Scholar 

    9.
    Sol, D., Székely, T., Liker, A. & Lefebvre, L. Big-brained birds survive better in nature. Proc. R. Soc. B Biol. Sci. 274, 763–769 (2007).
    Article  Google Scholar 

    10.
    Benson-Amram, S., Dantzer, B., Stricker, G., Swanson, E. M. & Holekamp, K. E. Brain size predicts problem-solving ability in mammalian carnivores. Proc. Natl. Acad. Sci. USA 113, 2532–2537 (2016).
    ADS  CAS  PubMed  Article  Google Scholar 

    11.
    Cartmill, M. New views on primate origins. Evol. Anthropol. Issues News Rev. 1, 105–111 (2005).
    Article  Google Scholar 

    12.
    Allman, J., McLaughlin, T. & Hakeem, A. Brain weight and life-span in primate species. Proc. Natl. Acad. Sci. 90, 118–122 (1993).
    ADS  CAS  PubMed  Article  Google Scholar 

    13.
    González-Lagos, C., Sol, D. & Reader, S. M. Large-brained mammals live longer. J. Evol. Biol. 23, 1064–1074 (2010).
    PubMed  Article  Google Scholar 

    14.
    Harvey, P. H. & Bennett, P. M. Evolutionary biology: Brain size, energetics, ecology and life history patterns. Nature 306, 314–315 (1983).
    ADS  CAS  PubMed  Article  Google Scholar 

    15.
    Aiello, L. C. & Wheeler, P. The expensive-tissue hypothesis: The brain and the digestive system in human and primate evolution. Curr. Anthropol. 36, 199–221 (1995).
    Article  Google Scholar 

    16.
    Kudo, H. & Dunbar, R. I. M. Neocortex size and social network size in primates. Anim. Behav. 62, 711–722 (2001).
    Article  Google Scholar 

    17.
    Schillaci, M. A. Sexual selection and the evolution of brain size in primates. PLoS ONE 1, e62 (2006).
    ADS  PubMed  PubMed Central  Article  Google Scholar 

    18.
    Shultz, S. & Dunbar, R. I. M. The evolution of the social brain: anthropoid primates contrast with other vertebrates. Proc. R. Soc. B Biol. Sci. 274, 2429–2436 (2007).
    Article  Google Scholar 

    19.
    King, B. J. Extractive foraging and the evolution of primate intelligence. Hum. Evol. 1, 361–372 (1986).
    Article  Google Scholar 

    20.
    Barton, R. A. Neocortex size and behavioural ecology in primates. Proc. R. Soc. Lond. B 263, 173–177 (1996).
    ADS  CAS  Article  Google Scholar 

    21.
    DeCasien, A. R., Williams, S. A. & Higham, J. P. Primate brain size is predicted by diet but not sociality. Nat. Ecol. Evol. 1, 0112 (2017).
    Article  Google Scholar 

    22.
    Powell, L. E., Isler, K. & Barton, R. A. Re-evaluating the link between brain size and behavioural ecology in primates. Proc. R. Soc. B Biol. Sci. 284, 20171765 (2017).
    Article  Google Scholar 

    23.
    Dunbar, R. I. M. & Shultz, S. Why are there so many explanations for primate brain evolution?. Philos. Trans. R. Soc. B Biol. Sci. 372, 20160244 (2017).
    Article  Google Scholar 

    24.
    Van Schaik, C. P. Why are diurnal primates living in groups?. Behaviour 87, 120–144 (1983).
    Article  Google Scholar 

    25.
    Van Schaik, C. P. & Van Hooff, J. A. R. A. M. On the ultimate causes of primate social systems. Behaviour 85, 91–117 (1983).
    Article  Google Scholar 

    26.
    Wrangham, R. W. An ecological model of female-bonded primate groups. Behaviour 75, 262–300 (1980).
    Article  Google Scholar 

    27.
    Atchley, W. R., Riska, B., Kohn, L. A. P., Plummer, A. A. & Rutledge, J. J. A quantitative genetic analysis of brain and body size associations, their origin and ontogeny: Data from mice. Evolution 38, 1165 (1984).
    PubMed  Article  Google Scholar 

    28.
    Riska, B. & Atchley, W. R. Genetics of growth predict patterns of brain-size evolution. Science 229, 668–671 (1985).
    ADS  CAS  PubMed  Article  Google Scholar 

    29.
    Rogers, J. et al. Heritability of brain volume, surface area and shape: An MRI study in an extended pedigree of baboons. Hum. Brain Mapp. 28, 576–583 (2007).
    PubMed  PubMed Central  Article  Google Scholar 

    30.
    Gómez-Robles, A., Hopkins, W. D., Schapiro, S. J. & Sherwood, C. C. Relaxed genetic control of cortical organization in human brains compared with chimpanzees. Proc. Natl. Acad. Sci. 112, 14799–14804 (2015).
    ADS  PubMed  Article  CAS  Google Scholar 

    31.
    DeCasien, A. R., Sherwood, C. C., Schapiro, S. J. & Higham, J. P. Greater variability in chimpanzee (Pan troglodytes) brain structure among males. Proc. R. Soc. B 287, 20192858 (2020).
    PubMed  Article  Google Scholar 

    32.
    Fears, S. C. et al. Identifying heritable brain phenotypes in an extended pedigree of vervet monkeys. J. Neurosci. 29, 2867–2875 (2009).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    33.
    Noreikiene, K. et al. Quantitative genetic analysis of brain size variation in sticklebacks: Support for the mosaic model of brain evolution. Proc. R. Soc. B Biol. Sci. 282, 20151008 (2015).
    Article  Google Scholar 

    34.
    Kotrschal, A. et al. Artificial selection on relative brain size in the guppy reveals costs and benefits of evolving a larger brain. Curr. Biol. 23, 168–171 (2013).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    35.
    Cheverud, J. M. et al. Heritability of brain size and surface features in rhesus macaques (Macaca mulatta). J. Hered. 81, 51–57 (1990).
    CAS  PubMed  Article  Google Scholar 

    36.
    de Villemereuil, P. Tutorial estimation of a biological trait heritability using the animal model How to use the MCMCglmm R package. (2012).

    37.
    Axelrod, C. J., Laberge, F. & Robinson, B. W. Intraspecific brain size variation between coexisting sunfish ecotypes. Proc. R. Soc. B Biol. Sci. 285, 20181971 (2018).
    Article  Google Scholar 

    38.
    Blomquist, G. E. Fitness-related patterns of genetic variation in rhesus macaques. Genetica 135, 209–219 (2009).
    PubMed  Article  Google Scholar 

    39.
    Brent, L. J. N. et al. Personality traits in rhesus macaques (Macaca mulatta) are heritable but do not predict reproductive output. Int. J. Primatol. 35, 188–209 (2014).
    PubMed  Article  Google Scholar 

    40.
    Dubuc, C. et al. Sexually selected skin colour is heritable and related to fecundity in a non-human primate. Proc. R. Soc. B Biol. Sci. 281, 20141602 (2014).
    Article  Google Scholar 

    41.
    Kimock, C. M., Dubuc, C., Brent, L. J. N. & Higham, J. P. Male morphological traits are heritable but do not predict reproductive success in a sexually-dimorphic primate. Sci. Rep. 9, 19794 (2019).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    42.
    Kruuk, L. E. B. Estimating genetic parameters in natural populations using the ‘animal model’. Philos. Trans. R. Soc. B 359, 873–890 (2004).
    Article  Google Scholar 

    43.
    Falk, D., Froese, N., Sade, D. S. & Dudek, B. C. Sex differences in brain/body relationships of Rhesus monkeys and humans. J. Hum. Evol. 36, 233–238 (1999).
    CAS  PubMed  Article  Google Scholar 

    44.
    Herndon, J. G., Tigges, J., Anderson, D. C., Klumpp, S. A. & McClure, H. M. Brain weight throughout the life span of the chimpanzee. J. Comp. Neurol. 409, 567–572 (1999).
    CAS  PubMed  Article  Google Scholar 

    45.
    Iwaniuk, A. N. Interspecific variation in sexual dimorphism in brain size in Nearctic ground squirrels (Spermophilus spp.). Can. J. Zool. 79, 759–765 (2001).
    Article  Google Scholar 

    46.
    Towe, A. L. & Mann, M. D. Habitat-related variations in brain and body size of pocket gophers. J. Hirnforsch. 36, 195–201 (1995).
    CAS  PubMed  Google Scholar 

    47.
    Kotrschal, A., Räsänen, K., Kristjánsson, B. K., Senn, M. & Kolm, N. Extreme sexual brain size dimorphism in sticklebacks: A consequence of the cognitive challenges of sex and parenting?. PLoS ONE 7, e30055 (2012).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    48.
    Ritchie, S. J. et al. Sex differences in the adult human brain: Evidence from 5216 uk biobank participants. Cereb. Cortex 28, 2959–2975 (2018).
    PubMed  PubMed Central  Article  Google Scholar 

    49.
    Whitten, P. L. Diet and dominance among female vervet monkeys (Cercopithecus aethiops). Am. J. Primatol. 5, 139–159 (1983).
    PubMed  Article  Google Scholar 

    50.
    Mori, A. Analysis of population changes by measurement of body weight in the Koshima troop of Japanese monkeys. Primates 20, 371–397 (1979).
    Article  Google Scholar 

    51.
    Small, M. F. Body fat, rank, and nutritional status in a captive group of Rhesus Macaques. Int. J. Primatol. 2, 91–95 (1981).
    Article  Google Scholar 

    52.
    Sade, D. S. Population dynamics in relation to social structure on Cayo Santiago. Ybk. Phys. Anthr. 20, 253–262 (1976).
    Google Scholar 

    53.
    Silk, J. B., Clark-Wheatley, C. B., Rodman, P. S. & Samuels, A. Differential reproductive success and facultative adjustment of sex ratios among captive female bonnet macaques (Macaca radiata). Anim. Behav. 29, 1106–1120 (1981).
    Article  Google Scholar 

    54.
    Rawlins, R. G. & Kessler, M. J. The Cayo Santiago macaques: History, behavior, and biology (SUNY Series Primatology, Suny, 1986).
    Google Scholar 

    55.
    Kessler, M. J. & Rawlins, R. G. A 75-year pictorial history of the Cayo Santiago rhesus monkey colony. Am. J. Primatol. 78, 6–43 (2016).
    PubMed  Article  Google Scholar 

    56.
    Widdig, A. et al. Genetic studies on the Cayo Santiago rhesus macaques: A review of 40 years of research. Am. J. Primatol. 78, 44–62 (2016).
    PubMed  Article  Google Scholar 

    57.
    Widdig, A. et al. Low incidence of inbreeding in a long-lived primate population isolated for 75 years. Behav. Ecol. Sociobiol. 71, 18 (2017).
    PubMed  Article  Google Scholar 

    58.
    Cheverud, J. M. Epiphyseal union and dental eruption in Macaca mulatta. Am. J. Phys. Anthropol. 56, 157–167 (1981).
    CAS  PubMed  Article  Google Scholar 

    59.
    Turnquist, J. E. & Kessler, M. J. Free-ranging Cayo Santiago rhesus monkeys (Macaca mulatta): I. Body size, proportion, and allometry. Am. J. Primatol. 19, 1–13 (1989).
    PubMed  Article  Google Scholar 

    60.
    Havill, L. M. Osteon remodeling dynamics in macaca mulatta: Normal variation with regard to age, sex, and skeletal maturity. Calcif. Tissue Int. 74, 95–102 (2004).
    CAS  PubMed  Article  Google Scholar 

    61.
    Konigsberg, L. et al. External brain morphology in rhesus macaques (Macaca mulatta). J. Hum. Evol. 19, 269–284 (1990).
    Article  Google Scholar 

    62.
    Logan, C. J. & Clutton-Brock, T. H. Validating methods for estimating endocranial volume in individual red deer (Cervus elaphus). Behav. Process. 92, 143–146 (2013).
    Article  Google Scholar 

    63.
    Jolly, C. The classification and natural history of Theropithecus (Simopithecus) (Andrew, 1916) baboons of the African Plio-Pleistocene. (Bull. Brit. Mus. Nat. Hist., 1972).

    64.
    Delson, E. et al. Body mass in Cercopithecidae (Primates, mammalia): Estimation and scaling in extinct and extant taxa. (American Museum of Natural History, 2000).

    65.
    Hadfield, J. D., Richardson, D. S. & Burke, T. Towards unbiased parentage assignment: Combining genetic, behavioural and spatial data in a Bayesian framework. Mol. Ecol. 15, 3715–3730 (2006).
    CAS  PubMed  Article  Google Scholar 

    66.
    Hadfield, J. D. MCMCglmm Course Notes. (2016).

    67.
    Morrissey, M. B. & Wilson, A. J. pedantics: An r package for pedigree-based genetic simulation and pedigree manipulation, characterization and viewing: Computer program article. Mol. Ecol. Resour. 10, 711–719 (2009).
    PubMed  Article  Google Scholar 

    68.
    Hadfield, J. D. MCMC methods for multi-response generalized linear mixed models: The MCMCglmm R package. J. Stat. Softw. 33, 1–22 (2010).
    Article  Google Scholar 

    69.
    Hadfield, J. D. & Nakagawa, S. General quantitative genetic methods for comparative biology: Phylogenies, taxonomies and multi-trait models for continuous and categorical characters. J. Evol. Biol. 23, 494–508 (2010).
    CAS  PubMed  Article  Google Scholar 

    70.
    Wilson, A. J. et al. An ecologist’s guide to the animal model. J. Anim. Ecol. 79, 13–26 (2010).
    PubMed  Article  Google Scholar 

    71.
    Kuznetsova, A., Brockhoff, P. B. & Christensen, R. H. B. lmerTest package: Tests in linear mixed effects models. J. Stat. Softw. 82, 13 (2017).
    Article  Google Scholar 

    72.
    Lande, R. & Arnold, S. J. The measurement of selection on correlated characters. Evolution 37, 1210–1226 (1983).
    PubMed  Article  Google Scholar 

    73.
    Morrissey, M. B. & Sakrejda, K. Unification of regression-based methods for the analysis of natural selection. Evolution 67, 2094–2100 (2013).
    PubMed  Article  Google Scholar 

    74.
    Stinchcombe, J., Agrawal, A., Hohenlohe, P., Arnold, S. & Blows, M. Estimating nonlinear selection gradients using quadratic regression coefficients: Double or nothing?. Evolution 62, 2435–2440 (2008).
    PubMed  Article  Google Scholar 

    75.
    Matsumura, S., Arlinghaus, R. & Dieckmann, U. Standardizing selection strengths to study selection in the wild: A critical comparison and suggestions for the future. Bioscience 62, 1039–1054 (2012).
    Article  Google Scholar  More

  • in

    Investigating an increase in Florida manatee mortalities using a proteomic approach

    This proteomic survey was conducted to identify proteins that were differentially expressed in the serum of manatees affected by two distinct mortality episodes: a red tide group and an unknown mortality episode group in the IRL. These groups were compared to a control group sampled at Crystal River. The red tide group’s exposure was evidenced by the presence of the PbTx antigen, with brevetoxin values in the 4.3 to 14.4 ng/ml range. The other group did not present with clinical symptoms except for mild cold stress in some animals. Two proteomics approaches were employed, 2D-DIGE and shot gun proteomics using LC–MS/MS, which provided similar results, suggesting that several serum proteins were specifically altered in each of the manatee mortality episode groups compared to the Crystal River control group. The differentially expressed serum proteins were cautiously identified based on annotation of the manatee genome6,7 and their amino acid sequence homologies with human serum proteins. While additional work still needs to be done to confirm that the identified manatee proteins function similarly to their human homologs, possible insight on the function of the proteins can be derived from human studies.
    The two proteomics methods used, 2D-DIGE and iTRAQ LC–MS/MS are complementary and both rely on LC–MS/MS for protein identification. 2D-DIGE is a top-down approach, quantifying the differentially expressed proteins at the protein level before identifying the protein by LC–MS/MS, while the iTRAQ method is a bottom-up approach, where the whole proteome is first digested with trypsin, the generated peptides are separated by chromatography and identified and measured by mass spectrometry. Mass spectrometry has become the primary method to analyze proteomes, benefitting from genomic sequences and bioinformatics tools that can translate the sequences into predicted proteins. There are excellent reviews of proteomics methods and how they may be used across species8,9.
    In total, 19 of the 26 proteins identified using the 2D-DIGE method were also identified by iTRAQ (Supplementary Table 1) which showed that these findings were replicated using two complementary experimental methods. In the 2D-DIGE method, most of the proteins were found in multiple spots, suggesting that they were differentially modified. 2D-DIGE can separate proteins based on a single charge difference. Some of the spots contained multiple proteins so it was difficult to determine the fold change of each of the proteins in these spots. For example, protein C4A was identified in 7 different spots, likely representing multiple isoforms. We were not able to corroborate the different post-translational modifications (PTMs) with iTRAQ, as the experiment was not designed to look for PTMs, only total protein quantitation. A drawback of 2D-DIGE is that keratin introduced into the sample from reagents at the time of electrophoresis or through the multiple steps required for protein extraction is also seen in the gels10,11,12. It is unlikely that the keratins were from the serum samples, as blood was collected directly into vacuum tubes. Because of the issue of keratin contamination, the 2D-DIGE method is considered more qualitative in its determination and thus in this study, iTRAQ data were the primary basis for quantitation.
    Pathway analysis detects groups of proteins that are linked in pathways that may be related to disease processes. We used Pathway Studio using subnetwork enrichment analysis to determine disease pathways potentially in place for the red tide and IRL manatees. The Pathway Studio database is constructed from relationships detected between proteins and diseases from articles present in Pubmed but is heavily directed towards human and rodent proteomes. To be able to use this tool, we assigned human homologs to the identified manatee proteins, assuming that based on their sequence homology the proteins would function in a similar way. There are many studies that suggest this assumption has merit, for example Nonaka and Kimura have examined the evolution of the complement system and found clear indications of homology among vertebrates13.
    The top 20 pathways for the red tide group (Table 3) and the IRL group (Table 4) show the diverse set of molecular pathways that may be affected by the exposures. Many of the same pathways appeared for both groups including thrombophilia, inflammation, wounds and injuries, acute phase reaction and amyloidosis. Thrombophilia was the most upregulated pathway for the IRL group (p-value 1.10E-19) and the second most upregulated pathway for the red tide group (p-value 4.1E-19). Thrombophilia, a condition in which blood clots occur in the absence of injury, happens when clotting factors become unbalanced. We obtained proteomics information on 12 of the proteins in this pathway, with some moving in opposing directions. The dysregulated proteins that were increased for both red tide and the IRL groups were SERPIN D1 (Serpin family member D 1), CRP (C-reactive protein), and PLAT (plasminogen activator) and the ones that were decreased in both groups, were SERPIN C1 (Serpin family member C 1), F5 (coagulation factor 5), and ALB (albumin). One protein, AGT (angiotensinogen), was upregulated in the red tide group but downregulated in the IRL. HRG (histidine rich glycoprotein), PROS1 (Protein S), C4BPA (complement component 4 binding protein alpha, and F2 (coagulation factor 2, also known as prothrombin) were downregulated in the red tide group but upregulated in the IRL group. The disparate regulation of proteins in this pathway suggests that clotting was among the pathways disrupted in the affected manatees. Red tide exposed manatees often present with hemorrhagic issues in their intestines, lungs and the brain (14), suggesting that downregulation of coagulation factors may be responsible for this clinical evaluation. Interestingly HRG was upregulated in the IRL by 1.34-fold and downregulated in the red tide group by 0.56-fold, making this protein a good biomarker to distinguish the two events.
    Table 3 Subnetwork enrichment pathways for serum proteins obtained from manatees exposed to red tide.
    Full size table

    Table 4 Subnetwork enrichment pathways for serum proteins obtained from manatees sampled in the IRL.
    Full size table

    Among the manatees in the red tide group, inflammation was ranked 3rd (p-value  More

  • in

    Evolutionary history and genetic connectivity across highly fragmented populations of an endangered daisy

    Aægisdóttir HH, Kuss P, Stöcklin J (2009) Isolated populations of a rare alpine plant show high genetic diversity and considerable population differentiation. Ann Bot 104:1313–1322
    Article  CAS  Google Scholar 

    Ahrens CW, James EA, Botanic R, Melbourne G, Ave B, Yarra S (2015) Range-wide genetic analysis reveals limited structure and suggests asexual patterns in the rare forb Senecio macrocarpus. Biol J Linn Soc 115:256–269
    Article  Google Scholar 

    Bouckaert R (2010) DensiTree: making sense of sets of phylogenetic trees. Bioinformatics 26:1372–137
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Bouckaert R, Vaughan TG, Barido-Sottani J, Duchene S, Fourmet M, Gavryushkina A et al. (2019) BEAST 2.5: an advanced software platform for Bayesian evolutionary analysis. PLoS Comput Biol 15:1–28
    Article  CAS  Google Scholar 

    Bowler J (1982) Aridity in the late tertiary and quaternary of Australia. In: Barker W, Greenslade P (eds) Evolution of the flora and fauna of arid Australia. Peacock Publications, Adelaide, p 35–45
    Google Scholar 

    Breed MF, Harrison PA, Blyth C, Byrne M, Gaget V, Gellie NJC et al. (2019) The potential of genomics for restoring ecosystems and biodiversity. Nat Rev Genet 20:615–628
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Brown AHD, Young AG (2000) Genetic diversity in tetraploid populations of the endangered daisy Rutidosis leptorrhynchoides and implications for its conservation. Heredity (Edinb) 85:122–129
    CAS  Article  Google Scholar 

    Bryant D, Bouckaert R, Felsenstein J, Rosenberg NA, Roychoudhury A (2012) Inferring species trees directly from biallelic genetic markers: bypassing gene trees in a full coalescent analysis. Mol Biol Evol 29:1917–1932
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Bull M, Stolfo G (2014) Flora of Melbourne. A guide to the indigenous plants of the greater Melbourne area, 4th edn. Hyland House, Melbourne
    Google Scholar 

    Buza L, Young A, Thrall P (2000) Genetic erosion, inbreeding and reduced fitness in fragmented populations of the endangered tetraploid pea Swainsona recta. Biol Conserv 93:177–186
    Article  Google Scholar 

    Charlesworth D (2006) Balancing selection and its effects on sequences in nearby genome regions. PLoS Genet 2:379–384
    CAS  Article  Google Scholar 

    Chen C, Lu RS, Zhu SS, Tamaki I, Qiu YX (2017) Population structure and historical demography of Dipteronia dyeriana (Sapindaceae), an extremely narrow palaeoendemic plant from China: implications for conservation in a biodiversity hot spot. Heredity (Edinb) 119:95–106
    CAS  Article  Google Scholar 

    Clarke GM, O’Dwyer C (2000) Genetic variability and population structure of the endangered golden sun moth, Synemon plana. Biol Conserv 92:371–381
    Article  Google Scholar 

    Cole CT (2003) Genetic variation in rare and common plants. Annu Rev Ecol Evol Syst 34:213–237
    Article  Google Scholar 

    Coleman RA, Weeks AR, Hoffmann AA (2013) Balancing genetic uniqueness and genetic variation in determining conservation and translocation strategies: a comprehensive case study of threatened dwarf galaxias, Galaxiella pusilla (Mack) (Pisces: Galaxiidae). Mol Ecol 22:1820–1835
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Courtice B, Hoebee SE, Sinclair S, Morgan JW (2020) Local population density affects pollinator visitation in the endangered grassland daisy Rutidosis leptorhynchoides (Asteraceae). Aust J Bot 67:638–648
    Article  Google Scholar 

    Crandall KA, Bininda-Emonds ORP, Mace GM, Wayne RK (2000) Considering evolutionary processes in conservation biology. TREE 15:290–295
    CAS  PubMed  PubMed Central  Google Scholar 

    Delph LF, Kelly JK (2014) On the importance of balancing selection in plants. N Phytol 201:45–56
    Article  Google Scholar 

    DeMauro MM (1993) Relationship of breeding system to rarity in the Lakeside Daisy (Hymenoxys acaulis var. glabra). Conserv Biol 7:542–550
    Article  Google Scholar 

    Department of the Environment (2020) Senecio macrocarpus in Species Profile and Threats Database, Department of the Environment, Canberra. Available from: http://www.environment.gov.au/sprat. Accessed 27 May 2020.

    Diekmann OE, Gouveia L, Perez JA, Gil-Rodriguez C, Serrão EA (2010) The possible origin of Zostera noltii in the Canary Islands and guidelines for restoration. Mar Biol 157:2109–2115
    Article  Google Scholar 

    Dorrough J, Ash JE (1999) Using past and present habitat to predict the current distribution and abundance of a rare cryptic lizard, Delma impar (Pygopodidae). Austral Ecol 24:614–624
    Article  Google Scholar 

    Drummond AJ, Rambaut A (2007) BEAST: Bayesian evolutionary analysis by sampling trees. BMC Evol Biol 7:214
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    Ellstrand NC, Elam DR (1993) Population genetic consequences of small population size: implications for plant conservation. Annu Rev Ecol Syst 24:217–241

    Evanno G, Regnaut S, Goudet J (2005) Detecting the number of clusters of individuals using the software STRUCTURE: a simulation study. Mol Ecol 14:2611–2620
    CAS  PubMed  PubMed Central  Google Scholar 

    Foll M, Gaggiotti OE (2006) Identifying the environmental factors that determine the genetic structure of populations. Genetics 174:875–891
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Foll M, Gaggiotti O (2008) A genome-scan method to identify selected loci appropriate for both dominant and codominant markers: a Bayesian perspective. Genetics 180:977–993
    PubMed  PubMed Central  Article  Google Scholar 

    Frankham R (1996) Relationship between genetic variation and populations size in wildlife. Conserv Biol 10:1500–1508
    Article  Google Scholar 

    Frankham R (2005) Genetics and extinction. Biol Conserv 126:131–140

    Frankham R (2015) Genetic rescue of small inbred populations: meta-analysis reveals large and consistent benefits of gene flow. Mol Ecol 24:2610–2618
    PubMed  Article  PubMed Central  Google Scholar 

    Frankham R, Ballou JD, Eldridge MDB, Lacy RC, Ralls K, Dudash MR et al. (2011) Predicting the probability of outbreeding depression. Conserv Biol 25:465–475
    PubMed  Article  PubMed Central  Google Scholar 

    Frankham R, Ballou JD, Ralls K, Eldridge MDB, Dudash MR, Fenster CB, et al. (2017) Genetic management of fragmented animal and plant populations, 1st edn. Oxford University Press, Oxford

    Frankham R, Bradshaw CJA, Brook BW (2014) Genetics in conservation management: Revised recommendations for the 50/500 rules, Red List criteria and population viability analyses. Biol Conserv 170:56–63
    Article  Google Scholar 

    Frankham R, Lees K, Montgomery ME, England PR, Lowe EH, Briscoe DA (1999) Do population size bottlenecks reduce evolutionary potential? Anim Conserv 2:255–260
    Article  Google Scholar 

    Georges A, Gruber B, Pauly GB, White D, Adams M, Young MJ et al. (2018) Genomewide SNP markers breathe new life into phylogeography and species delimitation for the problematic short-necked turtles (Chelidae: Emydura) of eastern Australia. Mol Ecol 27:5195–5213
    PubMed  Article  PubMed Central  Google Scholar 

    Glémin S, Gaude T, Guillemin ML, Lourmas M, Olivieri I, Mignot A (2005) Balancing selection in the wild: testing population genetics theory of self-incompatibility in the rare species Brassica insularis. Genetics 171:279–289
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    Goudet J (2005) HIERFSTAT, a package for R to compute and test hierarchical F‐statistics. Mol Ecol Resour 5:184–186
    Article  Google Scholar 

    Gruber B, Unmack PJ, Berry OF, Georges A (2018) DARTR: an R package to facilitate analysis of SNP data generated from reduced representation genome sequencing. Mol Ecol Resour 18:691–699
    PubMed  Article  PubMed Central  Google Scholar 

    Jaccoud D, Peng K, Feinstein D, Kilian A (2001) Diversity arrays: a solid state technology for sequence information dependent genotyping. Nucl Acids Res 29:e25
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Janes JK, Malenfant M, Andrew RL, Miller JM, Dupuis JR, Gorrell JC et al. (2017) The K = 2 conundrum. Mol Ecol 26:3594–3602
    PubMed  Article  PubMed Central  Google Scholar 

    Jones RN (1997) The biogeography of the grasses and lowland grasslands of south-eastern Australia. Adv Nat Conserv 2:11–18
    Google Scholar 

    Kamvar ZN, Brooks JC, Grünwald NJ (2015) Novel R tools for analysis of genome-wide population genetic data with emphasis on clonality. Front Genet 6:1–10
    CAS  Article  Google Scholar 

    Knapp EE, Rice KJ (1996) Genetic structure and gene flow in Elymus glaucus (blue wildrye): implications for native grassland restoration. Restor Ecol 4:1–10
    Article  Google Scholar 

    Kopelman NM, Mayzel J, Jakobsson M, Rosenberg NA, Ro AY (2015) CLUMPAK: a program for identifying clustering modes and packaging population structure inferences across K. Mol Ecol Resour 15:1179–1191
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Kronenberger JA, Funk WC, Smith JW, Fitzpatrick SW, Angeloni LM, Broder ED et al. (2017) Testing the demographic effects of divergent immigrants on small populations of Trinidadian guppies. Anim Conserv 20:3–11
    Article  Google Scholar 

    Lande R, Shannon S (1996) The role of genetic variation in adaptation and population persistence in a changing environment. Evolution (NY) 50:434–437
    Article  Google Scholar 

    Liddell E, Cook CN, Sunnucks P (2020) Evaluating the use of risk assessment frameworks in the identification of population units for biodiversity conservation. Wildl Res 47:208–216
    Article  Google Scholar 

    Lippé C, Dumont P, Bernatchez L (2006) High genetic diversity and no inbreeding in the endangered copper redhorse, Moxostoma hubbsi (Catostomidae, Pisces): the positive sides of a long generation time. Mol Ecol 15:1769–1780
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    Lloyd MW, Burnett RK, Engelhardt KAM, Neel MC (2011) The structure of population genetic diversity in Vallisneria Americana in the Chesapeake Bay: implications for restoration. Conserv Genet 12:1269–1285
    Article  Google Scholar 

    Mable BK, Robertson AV, Dart S, Di Berardo C, Witham L (2005) Breakdown of self-incompatibility in the perennial Arabidopsis lyrata (Brassicaceae) and its genetic consequences. Evolution (NY) 59:1437–1448
    Article  Google Scholar 

    Markgraf V, McGlone M, Hope G (1995) Neogene paleoenvironmental and paleoclimatic change in southern temperate ecosystems—a southern perspective. Trends Ecol Evol 10:143–147
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Melville J, Goebel S, Starr C, Keogh JS, Austin JJ (2007) Conservation genetics and species status of an endangered Australian dragon, Tympanocryptis pinguicolla (Reptilia: Agamidae). Conserv Genet 8:185–195
    Article  Google Scholar 

    Mijangos JL, Pacioni C, Spencer PBS, Craig MD (2015) Contribution of genetics to ecological restoration. Mol Ecol 22:22–37
    Article  Google Scholar 

    Morgan JW (1995) Ecological studies of the endangered Rutidosis leptorrhynchoides: I. Seed production, soil seed bank dynamics, population density and their effects on recruitment. Aust J Bot 43:1–11
    Article  Google Scholar 

    Moritz C (1999) Conservation units and translocations: Strategies for conserving evolutionary processes. Hereditas 130:217–228
    Article  Google Scholar 

    Murray BG, Young AG (2001) Widespread chromosome variation in the endangered grassland forb Rutidosis leptorrhynchoides F. Muell. (Asteraceae: Gnaphalieae). Ann Bot 87:83–90
    Article  Google Scholar 

    NSW Office of Environment and Heritage (2012) National Recovery Plan for Button Wrinklewort Rutidosis leptorrhynchoides. NSW Office of Environment and Heritage, Hurstville

    Nybom H, Bartish I (2000) Effects of life history traits and sampling strategies on genetic diversity estimates obtained with RAPD markers in plants. Perspect Plant Ecol Evol Syst 3:93–114
    Article  Google Scholar 

    Pacioni C, Hunt H, Allentoft ME, Vaughan TG, Wayne AF, Baynes A et al. (2015) Genetic diversity loss in a biodiversity hotspot: ancient DNA quantifies genetic decline and former connectivity in a critically endangered marsupial. Mol Ecol 24:5813–5828
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Pavlova A, Selwood P, Harrisson KA, Murray N, Quin B, Menkhorst P et al. (2014) Integrating phylogeography and morphometrics to assess conservation merits and inform conservation strategies for an endangered subspecies of a common bird species. Biol Conserv 174:136–146
    Article  Google Scholar 

    Pickrell JK, Pritchard JK (2012) Inference of population splits and mixtures from genome-wide allele frequency data. PLoS Genet 8:1–17
    Article  CAS  Google Scholar 

    Pickup M, Field DL, Rowell DM, Young AG (2012) Predicting local adaptation in fragmented plant populations: Implications for restoration genetics. Evol Appl 5:913–924
    PubMed  PubMed Central  Article  Google Scholar 

    Pickup M, Field DL, Rowell DM, Young AG (2013) Source population characteristics affect heterosis following genetic rescue of fragmented plant populations. Proc R Soc B Biol Sci 280:20122058
    CAS  Article  Google Scholar 

    Pickup M, Young AG (2008) Population size, self-incompatibility and genetic rescue in diploid and tetraploid races of Rutidosis leptorrhynchoides (Asteraceae). Heredity (Edinb) 100:268–274
    CAS  Article  Google Scholar 

    Pimm SL, Jenkins CN, Abell R, Brooks TM, Gittleman JL, Joppa LN et al. (2015) The biodiversity of species and their rates of extinction, distribution, and protection. Science 344:1246752
    Article  CAS  Google Scholar 

    Potter S, Neaves LE, Lethbridge M, Eldridge MDB (2020) Understanding historical demographic processes to inform contemporary conservation of an arid zone specialist: the yellow-footed rock-wallaby. Genes (Basel) 11:1–24
    Article  CAS  Google Scholar 

    Powell JM (1969) The squatting occupation of Victoria 1834-60. Aust Geogr Stud 7:9–27
    Article  Google Scholar 

    Pritchard JK, Stephens M, Donnelly P (2000) Inference of population structure using multilocus genotype data. Genetics 155:945–959

    Pritchard JK, Wen W (2003) Documentation for STRUCTURE Software: Version 2.

    Raj A, Stephens M, Pritchard JK (2014) fastSTRUCTURE: variational inference of population structure in large SNP data sets. Genetics 197:573–589
    PubMed  PubMed Central  Article  Google Scholar 

    Ralls K, Ballou JD, Dudash MR, Eldridge MDB, Fenster CB, Lacy RC et al. (2018) Call for a paradigm shift in the genetic management of fragmented populations. Conserv Lett 11:1–6
    Article  Google Scholar 

    Rambaut A, Drummond AJ, Xie D, Baele G, Suchard M (2018) Posterior summarisation in Bayesian phylogenetics using Tracer 1.7. Syst Biol 67:901–904
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Rodger YS, Greenbaum G, Silver M, Bar-david S, Winters G (2018) Detecting hierarchical levels of connectivity in a population of Acacia tortilis at the northern edge of the species’ global distribution: combining classical population genetics and network analyses. PLoS ONE 13:1–16
    Article  CAS  Google Scholar 

    Rojas D, Lima AP, Momigliano P, Ivo P, Dudaniec RY, Sauer TC et al. (2020) The evolution of polymorphism in the warning coloration of the Amazonian poison frog Adelphobates galactonotus. Heredity 124:439–456

    Scarlett NH, Parsons RF (1990) Conservation biology of the southern Australian daisy Rutidosis leptorrhynchoides. In: Clark TW, Seebeck JH (eds) Management and conservation of small populations. Chicago Zoological Society, Chicago, p 195–205
    Google Scholar 

    Sinclair SJ (2010) National recovery plan for the large-fruit groundsel Senecio macrocarpus. Department of Sustainability and Environment, Melbourne

    Sjogren P, Wyoni PI (1994) Conservation genetics and detection of rare alleles in finite populations. Conserv Biol 8:267–270
    Article  Google Scholar 

    Spalink D, Mackay R, Sytsma KJ (2019) Phylogeography, population genetics and distribution modelling reveal vulnerability of Scirpus longii (Cyperaceae) and the Atlantic Coastal Plain Flora to climate change. Mol Ecol 28:2046–2061

    Team RC (2018) R: a language and environment for statistical computing

    Wagenius S, Lonsdorf E, Neuhauser C (2007) Patch aging and the S-Allee effect: breeding system effects on the demographic response of plants to habitat fragmentation. Am Nat 169:383–397
    PubMed  Article  PubMed Central  Google Scholar 

    Weaver JC (1996) Beyond the fatal shore: pastoral squatting and the occupation of Australia. Am Hist Rev 101:981–1007
    Article  Google Scholar 

    Weeks AR, Sgro CM, Young AG, Frankham R, Mitchell NJ, Miller KA et al. (2011) Assessing the benefits and risks of translocations in changing environments: A genetic perspective. Evol Appl 4:709–725
    PubMed  PubMed Central  Article  Google Scholar 

    Weeks AR, Stoklosa J, Hoffmann AA (2016) Conservation of genetic uniqueness of populations may increase extinction likelihood of endangered species: the case of Australian mammals. Front Zool 13:1–9
    Article  CAS  Google Scholar 

    Weir BS, Cockerham CC (1984) Estimating F-statistics for the analysis of population structure. Evolution (NY) 38:1358–1370
    CAS  Google Scholar 

    Wells GP, Young AG (2002) Effects of seed dispersal on spatial genetic structure in populations of Rutidosis leptorrhychoides with different levels of correlated paternity. Genet Res 79:219–226

    Whiteley AR, Fitzpatrick SW, Funk WC, Tallmon DA (2015) Genetic rescue to the rescue. Trends Ecol Evol 30:42–49
    PubMed  Article  PubMed Central  Google Scholar 

    Young AG, Brown AHD, Murray BG, Thrall PH, Miller CH (2000) Genetic erosion, restricted mating and reduced viability in fragmented populations of the endangered grassland herb Rutidosis leptorrhynchoides. In: Young AG, Clarke G (eds) Genetics, demography and viability of fragmented populations, Cambridge University Press, London, p 335–359

    Young AG, Brown AHD, Zich FC (1999) Genetic structure of fragmented populations of the endangered Daisy Rutidosis leptorrhynchoides. Cons Biol 13:256–265

    Young AG, Miller C, Gregory E, Langston A (2000) Sporophytic self-incompatibility in diploid and tetraploid races of Rutidosis leptorrhynchoides (Asteraceae). Aust J Bot 48:667–672

    Young AG, Murray BG (2000) Genetic bottlenecks and dysgenic gene flow into re-established populations of the grassland daisy, Rutidosis leptorrhynchoides. Aust J Bot 48:409–416

    Young AG, Pickup M (2010) Low S-allele numbers limit mate availability, reduce seed set and skew fitness in small populations of a self-incompatible plant. J Appl Ecol 47:541–548
    Article  Google Scholar  More

  • in

    Epigenetic responses of hare barley (Hordeum murinum subsp. leporinum) to climate change: an experimental, trait-based approach

    Alsdurf J, Anderson C, Siemens DH (2016) Epigenetics of drought-induced trans-generational plasticity: consequences for range limit development. Ann Bot 8:plv146
    Google Scholar 

    Anderson JT, Willis JH, Mitchell-Olds T (2011) Evolutionary genetics of plant adaptation. Trends Genet 27:258–66
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Aragón-Gastélum JL, Flores J, Yáñez-Espinosa L, Badano E, Ramírez-Tobías HM, Rodas-Ortíz JP et al. (2014) Induced climate change impairs photosynthetic performance in Echinocactus platyacanthus, an especially protected Mexican cactus species. Flora—Morphol Distrib Funct Ecol Plants 209:499–503
    Article  Google Scholar 

    Banerjee A, Roychoudhury A (2017) Epigenetic regulation during salinity and drought stress in plants: histone modifications and DNA methylation. Plant Gene 11:199–204
    CAS  Article  Google Scholar 

    Bartels A, Han Q, Nair P, Stacey L, Gaynier H, Mosley M et al. (2018) Dynamic DNA methylation in plant growth and development. Int J Mol Sci 19:2144
    PubMed Central  Article  CAS  Google Scholar 

    Benjamini Y, Hochberg Y (1995) Controlling the False Discovery Rate: a practical and powerful approach to multiple testing. J R Stat Soc 57:289–300
    Google Scholar 

    Bonasio R, Tu S, Reinberg D (2010) Molecular signals of epigenetic states. Science 330:612–6
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Bongers FJ, Olmo M, Lopez-Iglesias B, Anten NPR, Villar R (2017) Drought responses, phenotypic plasticity and survival of Mediterranean species in two different microclimatic sites. Plant Biol 19:386–395
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Bossdorf O, Richards CL, Pigliucci M (2008) Epigenetics for ecologists. Ecol Lett 11:106–115
    PubMed  PubMed Central  Google Scholar 

    Bossdorf O, Zhang Y (2011) A truly ecological epigenetics study. Mol Ecol 20:1572–1574
    PubMed  Article  PubMed Central  Google Scholar 

    Chapin FS, Autumn K, Pugnaire F (1993) Evolution of suites of traits in response to environmental stress. Am Nat 142:78–92
    Article  Google Scholar 

    Conrath U, Pieterse CMJ, Mauch-Mani B (2002) Priming in plant-pathogen interactions. Trends Plant Sci 7:210–6
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Dabros A, Fyles JW (2010) Effects of open-top chambers and substrate type on biogeochemical processes at disturbed boreal forest sites in northwestern Quebec. Plant Soil 327:465–479
    CAS  Article  Google Scholar 

    Delgado-Baquerizo M, Maestre FT, Rodríguez JGP, Gallardo A (2013) Biological soil crusts promote N accumulation in response to dew events in dryland soils. Soil Biol Biochem 62:22–27
    CAS  Article  Google Scholar 

    Diez CM, Meca E, Tenaillon MI, Gaut BS (2014) Three groups of transposable elements with contrasting copy number dynamics and host responses in the maize (Zea mays ssp. mays) genome. PLoS Genet 10:e1004298
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    Doyle JJ, Doyle JL (1987) A rapid DNA isolation procedure for small quantities of fresh leaf tissue. Phytochem Bull 19:11–15
    Google Scholar 

    Ewens WJ (2013) Genetic variation. In: Maloy S, Hughes K (eds) Brenner’s encyclopedia of genetics, pp 290–291

    Excoffier L, Smouse PE, Quattro JM (1992) Analysis of molecular variance inferred from metric distances among DNA haplotypes. Genetics 131:479–91
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Fernández-Pascual E, Jiménez-Alfaro B, Caujapé-Castells J, Jaén-Molina R, Díaz TE (2013) A local dormancy cline is related to the seed maturation environment, population genetic composition and climate. Ann Bot 112:937–45
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    Forestan C, Aiese Cigliano R, Farinati S, Lunardon A, Sanseverino W, Varotto S (2016) Stress-induced and epigenetic-mediated maize transcriptome regulation study by means of transcriptome reannotation and differential expression analysis. Sci Rep 6:30446
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Fotiou C, Damialis A, Krigas N, Vokou D (2007) Hordeum murinum pollen as a contributor to pollinosis: important or trivial aeroallergen? Allergy 62:180
    Google Scholar 

    Freschet GT, Violle C, Bourget MY, Scherer-Lorenzen M, Fort F (2018) Allocation, morphology, physiology, architecture: the multiple facets of plant above- and below-ground responses to resource stress. N Phytol 219:1338–52
    Article  Google Scholar 

    Garnier E, Shipley B, Roumet C, Laurent G (2001) A standardized protocol for the determination of specific leaf area and leaf dry matter content. Funct Ecol 15:688–95
    Article  Google Scholar 

    Gayacharan JA (2013) Epigenetic responses to drought stress in rice (Oryza sativa L.). Physiol Mol Biol Plants 19:379–87
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Gómez JM (2004) Importance of microhabitat and acorn burial on Quercus ilex early recruitment: Non-additive effects on multiple demographic processes. Plant Ecol 172:287–297
    Article  Google Scholar 

    Gower JC (1971) A general coefficient of similarity and some of its properties. Biometrics 27:857–71
    Article  Google Scholar 

    Grigorova B, Vaseva I, Demirevska K, Feller U (2011) Combined drought and heat stress in wheat: changes in some heat shock proteins. Biol Plant 55:105–111
    CAS  Article  Google Scholar 

    Gu Z, Eils R, Schlesner M (2016) Complex heatmaps reveal patterns and correlations in multidimensional genomic data. Bioinformatics 32:2847–9
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Herrera CM, Bazaga P (2010) Epigenetic differentiation and relationship to adaptive genetic divergence in discrete populations of the violet Viola cazorlensis. N Phytol 187:867–76
    CAS  Article  Google Scholar 

    Herrera CM, Pozo MI, Bazaga P (2012) Jack of all nectars, master of most: DNA methylation and the epigenetic basis of niche width in a flower-living yeast. Mol Ecol 21:2602–16
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Hulting AG, Haavisto JL (2013) Hare barley (Hordeum murinum ssp. leporinum) biology and management in cool season perennial grass pastures of Western Oregon. J Chem Inform Model 33:1689–99
    Google Scholar 

    Hussain F, Durrani MJ (2009) Seasonal availability, palatability and animal preferences of forage plants in Harboi arid range land, Kalat, Pakistan. Pak J Bot 41:539–554
    Google Scholar 

    Ibáñez I, Schupp EW (2001) Positive and negative interactions between environmental conditions affecting Cercocarpus ledifolius seedling survival. Oecologia 129:543–550
    PubMed  Article  PubMed Central  Google Scholar 

    Jakob SS, Meister A, Blattner FR (2004) The considerable genome size variation of Hordeum species (Poaceae) is linked to phylogeny, life form, ecology, and speciation rates. Mol Biol Evol 21:860–9
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Jaskiewicz M, Conrath U, Peterhälnsel C (2011) Chromatin modification acts as a memory for systemic acquired resistance in the plant stress response. EMBO Rep 12:50–55
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Jeremias G, Barbosa J, Marques SM, Asselman J, Gonçalves FJM, Pereira JL (2018) Synthesizing the role of epigenetics in the response and adaptation of species to climate change in freshwater ecosystems. Mol Ecol 27:2790–2806
    PubMed  Article  PubMed Central  Google Scholar 

    Jiang Y, Huang B (2000) Effects of drought or heat stress alone and in combination on Kentucky bluegrass. Crop Sci 40:1358–62
    Article  Google Scholar 

    Kaur A, Grewal A, Sharma P (2018) Comparative analysis of DNA methylation changes in two contrasting wheat genotypes under water deficit. Biol Plant 62:471–8
    CAS  Article  Google Scholar 

    Kronholm I, Bassett A, Baulcombe D, Collins S (2017) Epigenetic and genetic contributions to adaptation in Chlamydomonas. Mol Biol Evol 34:2285–2306
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    de la Riva EG, Tosto A, Pérez-Ramos IM, Navarro-Fernández CM, Olmo M, Anten NPR et al. (2016) A plant economics spectrum in Mediterranean forests along environmental gradients: is there coordination among leaf, stem and root traits? J Veg Sci 27:187–199
    Article  Google Scholar 

    Lamaoui M, Jemo M, Datla R, Bekkaoui F (2018) Heat and drought stresses in crops and approaches for their mitigation. Front Chem 6:26
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    Lampei C (2019) Multiple simultaneous treatments change plant response from adaptive parental effects to within-generation plasticity, in Arabidopsis thaliana. Oikos 128:368–379
    Article  Google Scholar 

    Latzel V, Allan E, Bortolini Silveira A, Colot V, Fischer M, Bossdorf O (2013) Epigenetic diversity increases the productivity and stability of plant populations. Nat Commun 4:2875
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    Laughlin DC, Leppert JJ, Moore MM, Sieg CH (2010) A multi-trait test of the leaf-height-seed plant strategy scheme with 133 species from a pine forest flora. Funct Ecol 24:493–501
    Article  Google Scholar 

    Li X, Zhu J, Hu F, Ge S, Ye M, Xiang H et al. (2012) Single-base resolution maps of cultivated and wild rice methylomes and regulatory roles of DNA methylation in plant gene expression. BMC Genom 2:300
    Article  CAS  Google Scholar 

    Lindner M, Maroschek M, Netherer S, Kremer A, Barbati A, Garcia-Gonzalo J et al. (2010) Climate change impacts, adaptive capacity, and vulnerability of European forest ecosystems. Ecol Manag 259:698–709
    Article  Google Scholar 

    Lira-Medeiros CF, Parisod C, Fernandes RA, Mata CS, Cardoso MA, Ferreira PCG (2010) Epigenetic variation in mangrove plants occurring in contrasting natural environment. PLoS One 5:e10326
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    Liu J, Feng L, Li J, He Z (2015) Genetic and epigenetic control of plant heat responses. Front Plant Sci 6:267
    PubMed  PubMed Central  Google Scholar 

    Liu G, Xia Y, Liu T, Dai S, Hou X (2018) The DNA methylome and association of differentially methylated regions with differential gene expression during heat stress in Brassica rapa. Int J Mol Sci 19:1414
    PubMed Central  Article  CAS  Google Scholar 

    Liu Z, Xin M, Qin J, Peng H, Ni Z, Yao Y et al. (2015) Temporal transcriptome profiling reveals expression partitioning of homoeologous genes contributing to heat and drought acclimation in wheat (Triticum aestivum L.). BMC Plant Biol 15:1
    Article  CAS  Google Scholar 

    Maestre FT, Escolar C, de Guevara ML, Quero JL, Lázaro R, Delgado-Baquerizo M et al. (2013) Changes in biocrust cover drive carbon cycle responses to climate change in drylands. Glob Chang Biol 19:3835–3847
    PubMed  PubMed Central  Article  Google Scholar 

    Marion GM, Henry GHR, Freckman DW, Johnstone J, Jones G, Jones MH et al. (1997) Open-top designs for manipulating field temperature in high-latitude ecosystems. Glob Chang Biol 3:20–32
    Article  Google Scholar 

    Mariotte P, Vandenberghe C, Kardol P, Hagedorn F, Buttler A (2013) Subordinate plant species enhance community resistance against drought in semi‐natural grasslands (S Schwinning, Ed.). J Ecol 101:763–773
    Article  Google Scholar 

    Mastan SG, Rathore MS, Bhatt VD, Yadav P, Chikara J (2012) Assessment of changes in DNA methylation by methylation-sensitive amplification polymorphism in Jatropha curcas L. subjected to salinity stress. Gene 508:125–9
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Matías L, Godoy O, Gómez-Aparicio L, Pérez-Ramos IM (2018) An experimental extreme drought reduces the likelihood of species to coexist despite increasing intransitivity in competitive networks. J Ecol 106:826–837
    Article  Google Scholar 

    Metzger DCH, Schulte PM (2017) Persistent and plastic effects of temperature on DNA methylation across the genome of threespine stickleback (Gasterosteus aculeatus). Proc R Soc B Biol Sci 284:20171667
    Article  CAS  Google Scholar 

    Mitchell P, Wardlaw T, Pinkard L (2015) Combined stresses in forests (R Mahalingam, Ed.). Springer International Publishing, Switzerland
    Google Scholar 

    Moles AT, Westoby M (2004) Seedling survival and seed size: a synthesis of the literature. J Ecol 92:372–383
    Article  Google Scholar 

    Moore LM, Lauenroth WK (2017) Differential effects of temperature and precipitation on early- vs. late-flowering species. Ecosphere 8:e01819
    Article  Google Scholar 

    Muller-Landau HC (2010) The tolerance-fecundity trade-off and the maintenance of diversity in seed size. Proc Natl Acad Sci USA 107:4242–4247
    PubMed  Article  CAS  PubMed Central  Google Scholar 

    Münzbergová Z, Latzel V, Šurinová M, Hadincová V (2019) DNA methylation as a possible mechanism affecting ability of natural populations to adapt to changing climate. Oikos 128:124–34
    Article  CAS  Google Scholar 

    Nicotra AB, Atkin OK, Bonser SP, Davidson AM, Finnegan EJ, Mathesius U et al. (2010) Plant phenotypic plasticity in a changing climate. Trends Plant Sci 15:684–92
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Ogaya R, Peñuelas J, Martínez-Vilalta J, Mangirón M (2003) Effect of drought on diameter increment of Quercus ilex, Phillyrea latifolia, and Arbutus unedo in a holm oak forest of NE Spain. Ecol Manag 180:175–184
    Article  Google Scholar 

    Olea L, San Miguel A (2006) The Spanish dehesa. A traditional Mediterranean silvopastoral system linking production and nature conservation. In: Sustainable grassland productivity: Proceedings of the 21st General Meeting of the European Grassland Federation

    Paun O, Bateman RM, Fay MF, Hedrén M, Civeyrel L, Chase MW (2010) Stable epigenetic effects impact adaptation in allopolyploid orchids (Dactylorhiza: Orchidaceae). Mol Biol Evol 27:2465–73
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Pérez-Figueroa A (2013) msap: a tool for the statistical analysis of methylation-sensitive amplified polymorphism data. Mol Ecol Resour 13:522–527
    PubMed  Article  PubMed Central  Google Scholar 

    Pérez-Ramos IM, Cambrollé J, Hidalgo-Galvez MD, Matías L, Montero-Ramírez A, Santolaya S et al. (2020) Phenological responses to climate change in communities of plants species with contrasting functional strategies. Environ Exp Bot 170:103852
    Article  CAS  Google Scholar 

    Pérez-Ramos IM, Díaz-Delgado R, de la Riva EG, Villar R, Lloret F, Marañon T (2017) Climate variability and community stability in Mediterranean shrublands: the role of functional diversity and soil environment. J Ecol 105:1335–1346
    Article  Google Scholar 

    Pérez-Ramos IM, Matías L, Gómez-Aparicio L, Godoy Ó (2019) Functional traits and phenotypic plasticity modulate species coexistence across contrasting climatic conditions. Nat Commun 10:2555
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    Piikkia K, De Temmerman L, Högy P, Pleijel H (2008) The open-top chamber impact on vapour pressure deficit and its consequences for stomatal ozone uptake. Atmos Environ 42:6513–22
    Article  CAS  Google Scholar 

    Poorter H, Niinemets Ü, Poorter L, Wright IJ, Villar R (2009) Causes and consequences of variation in leaf mass per area (LMA): A meta-analysis. N Phytol 182:565–588
    Article  Google Scholar 

    R Core Team (2013) R: a language and environment for statistical computing. 55: 275–286.

    Reyna-López GE, Simpson J, Ruiz-Herrera J (1997) Differences in DNA methylation patterns are detectable during the dimorphic transition of fungi by amplification of restriction polymorphisms. Mol Gen Genet 253:703–710
    PubMed  Article  PubMed Central  Google Scholar 

    Richards CL, Verhoeven KJF, Bossdorf O (2012) Evolutionary significance of epigenetic variation. In: Plant Genome Diversity Volume 1: Plant Genomes, their Residents, and their Evolutionary Dynamics, pp 257–274

    Rizhsky L (2004) When defense pathways collide. the response of Arabidopsis to a combination of drought and heat stress. Plant Physiol 134:1683–96
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Rodríguez-Calcerrada J, Letts MG, Rolo V, Roset S, Rambal S (2013) Multiyear impacts of partial throughfall exclusion on Buxus sempervirens in a Mediterranean forest. Syst 22:202–213
    Google Scholar 

    Seifan M, Tielbörger K, Kadmon R (2010) Direct and indirect interactions among plants explain counterintuitive positive drought effects on an eastern Mediterranean shrub species. Oikos 119:1601–9
    Article  Google Scholar 

    Sharifi-Rigi P, Saeidi H, Rahiminejad MR (2014) Genetic diversity and geographic distribution of variation of Hordeum murinum as revealed by retroelement insertional polymorphisms in Iran. Biology 69:469–77
    Google Scholar 

    Shen X, De Jonge J, Forsberg SKG, Pettersson ME, Sheng Z, Hennig L et al. (2014) Natural CMT2 variation is associated with genome-wide methylation changes and temperature seasonality. PLoS Genet 10:e1004842
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    Suzuki MM, Bird A (2008) DNA methylation landscapes: Provocative insights from epigenomics. Nat Rev Genet 9:465–76
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Tan MP (2010) Analysis of DNA methylation of maize in response to osmotic and salt stress based on methylation-sensitive amplified polymorphism. Plant Physiol Biochem 48:21–6
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Tani E, Polidoros AN, Nianiou-Obeidat I, Tsaftaris AS (2005) DNA methylation patterns are differently affected by planting density in maize inbreeds and their hybrids. Maydica 50:19–23
    Google Scholar 

    Valencia E, Méndez M, Saavedra N, Maestre FT (2016) Plant size and leaf area influence phenological and reproductive responses to warming in semiarid Mediterranean species. Perspect Plant Ecol Evol Syst 21:31–40
    PubMed  PubMed Central  Article  Google Scholar 

    Verhoeven KJF, Jansen JJ, van Dijk PJ, Biere A (2010) Stress-induced DNA methylation changes and their heritability in asexual dandelions. N Phytol 185:1108–18
    CAS  Article  Google Scholar 

    Vos P, Hogers R, Bleeker M, Reijans M, Van De Lee T, Hornes M et al. (1995) AFLP: a new technique for DNA fingerprinting. Nucleic Acids Res 23:4407–14
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    Wang WS, Pan YJ, Zhao XQ, Dwivedi D, Zhu LH, Ali J et al. (2011) rought-induced site-specific DNA methylation and its association with drought tolerance in rice (Oryza sativa) L.). J Exp Bot 62:1951–60
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Watson RGA, Baldanzi S, Pérez-Figueroa A, Gouws G, Porri F (2018) Morphological and epigenetic variation in mussels from contrasting environments. Mar Biol 165:50
    Article  CAS  Google Scholar 

    Westoby M (1998) A Leaf-Height-Seed (LHS) plant ecology strategy scheme. Plant Soil 199:213–227
    CAS  Article  Google Scholar 

    Whittington HR, Tilman D, Wragg PD, Powers JS, Browning DM (2015) Phenological responses of prairie plants vary among species and year in a three-year experimental warming study. Ecosphere 6:1–15
    Article  Google Scholar 

    Wolkovich EM, Cleland EE (2014) Phenological niches and the future of invaded ecosystems with climate change. AoB Plants 6:plu013
    PubMed  PubMed Central  Article  Google Scholar 

    Wright IJ, Reich PB, Westoby M, Ackerly DD, Baruch Z, Bongers F et al. (2004) The worldwide leaf economics spectrum. Nature 428:821–7
    CAS  Article  Google Scholar 

    Zhang Y-Y, Parepa M, Fischer M, Bossdorf O (2016) Epigenetics of colonizing species? A study of japanese knotweed in central Europe. In: Barrett SC., Colautti RI, Dlugosch KM, Rieseberg LH (eds) Invasion genetics: the baker and Stebbins legacy, John Wiley & Sons, Ltd, pp 328–340

    Zhang X, Yazaki J, Sundaresan A, Cokus S, Chan SWL, Chen H et al. (2006) Genome-wide high-resolution mapping and functional analysis of DNA methylation in arabidopsis. Cell 126:1189–201
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    Zhu JK (2016) Abiotic stress signaling and responses in plants. Cell 167:313–24
    CAS  PubMed  PubMed Central  Article  Google Scholar  More

  • in

    System design for inferring colony-level pollination activity through miniature bee-mounted sensors

    Miniature flight recorders
    Honey bees regularly carry a payload of 55–65 mg6, but a mounted flight recorder should consume only a small fraction of this allowable weight to avoid significantly affecting bee behavior. The dimensions of the recorder must also be small, as the available mounting area on the bee thorax is limited. We propose a flight recorder consisting of a (2times 2times 0.3,text {mm}^3) ASIC mounted on a (3times 3times 0.4,text {mm}^3) printed circuit board, which is similar in size to 3 mm diameter, 1.5 mm tall conventional bee tags (“Queen number set,” Betterbee) and is on par with previous studies using honey bee tagging methods30. The ASIC provides most core functionality including signal detection, memory, power harvesting, and communications circuitry, and the PCB provides a magnetic backscatter coil for near-field wireless communication. The combined weight of our chip-PCB assembly is expected to be at most 10 mg, which is a small fraction of the honey bee payload (Fig. 1c). Based on previous studies, we expect this may slightly reduce foraging trip time but not significantly impact flight characteristics, which is the focus of our system31. Furthermore, future iterations of our flight recorder will have smaller size and weight, minimizing the overall impact on honey bees. Power will be harvested from sunlight, which can provide intensity greater than 1 mW/mm(^2). On-chip photovoltaics can offer power conversion efficiency on the order of 5%, supplying 50 μW of electrical power for the chip. This power budget, while low, is sufficient for the chip, since solar angle measurement and storage of data in memory are not energy-intensive operations and need only occur a few times per second. Furthermore, wireless communication for data upload is only used when the recorder is at the base station, thus allowing the base station to fully power the near-field wireless link. Additionally, IC technology is generally robust to the environmental factors likely to be encountered by honey bees. For instance, the variations in humidity level and temperature experienced by honey bees are not expected to affect chip operation. Our proposed design is a fully power-autonomous, environmentally-robust, miniature flight recorder well-suited to the task of recording honey bee activity.
    The orientation of a bee during flight can be described by the yaw ((gamma)), which represents the absolute heading relative to the sun, and the angle-of-incidence (AOI) ((psi)), which represents the overhead angle between the sun and the sensor (Fig. 1b). To record flights, the chip uses ASPs to measure the AOI of sunlight and stores these measurements in on-chip memory. ASPs achieve AOI-sensitivity via a pair of diffraction gratings stacked over a photodiode (Fig. 1b), wherein the first grating induces a diffraction pattern that shifts laterally across the second grating as AOI is swept, thus passing a periodically-varying intensity of light to the photodiode. The stored AOI measurements can be downloaded to a base station upon return to the hive, and from these data the heading throughout the flight can be extracted. Assuming a constant speed of 6.5 m/s32, we can use the sequence of recorded headings to reconstruct the honey bee’s trajectory in post-processing.
    The data taken by the flight recorder will be subject to measurement errors, and these errors will manifest in the reconstructed trajectory. Here, we identify and explore methods to mitigate the primary sources of error. We posit that errors will stem primarily from finite heading measurement resolution, finite sampling rate, and random fluctuations in sampling rate (jitter). Each of these error sources can be suppressed through careful chip design, but improvements in these performance variables can only be made at the expense of larger chip area. For instance, the heading measurement resolution will increase if more ASPs are used to measure AOI, but each pixel consumes significant silicon area and contributes additional data that must be stored in memory. Increasing the sampling rate and decreasing sampling rate jitter requires that more measurements be stored if flight time is unchanged, thus increasing the chip area required for memory. The size of the chip is thus inversely related to the severity of the expected measurement error, and trade-offs between chip size and achievable precision should be examined. A smaller sensor is feasible if trajectory reconstruction performance requirements are relaxed; more stringent requirements will necessitate a larger chip that will increase the burden on the bee. If the relationship between final uncertainty in the reconstructed position of the bee and the core sensor specifications is understood, then chip-level performance goals can be formed based on trajectory-level precision requirements.
    To determine required heading resolution, sampling rate, and sampling rate jitter, we first describe the procedure to reconstruct trajectories. We define the timestep estimate (hat{Delta t}) as the inverse of the sampling rate, and for known flight speed v the sequence of measured heading estimates ({hat{gamma }_0, hat{gamma }_1, …, hat{gamma }_{n-1}}) can be mapped to position estimate (hat{mathbf {p}}_n = [hat{x}_n, , hat{y}_n]^T) as a function of discrete time index n via the motion model

    $$begin{aligned} hat{mathbf {p}}_n = sum _{i=0}^{n-1} v, mathbf {h}(hat{gamma }_i) hat{Delta t} end{aligned}$$
    (1)

    where h is a unit vector pointing along the heading of the bee. Each sensor estimate of heading and timestep will be subject to errors, and by modeling these errors as additive white noise, we can evaluate a confidence region for each position estimate (hat{mathbf {p}}_n) (Fig. 2a,b). Each analog heading measurement (hat{gamma }_i) must be digitized for storage on-chip and will therefore suffer from quantization error, a form of rounding. We denote this error (epsilon _{gamma ,i}) and model it as a uniform random variable with variance (sigma ^2_gamma) on the interval (pm frac{Delta gamma }{2}), where (Delta gamma) is the heading bin width. Furthermore, the timestep estimate (hat{Delta t}) will be subject to random clock jitter that can be modeled as a Gaussian random variable (epsilon _{Delta t,i}) with variance (sigma ^2_{Delta t}). In this model, we posit for simplicity that the random error in timing scales linearly in proportion to oscillator frequency, thus maintaining a fixed ratio of (sigma _{Delta t}/hat{Delta t}). These measurement errors contribute random error to position estimate ({hat{mathbf {p}}}_n), and thus each position estimate should be viewed as a random variable (mathbf {p}_n). The confidence region surrounding ({hat{mathbf {p}}}_n) depends on the covariance matrix of ({mathbf {p}_n}), and we evaluate these terms by first using the small-angle approximation to linearize the motion model with respect to (epsilon _{gamma ,i}):

    $$begin{aligned} mathbf {p}_n = sum _{i=0}^{n-1} v , mathbf {h}(hat{gamma }_i + epsilon _{gamma ,i}) (hat{Delta t}+epsilon _{Delta t, i}) &approx sum _{i=0}^{n-1} v , left( mathbf {h}(hat{gamma }_i) + mathbf {h}^perp (hat{gamma }_i)epsilon _{gamma ,i}right) (hat{Delta t}+epsilon _{Delta t, i}) = sum _{i=0}^{n-1} v , mathbf {h}(hat{gamma }_i)(hat{Delta t}+epsilon _{Delta t,i}) + sum _{i=0}^{n-1} v , mathbf {h}^perp (hat{gamma }_i)epsilon _{gamma ,i} (hat{Delta t} + epsilon _{Delta t, i}) end{aligned}$$
    (2)

    Figure 2

    (a) Sensor measurement errors produce confidence regions surrounding each estimated position shown in an example circular trajectory. (b) Zoomed-in view of confidence region at end of flight, with principle components shown. (c) The two directed standard deviations grow throughout the duration of the trajectory, and are bounded above and below by the directed standard deviations computed from the case of the straight-line trajectory. (d–f) Both directional standard deviations characterizing the final error region will depend on all three of the core sensor specs ({Delta gamma , hat{Delta t}, sigma _{Delta t}}), and the max confidence region dimension will grow if these specs are relaxed. Plots were created in MATLAB33.

    Full size image

    This approximation is valid if (epsilon _{gamma ,i}) is kept small, which can be guaranteed by keeping heading bin width (Delta gamma) small. The covariance matrix of (mathbf {p}_n) is then given by

    $$begin{aligned} mathrm {Cov}(mathbf {p}_n, mathbf {p}_n) = varvec{ Sigma }_n approx sum _{i=0}^{n-1} R(hat{{gamma }}_{i}) begin{bmatrix} v^2sigma ^2_{Delta t} &{} 0 \ 0 &{} v^2sigma ^2_{gamma }(sigma _{Delta t}^2 + (hat{Delta t})^2) end{bmatrix} {R^T({hat{gamma }}_i)} end{aligned}$$
    (3)

    where R is the standard 2 × 2 rotation matrix (Appendix: Derivation of Trajectory Precision Equation). By the Central Limit Theorem, after a sufficient number of timesteps (mathbf {p}_n) will become Gaussian distributed. Thus, the confidence region will be an ellipse with major and minor axes spanned by the eigenvectors ({mathbf {v}_1, mathbf {v}_2}) of ({varvec{Sigma }}_n). The covariances of the confidence region along each of these axes are given by the eigenvalues ({lambda _1, lambda _2}) of (varvec{Sigma }_n), and directed standard deviations can then be defined as ({sigma _1,sigma _2}). A 99% confidence region for position estimate ({mathbf {hat{p}}}_{n}) is given by an ellipse with major and minor axes lengths ({3sigma _1mathbf {v}_1, 3sigma _2mathbf {v}_2}). We therefore conclude that (3sigma _1) and (3sigma _2) are critical values defining achievable trajectory reconstruction precision.
    These (3sigma)-bounds can be computed for any measured sequence of headings, but general upper and lower bounds for (3sigma _1) and (3sigma _2) across all possible trajectories can be derived from the (3sigma _1) and (3sigma _2) given by the case in which the bee flies in a straight line. In the straight-line case, the eigenvalues of (varvec{Sigma }_n) are given by n times the diagonal entries of the diagonal matrix in Eq. (3). The directed (3sigma)-bounds are then

    $$begin{aligned} 3sigma _{1,n}^*= & {} 3sqrt{n} v sigma _{Delta t} end{aligned}$$
    (4)

    $$begin{aligned} 3sigma _{2,n}^*= & {} 3sqrt{n} v sigma _{gamma }sqrt{sigma _{Delta t}^2 + (hat{Delta t})^2} end{aligned}$$
    (5)

    For some values of ({sigma _{gamma }, hat{Delta t}, sigma _{Delta t}}), (3sigma _{1,n}^*) will be larger than (3sigma _{2,n}^*); for others, the converse will be true. These equations provide simple bounds on achievable reconstruction precision that are valid for any trajectory and can be computed from sensor characteristics. Since heading error (epsilon _{gamma _i}) is uniformly distributed, heading variance (sigma ^2_gamma) is defined by heading bin width (Delta gamma), and thus the reconstruction precision is defined by a core suite of sensor specifications: ({Delta gamma , hat{Delta t}, sigma _{Delta t}}). An illustration of the trajectory reconstruction process, along with confidence regions, is shown in Fig. 2, as well as the relationship between reconstruction precision and each of the core chip specs.
    For a maximum specified chip sensing area, trajectory precision should be optimized through balanced allocation of area to solar AOI detection and to memory (Fig. 3). We evaluate the optimal area allocation by defining the standard deviation upper bound (3sigma ^*_{max} = max (3sigma _{1,f}^*, 3sigma _{2,f}^*)), where (3sigma _{1,f}^*) and (3sigma _{2,f}^*) are the directed (3sigma)-values computed from a straight-line trajectory that is long enough to completely fill the memory. If more area is spent on pixels for AOI detection, heading resolution can be increased, thus causing (sigma _gamma) to be reduced and correspondingly lowering (3sigma ^*_{2,f}). Conversely, if more area is spent on memory, measurements can be taken more frequently, and timestep and clock jitter can be reduced, thus reducing (hat{Delta t}) and (sigma _{Delta t}). This will cause (3sigma ^*_{1,f}) to decrease, but may cause an increase in (3sigma ^*_{2,f}) since less area is now available for heading sensors. As shown in Fig. 3, the upper bound (3sigma ^*_{max}) minimizes when the two counteracting variables (3sigma ^*_{1,f}) and (3sigma ^*_{2,f}) are equal, and this intersection point prescribes the optimal allocation of sensor area. Our proposed flight recorder features a 4 (mathrm {mm}^2) chip, which can offer sensing area of approximately 3 (mathrm {mm}^2). When this area is allocated optimally, the heading resolution is (2^circ) and timestep is approximately 240 ms at timestep jitter of (3sigma _{Delta t}/Delta t = 0.03). With these specifications, the maximum recordable trajectory length is approximately 4 km, with (3sigma) uncertainty of (pm 2.4) m.
    Figure 3

    (a) Area usage is optimized where (3sigma ^*_{1,f}) and (3sigma ^*_{2,f}) cross, as this design point co-minimizes the two directional standard deviations characterizing trajectory precision. (b) This design point specifies the heading resolution and number of data words in memory that minimize trajectory uncertainty given a fixed sensor area constraint. Plots were created in MATLAB33.

    Full size image

    Sensor calibration and noise modeling
    We next examined the output from existing ASP array sensors in order to create a model for future flight recorders. These particular sensors have 96 pixels, or 24 sets of 4 oriented in 90° angles. Specifically, we designed a calibration apparatus that consists of a platform holding the ASP array driven by a custom microcontroller PCB and an arm with a light emitting diode (LED) to imitate the sun. The platform rotates to mimic a change in yaw, while the arm rotates to mimic different AOI solar light at different times of the day (Fig. 4a). Using this apparatus, we measured light input in 0.9° increments across the entire hemisphere and record the ASP array response (Fig. 4a inset) for a total resolution of 40,000 measurements in a single sweep with 200 AOI angles and 200 yaw angles. Measurements sampled by the microcontroller were transmitted to a desktop computer for logging and processing via a custom MATLAB33 script. Each ASP array response consists of a 48-bit sequence. To interpret the output, we created a lookup table from the unique 48-bit sequence that is stored for each yaw-AOI angle pair. Future data was then compared to these stored sequences in parallel using an XOR operation and the pair with the least difference in bit values was returned.
    Figure 4

    (a) Calibration apparatus with inset example of a measured ASP quadrature response. (b) ASP array repeatability over a single sensor (left) and multiple sensors (right). The magenta lines denote the expected operating region. (c) Expected system operating region (45°–75°) shown in magenta given the peak foraging hours (10 a.m.–4 p.m.) shown in grey and the AOI solar light during the Summer in Ithaca, NY. Plots in (a-inset), (b) and (c) were created in MATLAB33.

    Full size image

    We characterized sensor repeatability by repeating a sweep three times with a single ASP array and, similarly, characterized precision by comparing sweeps from two additional ASP arrays. A sample curve from the calibration sweep seen in the inset in Fig. 4a shows the characteristic angle dependence. The sensor exhibits poor response uniqueness when the sun approaches zenith and when it nears the horizon; upwards of 100(^circ) error in yaw near 90(^circ) AOI (zenith) and up to 180(^circ) error in yaw at 0°–25° AOI (horizon) (Fig. 4b). The former occurs because the position of a light source directly overhead is ambiguous to the sensor across yaw and consequently, indeterminate. We find that the sensor simply does not operate well in the latter region where light is arriving nearly parallel to the surface of the chip. Furthermore, as is expected, the difference in response is generally greater when comparing different sensors. The horizontal dark bars in the right graph in Fig. 4b are examples of this increased error. We expect our system to operate under favorable foraging conditions. Based on previously published data, we estimate this operating region to be May through September at an example location of the authors’ hometown of Ithaca, New York, USA, with the most active foraging hours being from 10 a.m. to 4 p.m.34. During this time, the AOI spans (45^circ) to (75^circ) (Fig. 4c). Within this region, we see a significantly reduced same-chip error in yaw with a mean and standard deviation of (1.52^circ pm 1.23^circ). We compensate for the remaining error within the operating region as discussed in the following sections.
    In order to realistically simulate sensor output, we create a lookup table with an error model for each individual AOI value. Similar to our theoretical model, we fit a normal distribution to error in the yaw angle measured at each AOI. We use this error model to inform reconstruction of recorded bee paths as described in the following sections. For this work, we assume that we have access to calibration data for each particular sensor, however, given the low discrepancy between sensors (Fig. 4b right), we believe that it is possible to avoid individual calibration with more sophisticated data processing. We leave this aspect for future work.
    Honey bee foraging simulation
    To properly develop our methodology for using instrumented bees to monitor the state of pollination and bloom, we designed a colony foraging simulator with an example apple orchard. Central to our approach is an understanding of the behavior and environmental conditions surrounding honey bee foraging, summarized in Fig. 1c. The following subsections detail orchard, honey bee motion, and colony foraging models.
    Orchard model
    We modelled the orchard based on common characteristics seen in real orchards (Fig. 5a) as well as those reported by the University of Vermont Cooperative Extension for Growing Fruit Trees35. Specifically, these include a tree trunk radius of 0.15 m, separation between individual trees in a given planted row as 2.4 m, and separation between rows as 5 m. To make the model realistic to a variety of orchards, we add randomness to the trunk radius (0.15–0.30 m) and to the tree locations (up to 0.5 m in any direction). We further use a 60 × 60 m2 area with 200 trees, as is representative of the common grower practice utilizing a single colony per acre36. We account for the fact that trees can be in different stages of bloom by assigning each a randomly generated quality factor between 1 and 10; this quality factor affects the number of feeding events in a flight.
    Colony foraging model
    A high-quality honey bee colony for commercial apple pollination contains a laying queen, developing brood, and 20,000–40,000 worker bees, of which approximately 25% are “foragers”, or those that leave the hive to collect pollen, nectar, resin, and water29,37,38. Since resin and water foragers are a small proportion of the forager workforce, we expect our flight dataset to be largely from bees that are visiting flowers39. For this work, we assume favorable foraging conditions as previously described and a colony size of 35,000, 25% of which are foragers for a total of 8750 foragers conducting ~ 36,000 flights per day (an average of 4 flights per forager)37,40. In our model, a forager can perform either a learning-, return-, or scout flight. Note that we exclude orientation flights, which are conducted by new foragers, under the assumption that these can be easily classified given their tortuous nature23. A learning flight is when a bee orients to a feeding site it has not previously visited after learning the bearing and distance from one of its sisters in the hive41,42. A return flight occurs when a bee orients to a feeding site it has previously visited, and can be thought of as an optimized version of the learning flight in terms of distance flown43. A scout flight occurs when a forager leaves the hive to search independently for new feeding sites. In our model, we make the assumption based on published behavioral research that 20% of foragers are acting as scout foragers and the remaining 80% perform an initial learning flight followed by return flights to the same source41,42,43.
    We incorporate that return flights will frequent the same feeding sites and that neighboring trees are likely to bloom together by randomly assigning initial goal locations (trees that bees advertise in the colony as high quality food sources) to 5 neighboring trees. Bees will randomly choose between these 5, then continue feeding on neighboring trees until they have visited trees with quality factors accumulating to at least 10 before returning to the hive. Scout foragers randomly visit trees in the orchard. Realistically, not all bees will be tagged and some tags will be lost. Here, we consider a conservative estimate that at least 430 or 5% of all foragers will be tagged, leading us to 1750 recorded flights per day, and use accumulated data to overcome the loss of tagged bees, which we expect to occur as a result of predation, senescence, stress, and other factors. Note that honey bees have a pronounced division of labor associated with worker age41, making it easy to tag a cohort and wait for them to become foragers, or to identify foragers and tag them specifically. Tagging 430 bees would take our honey bee technician approximately a day; speeding up this process is an area of future investigation.
    Honey bee motion model
    To better illustrate the characteristics of foraging flights, we recorded activity between a queenright colony with about 10,000 workers and a nearby feeder station (Fig. 5b–d). Three distinct phases of the bee flights were recorded: an initial orientation flight upon leaving the hive, flights between the hive and the feeder, and search flights near the feeder. Flights near the entrance and the feeder were characterized by rapid turning, whereas flights in between the hive and the feeder were nearly straight “bee lines”. While at the feeder, bees crawl around at a significantly reduced velocity.
    Figure 5

    (a) Photo of a honey bee in a conventional apple orchard. (b–d) Setup to showcase different types of honey bee flights. Still images from videos recorded at the hive entrance, in between, and at the feeder station, with tracked paths overlaid. (e–f) Recorded heading over the course of a simulated foraging flight and feeding event. Straight line flights are marked in grey, turns in blue, and feeding events in green. Image overlays in (b), (c) and (d) were created in MATLAB33. Plots in (e) and (f) were created in Python 3.7.

    Full size image

    We use this study to inform our foraging simulation. We assume generally straight paths in obstacle-free environments, slow turns when avoiding obstacles, and rapid turns in AOI and yaw when nearing and crawling on a food source. We further base our flight model on the following assumptions, summarized in Fig. 1c. (1) We assume the starting location is well known since the flight path will always originate and terminate at the hive entrance. (2) Based on past studies and the fact that our simulation takes place in a dense apple orchard, we assume most flights will be within the 4 km range of our flight recorder44,45,46. When leaving the hive, bees will fly an average of ~ 7.5 m/s, but once loaded with nectar, flight speed is reduced to ~ 6.5 m/s32. Here, we assume a constant velocity of ~ 6.5 m/s. (3) Based on prior honey bee tracking studies23,47, we represent flight in only two dimensions. The apple trees in the orchards we model are not tall and bees will therefore experience much greater motion in the horizontal plane than the vertical. Regardless of whether a bee in reality will fly over or under the canopy, we can model this issue in two dimensions. Turns around tree trunks represent the largest source of error for flight reconstruction, therefore by modeling flight under the canopy, we model the “worst case” scenario. (4) We estimate that the AOI of sunlight with respect to the orchard will remain within a quantifiable margin throughout the duration of the simulated flights, as bees have been found to spend an average of 20–45 min on foraging flights48. (5) We model the yaw of a bee as constant during bee-line flights, i.e. given no nearby obstacles. When the bee changes its heading to circumvent obstacles, this causes a change in yaw. (6) Once implemented on bees, we expect to be able to add sensors near the hive which would help us acquire current temperature and weather patterns as well as other dynamic factors specific to a particular environment for calibrating our model.
    To simulate scouting, learning, and return foraging flights, we combine the honey bee motion model previously described, with the Bug2 algorithm and grid-based path planning49. Grid based path planning uses a discrete grid of points over which an agent searches to find obstacle-free path segments. We compute scout and learning flights as follows. Using the Bug2 algorithm, honey bee paths are generated by first assuming direct flight along a known heading from the hive. Once an obstacle is encountered, the bee searches for a path around it, until it can once more move unhindered toward the goal. The process is repeated until all goals are reached and the bee has returned to the hive. Since no two paths are identical in nature, we plan obstacle navigation with a randomly generated grid. Return flights are found by forming a graph of all the points visited during a learning flight and using a Dijkstra’s search algorithm49 to find the shortest path through these points from the hive to the goal. Once paths are generated, we compute the sequence of headings given the 240 ms sensor sampling frequency reported earlier. An example flight is shown in Fig. 5e,f. These headings are then discretized based on the ASP calibration data discussed earlier, and noise is added given the error shown in Fig. 4c left.
    When bees land at a feeding site, they tend to crawl on and among flowers to gather nectar and pollen. We simulate this by generating random motion centered around the feeding site. The average feeding time was reported to be 1–2 min per feeding site6. To make the simulation more realistic, we randomly generate a feeding time between 60 and 120 s for foragers, and between 20 and 130 s for scouts. We furthermore assume that the AOI of sunlight changes as the honey bee tilts up and down while crawling on flowers.
    Path reconstruction and generation of foraging activity maps
    Path reconstruction inherently depends on the accuracy with which our sensor is able to describe the motion of an instrumented bee. Beyond limited angle resolution, errors related to the sampling rate accumulate when turns occur, at worst (v_{loaded} Delta t) = 1.6 m. To increase the accuracy of our foraging activity maps, we use models of sensor noise and flight speed, and leverage all recorded flights. The full workflow is shown in Fig. 6, where the foraging simulation portion generates the data we expect if our sensors are placed on actual bees, and the remaining flow is the data processing portion of our methodology.
    Figure 6

    Flow chart combining the colony foraging simulation, simulated flight recorder, path reconstruction, and data processing to output the final foraging activity map. Recorded heading plot made in Python 3.7. Other plots were created in MATLAB33.

    Full size image

    We first identify feeding and turn features in our path data that stand out above the noise floor. Feeding features consist of crawling behavior in which the bee is moving at much lower speed compared to flying, but with rapidly varying yaw and AOI, inducing an elevated rate of change in measured AOI. A turn is marked by a significant change in yaw. Given the time of day and location, we can find the expected AOI on the orchard and use this to find the yaw from our lookup table. Our algorithm then indexes the sensor noise lookup table to find the related mean and standard deviation and uses this to estimate turns and feeding status. Our method marks a turn for a change greater than three standard deviations in yaw ((6.06^circ)). A feeding site is marked if the detected AOI deviates more than three standard deviations ((3.72^circ)) from the one expected, or if two consecutive AOI samples deviate more than 3 standard deviations, adjustable depending on the total flight time. The average detection accuracy of a turn is 99% with a standard deviation of 0.28% and average detection accuracy of a feeding site is 99% with a standard deviation of 0.29%.
    We explore three methods to generate activity maps: from the raw data, we classify feeding features using the aforementioned statistical approach and produce maps based on accumulated path reconstructions; in “iteration 1” and “iteration 4” we take a particle filter approach to improve path localization based on knowledge of the hive and frequented feeding sites respectively. Particle filters are used to track a variable of interest over time by creating many representative particles, generating predictions according to dynamics and error models, and then updating them according to observation models50. In this case, each particle forms a candidate trajectory and predictions are based on speed, sensor readings, and the sensor error model. Assuming that we start without knowledge of the orchard, we build up an observation model by reconstructing and accumulating the raw flight paths (Fig. 6, raw data). To account for the fact that bees may turn at any point between sensor readings, we then upsample our sensor readings by a factor of 8, essentially producing 8 guesses for where the bee actually turned. Based on the artificially upsampled data and a random sample from the sensor noise distribution at a given AOI, we then generate the displacement in each path segment as follows:

    $$begin{aligned} begin{bmatrix} Delta x_{t} \ Delta y_{t} end{bmatrix} = v Delta t begin{bmatrix} cos {(gamma _{t} + gamma _{noise})} \ sin {(gamma _{t} + gamma _{noise})} end{bmatrix} end{aligned}$$
    (6)

    The final path is found as a cumulative sum of these displacements. We repeat this process to generate 5000 particles for each flight. The choice of 5000 is guided by our variable dimensions; the upsampling rate of 8 was the highest we could handle on a quadcore desktop computer with 16 GB RAM—to truly represent all potential turns we would need a number of particles equal to 8 to the power of the number of turns per flight. For reference, the average number of turns per flight is 13, thus the true representation in our sampling approach would require (8^{13}) particles.
    In iteration 1, we choose ten of these particles according to proximity to the hive upon return, and use the feeding features from these to form an initial foraging activity map, represented by a discrete grid with computed visit numbers. Note that in a typical localization approach a single particle, or average of several particles, is chosen as the final reconstruction. In our approach, we retain 10 different reconstructions for each individual path in order to better represent the distribution of points in a path due to sensor noise.
    After this first pass, we repeat the process, but now filter particles by using the initial activity map as an observation model for the particle filter (Fig. 6, iteration 2–4). Specifically, we update particle probability at each detected feeding site by assigning the probability of the nearest grid cell in the map to the particle, and then sampling the particles by weight. At the end of the process, the ten particles with the highest probability product are used to construct an updated activity map. More

  • in

    Synthesis of novel phytol-derived γ-butyrolactones and evaluation of their biological activity

    Chemistry
    General
    Racemic mixture of cis/trans (35%:65%) isomers of phytol (1) (PYT) (97% purity), N-bromosuccinimide (NBS, 99% purity) and N-chlorosuccinimide (NCS, 98% purity) were purchased from Sigma-Aldrich Chemical Co. (St. Louis, MO, USA), while trimethylortoacetate was purchased from Fluka. Analytical grade acetic acid, sodium hydrogen carbonate, acetone, hexane, diethyl ether, tetrahydrofuran (THF), anhydrous magnesium sulfate, sodium chloride were purchased from Chempur (Poland).
    Analytical Thin Layer Chromatography (TLC) was carried out on silica gel coated aluminium plates (DC-Alufolien Kieselgel 60 F254, Merck, Darmstadt, Germany) with a mixture of hexane, acetone and diethyl ether in various ratios as the developing systems. Compounds were visualized by spraying the plates with solution of 1% Ce(SO4)2 and 2% H3[P(Mo3O10)4] (2 g) in 10% H2SO4, followed by heating to 120–200 °C.
    The products of chemical synthesis were purified by column chromatography on silica gel (Kieselgel 60, 230–400 mesh ASTM, 40–63 μm, Merck) using a mixture of hexane, acetone, and diethyl ether (in various ratios) as eluents.
    Gas chromatography (GC) analysis was carried out on an Agilent Technologies 6890 N Network GC instrument (Santa Clara, CA, USA) equipped with autosampler, split injection (20:1) and FID detector using a DB-5HT column (Agilent, Santa Clara, USA) (polyimide-coated fused silica tubing, 30 m × 0.25 mm × 0.1 µm) with hydrogen as the carrier gas. Products of the chemical reactions were analysed using the following temperature programme: injector 250 °C, detector (FID) 250 °C, initial column temperature: 100 °C, 100–300 °C (rate 30 °C/min), final column temperature 300 °C (hold 2 min).
    Nuclear magnetic resonance spectra 1H NMR, 13C NMR, DEPT 135, HSQC, 1H–1H COSY and NOESY were recorded in CDCl3 solutions with signals of residual solvent (δH = 7.26 δC = 77) on a Brüker Avance II 600 MHz (Brüker, Rheinstetten, Germany) spectrometer.
    High-resolution mass spectra (HRMS) were recorded using electron spray ionization (ESI) technique on spectrometer Waters ESI-Q-TOF Premier XE (Waters Corp., Milford, MA, USA).
    General procedure for the synthesis of compounds (2–7)
    The preparation of ester 2 and acid 3 has been illustrated in detail in our previous work39, and so the synthesis method would not be listed here.
    To a solution of acid 3 (7.8 mmol) in THF (30 mL) the N-bromosuccinimide (7.8 mmol) or N-chlorosuccinimide (7.8 mmol) was added. The mixture was stirred at room temperature for 48–96 h. When the substrate reacted completely (TLC, GC) the mixture was diluted with diethyl ether and washed with saturated NaHCO3 solution and brine. Organic layer of ether extract was separated and dired over anhydrous magnesium sulfate and evaporated on a rotary evaporator. New δ-halogeno-γ-lactones (4–7) were separated by silica gel column using for elution hexane/diethyl eter in gradient system. Bromo- and chlorolactonization afforded products with the following physical and spectral data presented below:
    trans-5-Bromomethyl-4-methyl-4-(4′,8′,12′-trimethyltridecyl)dihydrofuran-2-one ( 4 )
    (25% yield); 1H NMR (600 MHz, CDCl3): δ 0.85 (four t, J = 6.4 Hz, 12H, CH3-4′, CH3-8′, (CH3)2–12′), 1.05–1.55 (m, 21H, CH2-1′, CH2-2′, CH2-3′, CH2-5′, CH2-6′, CH2-7′, CH2-9′, CH2-10′, CH2-11′, H-4′, H-8′, H-12′), 1.24 (s, 3H, CH3-4), 2.30 and 2.60 (two d, J = 17.2 Hz, 2H, CH2-3), 3.47 (dd, J = 11.3, 7.3 Hz, 1H, one of CH2-Br), 3.55 (dd, J = 11.3, 4.5 Hz, 1H, one of CH2-Br), 4.39 (dd, J = 7.2, 4.5 Hz, 1H, H-5); 13C NMR (150 MHz, CDCl3): δ 19.61, 19.69 (CH3)2–12′), 22.65, 22.75 (CH3-4′, CH3-8′), 24.50 (CH3-4), 29.21 (CH2-Br), 41.49 (CH2-3), 42.86 (C-4), 22.00, 24.47, 24.82, 34.08, 37.28, 37.41, 37.61, 37.70, 39.38 (CH2-1′, CH2-2′, CH2-3′, CH2-5′, CH2-6′, CH2-7′, CH2-9′, CH2-10′, CH2-11′), 28.00, 32.69, 32.80 (H-4′, H-8′, H-12′), 87.66 (H-5), 174.92 (C-2); HRMS (ESI): m/z calcd. for C22H41BrO2 [M + Na]+ 439.2188; found 439.2182.
    cis-5-Bromomethyl-4-methyl-4-(4′,8′,12′-trimethyltridecyl)dihydrofuran-2-one ( 5 )
    (46% yield); 1H NMR (600 MHz, CDCl3): δ 0.85 (four t, J = 6.4 Hz, 12H, CH3-4′, CH3-8′, (CH3)2–12′), 1.04–1.55 (m, 21H, CH2-1′, CH2-2′, CH2-3′, CH2-5′, CH2-6′, CH2-7′, CH2-9′, CH2-10′, CH2-11′, H-4′, H-8′, H-12′), 1.08 (s, 3H, CH3-4), 2.38 and 2.49 (two d, J = 17.2 Hz, 2H, CH2-3), 3.48 (m, 2H, one CH2-Br), 4.41 (dd, J = 7.5, 4.5 Hz, 1H, H-5); 13C NMR (150 MHz, CDCl3): δ 18.96 (CH3-4), 19.69, 19.76 (CH3)2–12′), 22.65, 22.75 (CH3-4′, CH3-8′), 29.17 (CH2-Br), 42.70 (CH2-3), 43.09 (C-4), 22.20, 24.46, 24.82, 37.28, 37.39, 37.41, 37.49, 39.38, 39.96 (CH2-1′, CH2-2′, CH2-3′, CH2-5′, CH2-6′, CH2-7′, CH2-9′, CH2-10′, CH2-11′), 28.00, 30.95, 32.72 (H-4′, H-8′, H-12′), 86.46 (H-5), 174.76 (C-2); HRMS (ESI): m/z calcd. for C22H41BrO2 [M + Na]+ 439.2188; found 439.2183.
    trans-5-Chloromethyl-4-methyl-4-(4′,8′,12′-trimethyltridecyl)dihydrofuran-2-one ( 6 )
    (21% yield); 1H NMR (600 MHz, CDCl3): δ 0.84 (four t, J = 6.4 Hz, 12H, CH3-4′, CH3-8′, (CH3)2–12′), 1.03–1.54 (m, 21H, CH2-1′, CH2-2′, CH2-3′, CH2-5′, CH2-6′, CH2-7′, CH2-9′, CH2-10′, CH2-11′, H-4′, H-8′, H-12′), 1.23 (s, 3H, CH3-4), 2.23 and 2.60 (two d, J = 17.2 Hz, 2H, CH2-3), 3.67 (dd, J = 12.1, 6.1 Hz, 1H, one of CH2-Cl), 3.73 (dd, J = 12.1, 4.6 Hz, 1H, one of CH2-Cl), 4.33 (dd, J = 7.2, 4.6 Hz, 1H, H-5); 13C NMR (150 MHz, CDCl3): δ 19.61, 19.67 (CH3)2–12′), 22.65, 22.75 (CH3-4′, CH3-8′), 24.67 (CH3-4), 42.30 (CH2-Cl), 41.48 (CH2-3), 42.38 (C-4), 22.09, 24.46, 24.82, 34.21, 37.29, 37.39, 37.61, 37.70, 39.38 (CH2-1′, CH2-2′, CH2-3′, CH2-5′, CH2-6′, CH2-7′, CH2-9′, CH2-10′, CH2-11′), 28.00, 32.70, 32.80 (H-4′, H-8′, H-12′), 87.45 (H-5), 175.21 (C-2); HRMS (ESI): m/z calcd. for C22H41ClO2 [M + Na]+ 395.2693; found 395.2698.
    cis-5-chloromethyl-4-methyl-4-(4′,8′,12′-trimethyltridecyl)dihydrofuran-2-one ( 7 )
    (39% yield); 1H NMR (600 MHz, CDCl3): δ 0.85 (four t, J = 6.6 Hz, 12H, CH3-4′, CH3-8′, (CH3)2–12′), 1.04–1.56 (m, 21H, CH2-1′, CH2-2′, CH2-3′, CH2-5′, CH2-6′, CH2-7′, CH2-9′, CH2-10′, CH2-11′, H-4′, H-8′, H-12′), 1.06 (s, 3H, CH3-4), 2.38 and 2.47 (two d, J = 17.2 Hz, 2H, CH2-3), 3.67 (m, 2H, one CH2-Cl), 4.35 (dd, J = 6.4, 4.9 Hz, 1H, H-5); 13C NMR (150 MHz, CDCl3): δ 19.06 (CH3-4), 19.69, 19.76 (CH3)2–12′), 22.65, 22.74 (CH3-4′, CH3-8′), 42.52 (CH2-Cl), 42.39 (CH2-3), 42.54 (C-4), 22.14, 24.46, 24.83, 37.28, 37.37, 37.41, 37.49, 39.38, 40.16 (CH2-1′, CH2-2′, CH2-3′, CH2-5′, CH2-6′, CH2-7′, CH2-9′, CH2-10′, CH2-11′), 28.00, 32.64, 32.80 (H-4′, H-8′, H-12′), 86.24 (H-5), 175.04 (C-2); HRMS (ESI): m/z calcd. for C22H41ClO2 [M + Na]+ 395.2693; found 395.2697.
    Deterrent activity of phytol and its derivatives
    Aphids, plants and compound application
    The peach potato aphids Myzus persicae (Sulzer) and the Chinese cabbage Brassica rapa subsp. pekinensis (Lour.) Hanelt were reared in laboratory at 20 °C, 65% r.h., and L16:8D photoperiod. One to seven days old apterous females of M. persicae and 3-week old plants with 4–5 fully developed leaves were used for experiments. M. persicae were obtained from the laboratory culture maintained at the Department of Botany and Ecology for many generations since 2000. All experiments were carried out under the same conditions of temperature, relative humidity, and photoperiod. The bioassays were started at 10–11.a.m. Each compound was dissolved in 70% ethanol to obtain the recommended 0.1% solution40. All compounds were applied on the adaxial and abaxial leaf surfaces by immersing a leaf in the ethanolic solution of a given compound for 30 s.20. Control leaves of similar size were immersed in 70% ethanol that was used as a solvent for the studied compounds. Experiments were performed 1 h after the compounds application to allow the evaporation of the solvent. Every plant and aphid were used only once.
    Aphid settling (choice test)
    This bioassay allows the study of aphid host preferences under semi-natural conditions41. In the present study, aphids were given free choice between control and treated excised leaves that were placed in a Petri dish. Aphids were placed in the dish equidistance from treated and untreated leaves, so that aphids could choose between treated (on one half of a Petri dish) and control leaves (on the other half of the dish). Aphids that settled, i.e. they did not move, and the position of their antennae indicated feeding, on each leaf were counted at 1 h, 2 h, and 24 h intervals after access to the leaf. Each experiment was replicated 8 times (n = 8 replicates, 20 viviparous apterous females/replicate). Aphids that were moving or not on any of the leaves were not counted.
    Behavioral responses of aphids Myzus persicae during probing and feeding (no-choice test)
    Aphid probing and the phloem sap uptake by M. persicae was monitored using the technique of electronic registration of aphid probing in plant tissues, known as EPG (= Electrical Penetration Graph), that is frequently employed in insect–plant relationship studies considering insects with sucking-piercing mouthparts42,43,44. In this experimental set-up, aphid and plant are connected to electrodes and thus made parts of an electric circuit, which is completed when the aphid inserts its stylets into the plant. Weak voltage is supplied in the circuit, and all changing electric properties are recorded as EPG waveforms that can be correlated with aphid activities and stylet position in plant tissues45,46. The parameters describing aphid behaviour during probing and feeding, such as total time of probing, proportion of phloem patterns E1 and E2, number of probes, etc., are good indicators of plant suitability or interference of probing by chemical or physical factors in individual plant tissues44,45,46. In the present study, aphids were attached to a golden wire electrode with conductive silver paint (epgsystems. eu) and starved for 1 h prior to the experiment. Probing behaviour of 12 apterous females/studied compound and control was monitored for 8 h continuously with Giga-4 and Giga-8 DC EPG with 1 GΩ of input resistance recording equipment (EPG Systems, Wageningen, The Netherlands). Each aphid was given access to a freshly prepared plant and each aphid/plant combination was used only once. Various behavioural phases were labelled manually using the Stylet + software (www.epgsystems.eu). The following aphid behaviours were distinguished: no penetration (waveform ‘np’ – aphid stylets outside the plant), pathway phase—penetration of non-phloem tissues (waveforms ‘ABC’), phloem phase (salivation into sieve elements, waveform ‘E1’ and ingestion of phloem sap, waveform ‘E2’), and xylem phase (ingestion of xylem sap, waveform ‘G’). Waveform ‘G’ occurred rarely irrespective of the treatment. Therefore, in all calculations the xylem phase was added to the pathway phase and termed as probing in non-phloem tissues. The E1/E2 transition patterns were included in E2. Waveform patterns that were not terminated before the end of the experimental period (8 h) were not excluded from the calculations. The parameters derived from EPG recordings were analyzed according to their frequency and duration in configuration related to activities in peripheral and vascular tissues.
    Statistical analysis
    The data of the choice-test were analyzed using Student’s t-test (STATISTICA 13.1. package). If aphids showed clear preference for the leaf treated with the tested compound (p  More

  • in

    Factors influencing scavenger guilds and scavenging efficiency in Southwestern Montana

    1.
    Leroux, S. J. & Loreau, M. Subsidy hypothesis and strength of trophic cascades across ecosystems. Ecol. Lett. 11, 1147–1156 (2008).
    PubMed  Article  PubMed Central  Google Scholar 
    2.
    Moore, J. C. et al. Detritus, trophic dynamics and biodiversity. Ecol. Lett. 7, 584–600 (2004).
    ADS  Article  Google Scholar 

    3.
    Nowlin, W. H., Vanni, M. J. & Yang, L. H. Comparing resource pulses in aquatic and terrestrial ecosystems. Ecology 89, 647–659 (2008).
    PubMed  Article  PubMed Central  Google Scholar 

    4.
    Wilson, E. E. & Wolkovich, E. M. Scavenging: how carnivores and carrion structure communities. Trends Ecol. Evol. 26, 129–135 (2011).
    PubMed  Article  PubMed Central  Google Scholar 

    5.
    Margalida, A., Donázar, J. A., Carrete, M. & Sánchez-Zapata, J. A. Sanitary versus environmental policies: fitting together two pieces of the puzzle of European vulture conservation. J. Appl. Ecol. 47, 931–935 (2010).
    Article  Google Scholar 

    6.
    Margalida, A., Colomer, M. À. & Oro, D. Man-induced activities modify demographic parameters in a long-lived species: effects of poisoning and health policies. Ecol. Appl. 24, 436–444 (2014).
    PubMed  Article  PubMed Central  Google Scholar 

    7.
    Moreno-Opo, R. & Margalida, A. Carcasses provide resources not exclusively to scavengers: patterns of carrion exploitation by passerine birds. Ecosphere 4, art105 (2013).

    8.
    DeVault, T. L., Rhodes, O. E. Jr. & Shivik, J. A. Scavenging by vertebrates: behavioral, ecological, and evolutionary perspectives on an important energy transfer pathway in terrestrial ecosystems. Oikos 102, 225–234 (2003).
    Article  Google Scholar 

    9.
    Barton, P. S., Cunningham, S. A., Lindenmayer, D. B. & Manning, A. D. The role of carrion in maintaining biodiversity and ecological processes in terrestrial ecosystems. Oecologia 171, 761–772 (2013).
    ADS  PubMed  Article  PubMed Central  Google Scholar 

    10.
    Bump, J. K. et al. Ungulate carcasses perforate ecological filters and create biogeochemical hotspots in forest herbaceous layers allowing trees a competitive advantage. Ecosystems 12, 996–1007 (2009).
    Article  Google Scholar 

    11.
    Danell, K., Berteaux, D. & Bråthen, K. A. Effect of muskox carcasses on nitrogen concentration in tundra vegetation. Arctic 55, 389–392 (2002).
    Article  Google Scholar 

    12.
    Klink, R., Laar-Wiersma, J., Vorst, O. & Smit, C. Rewilding with large herbivores: positive direct and delayed effects of carrion on plant and arthropod communities. PLoS ONE 15, e0226946 (2020).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    13.
    Turner, W. C. et al. Fatal attraction: vegetation responses to nutrient inputs attract herbivores to infectious anthrax carcass sites. Proc. R. Soc. Lond. B Biol. Sci. 281, e20141785 (2014).

    14.
    Mateo-Tomás, P. et al. From regional to global patterns in vertebrate scavenger communities subsidized by big game hunting. Divers. Distrib. 21, 913–924 (2015).
    Article  Google Scholar 

    15.
    Markandya, A. et al. Counting the cost of vulture decline—an appraisal of the human health and other benefits of vultures in India. Ecol. Econ. 67, 194–204 (2008).
    Article  Google Scholar 

    16.
    Selva, N., Jędrzejewska, B., Jędrzejewski, W. & Wajrak, A. Factors affecting carcass use by a guild of scavengers in European temperate woodland. Can. J. Zool. 83, 1590–1601 (2005).
    Article  Google Scholar 

    17.
    DeVault, T. L., Brisbin, J., Lehr, I., Rhodes, J. & Olin, E. Factors influencing the acquisition of rodent carrion by vertebrate scavengers and decomposers. Can. J. Zool. 82, 502–509 (2004).
    Article  Google Scholar 

    18.
    Arrondo, E. et al. Rewilding traditional grazing areas affects scavenger assemblages and carcass consumption patterns. Basic Appl. Ecol. 41, 56–66 (2019).
    Article  Google Scholar 

    19.
    Morales-Reyes, Z. et al. Scavenging efficiency and red fox abundance in Mediterranean mountains with and without vultures. Acta Oecologica 79, 81–88 (2017).
    ADS  Article  Google Scholar 

    20.
    Ruzicka, R. E. & Conover, M. R. Does weather or site characteristics influence the ability of scavengers to locate food? Ethology 118, 187–196 (2012).
    Article  Google Scholar 

    21.
    Moleón, M., Sánchez-Zapata, J., Sebastián-González, E. & Owen-Smith, N. Carcass size shapes the structure and functioning of an African scavenging assemblage. Oikos 124, 1391–1403 (2015).

    22.
    Cornwell, W. K. et al. Plant species traits are the predominant control on litter decomposition rates within biomes worldwide. Ecol. Lett. 11, 1065–1071 (2008).
    PubMed  Article  PubMed Central  Google Scholar 

    23.
    Ogada, D. L., Torchin, M. E., Kinnaird, M. F. & Ezenwa, V. O. Effects of vulture declines on facultative scavengers and potential implications for mammalian disease transmission. Conserv. Biol. 26, 453–460 (2012).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    24.
    Sekercioglu, Ç. H., Wenny, D. G. & Whelan, C. J. Why Birds Matter: Avian Ecological Function and Ecosystem Services (University of Chicago Press, 2016).

    25.
    Pereira, L. M., Owen-Smith, N. & Moleón, M. Facultative predation and scavenging by mammalian carnivores: seasonal, regional and intra-guild comparisons. Mammal Rev. 44, 44–55 (2014).
    Article  Google Scholar 

    26.
    Selva, N. & Fortuna, M. A. The nested structure of a scavenger community. Proc. R. Soc. B Biol. Sci. 274, 1101–1108 (2007).
    Article  Google Scholar 

    27.
    Wolf, C. & Ripple, W. J. Range contractions of the world’s large carnivores. R. Soc. Open Sci. 4, 170052 (2017).
    ADS  PubMed  PubMed Central  Article  Google Scholar 

    28.
    Grimm, N. B. et al. The impacts of climate change on ecosystem structure and function. Front. Ecol. Environ. 11, 474–482 (2013).
    Article  Google Scholar 

    29.
    Lauenroth, W. et al. Potential effects of climate change on the temperate zones of North and South America. Rev. Chil. Hist. Nat. 77, 439–453 (2004).
    Article  Google Scholar 

    30.
    Shanley, C. S. et al. Climate change implications in the northern coastal temperate rainforest of North America. Clim. Change 130, 155–170 (2015).
    ADS  CAS  Article  Google Scholar 

    31.
    Wilmers, C. C. & Getz, W. M. Gray wolves as climate change buffers in yellowstone. PLOS Biol. 3, e92 (2005).
    PubMed  PubMed Central  Article  CAS  Google Scholar 

    32.
    Sebastián-González, E. et al. Network structure of vertebrate scavenger assemblages at the global scale: drivers and ecosystem functioning implications. Ecography 43, 1143–1155 (2020).
    Article  Google Scholar 

    33.
    Pardo-Barquín, E., Mateo-Tomás, P. & Olea, P. P. Habitat characteristics from local to landscape scales combine to shape vertebrate scavenging communities. Basic Appl. Ecol. 34, 126–139 (2019).
    Article  Google Scholar 

    34.
    Sebastián-González, E. et al. Scavenging in the Anthropocene: human impact drives vertebrate scavenger species richness at a global scale. Glob. Change Biol. 25, 3005–3017 (2019).
    ADS  Article  Google Scholar 

    35.
    Turner, K. L., Abernethy, E. F., Conner, L. M., Rhodes, O. E. & Beasley, J. C. Abiotic and biotic factors modulate carrion fate and vertebrate scavenging communities. Ecology 98, 2413–2424 (2017).
    PubMed  Article  PubMed Central  Google Scholar 

    36.
    Janßen, F., Treude, T. & Witte, U. Scavenger assemblages under differing trophic conditions: a case study in the deep Arabian Sea. Deep Sea Res. Part II Top. Stud. Oceanogr. 47, 2999–3026 (2000).
    ADS  Article  Google Scholar 

    37.
    Houston, D. C. To the vultures belong the spoils. Nat. Hist. 103, 34–41 (1994).
    Google Scholar 

    38.
    Houston, D. C. Scavenging efficiency of turkey vultures in tropical forest. The Condor 88, 318–323 (1986).
    Article  Google Scholar 

    39.
    Sauer, J. et al. The North American breeding bird survey, results and analysis 1966–2015. (2017).

    40.
    Hill, J. E., DeVault, T. L., Beasley, J. C., Rhodes, O. E. & Belant, J. L. Effects of vulture exclusion on carrion consumption by facultative scavengers. Ecol. Evol. 8, 2518–2526 (2018).
    PubMed  PubMed Central  Article  Google Scholar 

    41.
    Heinrich, B. Winter foraging at carcasses by three sympatric corvids, with emphasis on recruitment by the raven, Corvus corax. Behav. Ecol. Sociobiol. 23, 141–156 (1988).
    Article  Google Scholar 

    42.
    Bellan, S. E., Turnbull, P. C. B., Beyer, W. & Getz, W. M. Effects of experimental exclusion of scavengers from carcasses of anthrax-infected herbivores on bacillus anthracis sporulation, survival, and distribution. Appl. Environ. Microbiol. 79, 3756–3761 (2013).
    CAS  PubMed  PubMed Central  Article  Google Scholar 

    43.
    The IUCN Red List of Threatened Species. IUCN Red List of Threatened Species https://www.iucnredlist.org/en.

    44.
    Kiff, L. F. The current status of North American vultures. In Raptors at Risk 175–189 (World Working Group on Birds of Prey/Hancock House, 2000).

    45.
    Prasad, A. M., Iverson, L. R., Peters, M. P. & Matthews, S. N. Climate change tree atlas (Northern Research Station, US Forest Service, Delaware, OH, 2014).
    Google Scholar 

    46.
    Kiff, L. The current status of North American vultures. in 175–189 (2000).

    47.
    Houston, D. C. Competition for food between Neotropical vultures in forest. Ibis 130, 402–417 (1988).
    Article  Google Scholar 

    48.
    Gomez, L. G., Houston, D. C., Cotton, P. & Tye, A. The role of greater yellow-headed vultures Cathartes melambrotus as scavengers in neotropical forest. Ibis 136, 193–196 (1994).
    Article  Google Scholar 

    49.
    Ripple, W. J. et al. Status and ecological effects of the world’s largest carnivores. Science 343, e1241484 (2014).

    50.
    Tomberlin, J. K., Barton, B. T., Lashley, M. A. & Jordan, H. R. Mass mortality events and the role of necrophagous invertebrates. Curr. Opin. Insect Sci. 23, 7–12 (2017).
    PubMed  Article  PubMed Central  Google Scholar 

    51.
    Fey, S. B. et al. Recent shifts in the occurrence, cause, and magnitude of animal mass mortality events. Proc. Natl. Acad. Sci. 112, 1083–1088 (2015).
    ADS  CAS  PubMed  Article  PubMed Central  Google Scholar 

    52.
    Wikenros, C., Sand, H., Ahlqvist, P. & Liberg, O. Biomass flow and scavengers use of carcasses after re-colonization of an apex predator. PLoS ONE 8, e77373 (2013).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    53.
    Kočárek, P. Decomposition and Coleoptera succession on exposed carrion of small mammal in Opava, the Czech Republic. Eur. J. Soil Biol. 39, 31–45 (2003).
    Article  Google Scholar 

    54.
    Matuszewski, S., Bajerlein, D., Konwerski, S. & Szpila, K. Insect succession and carrion decomposition in selected forests of Central Europe. Part 1: pattern and rate of decomposition. Forensic Sci. Int. 194, 85–93 (2010).
    PubMed  Article  PubMed Central  Google Scholar 

    55.
    Reed, H. B. A study of dog carcass communities in tennessee, with special reference to the insects. Am. Midl. Nat. 59, 213–245 (1958).
    Article  Google Scholar 

    56.
    Bauer, J. W., Logan, K. A., Sweanor, L. L. & Boyce, W. M. Scavenging behavior in Puma. Southwest. Nat. 50, 466–471 (2005).
    Article  Google Scholar 

    57.
    Burkepile, D. E. et al. Chemically mediated competition between microbes and animals: microbes as consumers in food webs. Ecology 87, 2821–2831 (2006).
    PubMed  Article  PubMed Central  Google Scholar 

    58.
    Janzen, D. H. Why fruits rot, seeds mold, and meat spoils. Am. Nat. 111, 691–713 (1977).
    CAS  Article  Google Scholar 

    59.
    DeVault, T. L. & Rhodes, O. E. Identification of vertebrate scavengers of small mammal carcasses in a forested landscape. Acta Theriol. (Warsz.) 47, 185–192 (2002).
    Article  Google Scholar 

    60.
    Parker, K. L., Robbins, C. T. & Hanley, T. A. Energy expenditures for locomotion by Mule Deer and Elk. J. Wildl. Manag. 48, 474–488 (1984).
    Article  Google Scholar 

    61.
    Crête, M. & Larivière, S. Estimating the costs of locomotion in snow for coyotes. Can. J. Zool. 81, 1808–1814 (2003).
    Article  Google Scholar 

    62.
    Droghini, A. & Boutin, S. The calm during the storm: snowfall events decrease the movement rates of grey wolves (Canis lupus). PLoS ONE 13, e0205742 (2018).

    63.
    Green, G. I., Mattson, D. J. & Peek, J. M. Spring feeding on ungulate carcasses by grizzly bears in Yellowstone National Park. J. Wildl. Manag. 61, 1040–1055 (1997).
    Article  Google Scholar 

    64.
    De Jong, G. D. & Chadwick, J. W. Decomposition and arthropod succession on exposed rabbit carrion during summer at high altitudes in colorado, USA. J. Med. Entomol. 36, 833–845 (1999).
    PubMed  Article  PubMed Central  Google Scholar 

    65.
    Sun, S.-J. et al. Climate-mediated cooperation promotes niche expansion in burying beetles. Elife 3, e02440 (2014).
    PubMed  PubMed Central  Article  Google Scholar 

    66.
    Krofel, M. Monitoring of facultative avian scavengers on large mammal carcasses in Dinaric forest of Slovenia. Acrocephalus 32, 45–51 (2011).
    Article  Google Scholar 

    67.
    DeVault, T. L., Seamans, T. W., Linnell, K. E., Sparks, D. W. & Beasley, J. C. Scavenger removal of bird carcasses at simulated wind turbines: Does carcass type matter?. Ecosphere 8, e01994 (2017).
    Article  Google Scholar 

    68.
    Turner, K. L., Conner, L. M. & Beasley, J. C. Effect of mammalian mesopredator exclusion on vertebrate scavenging communities. Sci. Rep. 10, 2644 (2020).
    ADS  CAS  PubMed  PubMed Central  Article  Google Scholar 

    69.
    Abernethy, E. F., Turner, K. L., Beasley, J. C. & Rhodes, O. E. Scavenging along an ecological interface: utilization of amphibian and reptile carcasses around isolated wetlands. Ecosphere 8, e01989 (2017).
    Article  Google Scholar 

    70.
    Olson, Z. H., Beasley, J. C. & Rhodes, O. E. Carcass type affects local scavenger guilds more than habitat connectivity. PLoS ONE 11, (2016).

    71.
    Ragg, J., Mackintosh, C. & Moller, H. The scavenging behaviour of ferrets (Mustela furo), feral cats (Felis domesticus), possums (Trichosurus vulpecula), hedgehogs (Erinaceus europaeus) and harrier hawks (Circus approximans) on pastoral farmland in New Zealand: Implications for bovine tuberculosis transmission. N. Z. Vet. J. 48, 166–175 (2001).
    Article  Google Scholar 

    72.
    Laundré, J. W., Hernández, L. & Altendorf, K. B. Wolves, elk, and bison: reestablishing the” landscape of fear” in Yellowstone National Park, USA. Can. J. Zool. 79, 1401–1409 (2001).
    Article  Google Scholar 

    73.
    Ripple, W. J. & Beschta, R. L. Linking wolves to willows via risk-sensitive foraging by ungulates in the northern Yellowstone ecosystem. For. Ecol. Manag. 230, 96–106 (2006).
    Article  Google Scholar 

    74.
    Ripple, W. J. & Beschta, R. L. Trophic cascades in Yellowstone: the first 15 years after wolf reintroduction. Biol. Conserv. 145, 205–213 (2012).
    Article  Google Scholar 

    75.
    Smith, D. W., Peterson, R. O. & Houston, D. B. Yellowstone after wolves. Bioscience 53, 330–340 (2003).
    Article  Google Scholar 

    76.
    White, P. J. & Garrott, R. A. Northern Yellowstone elk after wolf restoration. Wildl. Soc. Bull. 33, 942–955 (2005).
    Article  Google Scholar 

    77.
    Clark, P. J. & Evans, F. C. Distance to nearest neighbor as a measure of spatial relationships in populations. Ecology 35, 445–453 (1954).
    Article  Google Scholar 

    78.
    Cook, R. C., Cook, J. G. & Irwin, L. L. Estimating elk body mass using chest-girth circumference. Wildl. Soc. Bull. 1973-2006 31, 536–543 (2003).
    Google Scholar 

    79.
    Craine, J. M., Towne, E. G. & Elmore, A. Intra-annual bison body mass trajectories in a tallgrass prairie. Mammal Res. 60, 263–270 (2015).
    Article  Google Scholar 

    80.
    Lott, D. F. & Galland, J. C. Body mass as a factor influencing dominance status in American Bison Cows. J. Mammal. 68, 683–685 (1987).
    Article  Google Scholar 

    81.
    Fox, J. & Weisberg, S. An R Companion to Applied Regression. (Sage Publications, 2018).

    82.
    R Core Team. R: A Language and Environment for Statistical Computing. (R Foundation for Statistical Computing, 2020).

    83.
    Pan, Y. & Jackson, R. T. Ethnic difference in the relationship between acute inflammation and serum ferritin in US adult males. Epidemiol. Infect. 136, 421–431 (2008).
    CAS  PubMed  Article  PubMed Central  Google Scholar 

    84.
    Brewer, M. J., Butler, A. & Cooksley, S. L. The relative performance of AIC, AICC and BIC in the presence of unobserved heterogeneity. Methods Ecol. Evol. 679, 692. https://doi.org/10.1111/2041-210X.12541 (2016).
    Article  Google Scholar 

    85.
    Burnham, K. P. & Anderson, D. R. Multimodel inference: understanding AIC and BIC in model selection. Sociol. Methods Res. 33, 261–304 (2004).
    MathSciNet  Article  Google Scholar 

    86.
    Franklin, J. Mapping Species Distributions: Spatial Inference and Prediction (Cambridge University Press, Cambridge, 2010).
    Google Scholar 

    87.
    Kleiber, C. & Zeileis, A. Applied Econometrics with R (Springer, Berlin, 2008).
    Google Scholar 

    88.
    Cameron, A. C. & Trivedi, P. K. Regression-based tests for overdispersion in the Poisson model. J. Econom. 46, 347–364 (1990).
    MathSciNet  Article  Google Scholar  More