More stories

  • in

    Ant milk: The mysterious fluid that helps them thrive

    Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain
    the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in
    Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles
    and JavaScript. More

  • in

    Vulcanimicrobium alpinus gen. nov. sp. nov., the first cultivated representative of the candidate phylum “Eremiobacterota”, is a metabolically versatile aerobic anoxygenic phototroph

    Sampling, bacterial isolation, and colony screeningSamples were collected from thermally active sediments as evidenced by temperature and/or emission of steam in November 2010 and 2012 from Harry’s Dream (HD2 and HD3), Warren Cave (WC1, WC2, WC3, WC4, WC7, and WC8), Haggis Hole, (HH), Mammoth Cave (MC), Hut Cave (Hut), and Heroine Cave (HC) (fumarolic ice caves on Mt. Erebus volcano [21]). The samples were divided into four types: Mainly weathered basaltic/phonolitic sand (HD2, HD3, WC3, WC4, WC8 and Hut); pebbles and rock fragments (WC7, HH, and MC); black porous glassy materials that appeared to be solidified lava (WC1, WC2, and HC); and ash sediment (WC10) with little to no organic material (Table S1). The sediments were collected aseptically using sterile 50 mm conical tubes and immediately sealed. Additional information and environmental parameters on the sampling locations are provided in Table S1 and in Tebo et al. [21].To isolate bacteria found in these oligotrophic environments, we employed a culture strategy of long-term incubation in a nutrient-poor medium and screening of slow-growing colonies by direct PCR identification. Reasoner’s 2 A gellan gum medium (10% R2AG) [23] and FS1VG medium [24] were used for bacterial isolation. The 10% R2AG is a 10-fold diluted R2A broth (Nihon Seiyaku, Tokyo, Japan), solidified with 15 g/L gellan gum (Kanto chemical, Tokyo, Japan) with 2 g/L CaCl2. The 10% R2AG and FS1VG media were adjusted to a pH of 4.5 or 6.0, with or without 30 mg/L sodium azide. Samples used for isolation were selected from Warren Cave and Harry’s Dream, where the presence of relatively large numbers of bacteria ( >107/g) was confirmed in a previous report [21], and a total of nine samples, WD1, 2, 3, 4, 7, 8, and 10 and HD2 and 3, were used without any other pretreatment such as drying or dilution. For the sandy and ash samples with fine particles (WC3, 4, 8, 10, and HD2, 3), approximately 50 mg of sample were spread directly on plates. Glassy materials (WC1 and WC2) were embedded directly in plates using 2-3 pieces (approx. 200 mg). Pebbles and rock fragments (WC7) were crushed and approximately 50 mg of debris were spread directly onto the plates. All plates were incubated at 15 °C, 30 °C or 37 °C under dark conditions. New colonies were marked as they appeared and selection of the isolates was performed by picking only colonies that appeared after four weeks of incubation. We selected colonies around the sediment or in the cracks created during spreading using a magnifying glass.For identification, colonies were picked with a sterile toothpick, re-streaked, stabbed on fresh medium, and subsequently suspended in 20 μL sterilized 0.05 M NaOH. Suspensions were heated at 100 °C for 15 min, and supernatants were used as template DNAs for PCR. Partial 16 S rRNA gene sequences were amplified by PCR using commonly used bacterial primer set 27 F (5′-AGATTTGATCCTGGCTCAG-3′) and 1492 R (5′-GGTTACCTTGTTACGACTT-3′) or 536 R (5′-GTA TTA CCG CGG CTG CTG-3′) with TaKaRa ExTaq DNA polymerase (Takara Bio, Shiga, Japan) as previously described [23]. Sequencing was performed at Eurofins Genomics (Louisville, KY, USA), using a 3730xl DNA analyzer (Applied Biosystems, CA, USA). Sequence similarities with closest species were calculated using EZbiocloud’s Identify Service (https://www.ezbiocloud.net/identify). Subsequently, cells in the stab identified as “Ca. Eremiobacterota” by the direct PCR identification were serially diluted and stabbed onto new plates until multiple pure cultures of Eremiobacterota were obtained. The isolate was designated as strain WC8-2.Whole genome sequencing and annotationGenomic DNA was extracted from WC8-2 cells grown in 10% R2A broth (pH6.0) with air/CO2 (90:10, v/v) at 30 °C for 30 days under a 12/12 h light/dark regime with incandescent light (250 μmol m–2s–1), using a Puregene Yeast/Bact. Kit B (Qiagen, Germantown, MD, USA) [25]. Sequencing was performed by Macrogen Japan Corp., on a NovaSeq 6000 (Illumina, Inc., San Diego, CA, USA) and PacBio RSII (Pacific Biosciences of California, Inc. Menlo Park, CA, USA). The gap-free complete genome was assembled de novo using the Unicycler version 0.4.8 hybrid assembly pipeline with default settings [26]. Completeness and contamination levels were estimated using CheckM [27]. The genome was annotated using the DDBJ Fast Annotation and Submission Tool (DFAST) [28] and the BlastKOALA web server version 2.2, and was visualized using CGView Server [29] (http://cgview.ca/).Phylogenetic analysis of “Ca. Eremiobacterota”Identification of strain WC8-2 was performed using the Genome Taxonomy Database Toolkit (GTDB-Tk) (ver. 2.1.0), which produces standardized taxonomic labels that are based on those used in the Genome Taxonomy Database [30]. Terrabacterial genomes including “Ca. Eremiobacterota” MAGs and related genomes were retrieved from the Genome Taxonomy Database (GTDB) (July 2022) and the NCBI RefSeq database (July 2022). Full-length 16 S rRNA gene sequences were retrieved from the WC8-2 genome (WPS_r00030) and the NCBI database (Table S2). Multiple sequences were aligned using SINA (version 1.2.11) [31]. IQ-TREE version 1.6.12 [32] was used to build the phylogeny. ModelFinder [33] was used to determine the optimal evolutionary model for phylogeny building, which selected the TNe+I + G4 model. Branch support was calculated using 1000 ultrafast bootstraps [34]. The pairwise 16 S rRNA gene sequence similarities were determined using SDT software. Phylogenomic analysis based on 400 marker proteins was carried out using PhyloPhlAn v3.0 [35]. Diamond v5.2.32 [36], MAFFT v7.453 [37], and TrimAI were utilized for orthologs searching, multiple sequence alignment within the superphylum Terrabacteria, and gap-trimming, respectively. Gappy sites and sequences with >50% gaps were deleted from the alignments. IQ-TREE version 1.6.12 [32] was used to build the phylogenomic tree. ModelFinder [33] was used to determine the optimal evolutionary model for phylogeny building, which selected the LG + F + R9 model. These analyses were conducted using the “AOBA-B” super- computer (NEC, Tokyo, Japan) with 2CPUs (EPYC7702, AMD, CA, US) and 256GB of RAM. The related similarity of genomes between strain WC8-2 and relatives was estimated using average nucleotide identity (ANI) values, which were calculated using OrthoANI calculator in the EzBio-Cloud web service [38]. The related similarity between strain WC8-2 and its sister phyla with one representative from each class was assessed by pairwise Average Amino acid Identity (AAI) values using the online tools at the Kostas Konstantinidis Lab website Environmental Microbial Genomics Laboratory (http://enve-omics.ce.gatech.edu/aai/). The MAGs used in the tree are listed in Table S3.Phylogeny of photosynthesis- and “atmospheric chemosynthesis”- associated genesWe retrieved phototrophy- and “atmospheric chemosynthesis”-related protein sequences from the WC8-2 genome, “Ca. Eremiobacterota” MAGs, and known phototrophs genomes (Table S3) using the local BLAST server (SequenceServer 1.0.14 [39]) with reference sequences (Table S3) or annotated sequences as queries. Sequences were aligned using MUSCLE and poorly aligned regions were removed using Gblocks version 0.91b [40] or by manual inspection. The alignment sequences were concatenated into a single sequence. The ML tree was constructed using IQ-TREE with the best-hit evolutionary rate model: LG + I + G4 for HhyL, CbbL, BchXYZ and CbbL, LG + F + I + G4 for BchLNB, and LG + F + G4 for PufML, BchI, and BchD. All trees were visualized using iTOL (version 5.0) [41]. All sequences used in the trees are listed in Table S3.Analysis of bacterial communities in the fumarolic ice cavesEnvironmental DNA was extracted from the samples (0.1 g) (WC7, WC8, HD3, MC, Hut, HH, HC) using a DNeasy PowerSoil Kit (Qiagen, Valencia, CA, USA) following the manufacturer’s instructions. The V4 region of the 16 S rRNA gene was PCR amplified using primers with adapter sequences (V3-V4f_MIX) [42]. The PCR cycling was carried out using the following parameters: 94 °C for 2 min followed by 25 cycles at 94 °C for 30 s, 56 °C for 30 s, and 72 °C for 1 min with a final extension at 72 °C for 2 min. Library construction and sequencing were performed at the Bioengineering Lab (Kanagawa, Japan) using MiSeq (Illumina). Briefly, adaptor and primer regions were trimmed using the FASTX-Toolkit v0.0.13 (http://hannonlab.cshl.edu/fastx_toolkit). Read sequences of ≤40 bp with ambiguous bases and low-quality sequences (quality score, ≤Q20), together with their paired-end reads, were filtered out using Sickle v1.33 (https://github.com/najoshi/sickle). High-quality paired-end reads were merged using PEAR v0.9.10 with default settings [43]. Merged sequences of ≤245 and ≥260 bp were discarded using SeqKit v0.8.0 [44]. Operational taxonomic units (OTUs) were classified using QIIME v1.9.1 and the SILVA database (release 132) with 97% identity. To study the phylogeny of the OTUs assigned to the “Ca. Eremiobacterota”, a neighbor-joining (NJ) tree of the OTUs was constructed [45] as described above.Microscopic observationElectron microscopic observations of the cells were performed at Tokai Electron Microscopy (Nagoya, Japan). Briefly, the cells grown in 10% R2A broth (pH6.0) with air/CO2 (90:10, v/v) at 30 °C for 15 days under a 12/12 h light/dark regime with incandescent light (250 μmol m–2s–1) were fixed with 4% paraformaldehyde (PFA) and 4% glutaraldehyde (GA) in 0.1 M phosphate buffer (PB) at pH7.4, and postfixed with 2% OsO4 in 0.1 M PB. Cells were then dehydrated using graded ethanol solutions. The dehydrated cells were polymerized with resin, ultrathin sectioned, stained with 2% uranyl acetate, then secondary-stained with Lead stain solution. A transmission electron microscope (TEM) (JEM-1400Pus; JEOL, Tokyo, Japan) was used to observe the ultrathin sectioned cells at 100 kV acceleration voltage. To observe the negative-stained cells, PFA- and GA-fixed cells were adsorbed on formvar film-coated copper grids and stained with 2% phosphotungstic acid solution (pH 7.0) and observed using a TEM at 100 kV. DAPI (4,6-diamidino2-phenylindole) and Nile-Red staining was performed by incubating 0.1 mL cell suspension with a 1 mL staining solution (1 mg/L DAPI and 1 mg/L Nile Red in PBS buffer) for 10 min. The stained cells were observed under a fluorescence microscope (Olympus AX80T; Olympus Optical; Tokyo, Japan).Growth assayUnless otherwise noted, all cultures were grown in 100 mL butyl stopper- and screw-cap-sealed glass vials containing 50 mL liquid medium (pH6.0) at 30 °C. Growth was monitored by optical density at 600 nm (OD600) using a spectrophotometer (BioSpectrometer Basic; Eppendorf; Tokyo, Japan). Initial cell density was adjusted to 0.005 (OD600). Specific culture conditions are described below. All anaerobic growth tests were conducted with 100% N2 gas in the headspaces and supplemented with a reducing agent (0.3 g/L cysteine-HCl) and a redox indicator (1 mg/L resazurin).Optimal culture conditionsCell growth in different media was examined using 1, 10, 20, 100% (a full strength) R2A broth and Basal_YE with air/CO2 (90:10, v/v) under a 12/12 h light/dark regime with incandescent light (250 μmol m–2s–1) for 25 days. Basal_YE contained (l−1) 0.44 g KH2PO4, 0.1 g (NH4)2SO4, 0.1 g MgSO4.7H2O, 0.3 g yeast extract, and 1 ml trace element SL-8 [refer to DSMZ745]. Cell growth at different temperatures (10, 13, 20, 25, 30, 33 and 37 °C; pH6.0), pH (3.7, 4.5, 6.0, 7.0, 8.0, and 9.0), and NaCl (0, 1, 10, 20, and 30 g [l−1]; pH6.0) was examined using Basal_YE with air/CO2 (90:10, v/v) for 25 days under a 12/12 h light/dark regime with incandescent light (250 μmol m−2s−1). The following pH buffer solutions were used: acetic acid/sodium acetate for pH 4–6, K2HPO4/KH2PO4 for pH 6–8, sodium bicarbonate/sodium carbonate for pH 9–10. To examine colony formation on solid media, WC8-2 was streaked or stabbed on 10% R2AG medium (pH 6.0) and incubated at 30 °C for 30 days under a 100% air atmosphere.Optimal oxygen and carbon dioxide conditionsTo determine the preferred O2 concentration, WC8-2 was grown in Basal_YE (pH 6.0, 30 °C, no NaCl) in butyl stopper-sealed glass bottles with the atmosphere in the headspace adjusted to different N2/O2/CO2 ratios (70:20:10%, 80:10:10%, 89:1:10%, 90:0:10% v/v) after removing oxygen with 100% N2 gas. Cultures were incubated for 25 days under a 12/12 h light/dark regime with incandescent light (250 μmol m–2s–1). To determine the CO2 preference, the gas phase was adjusted to different N2/O2/CO2 ratios (70:20:10%,75:20:5%, 80:20:0% v/v) in sealed bottles or 100% air (plugged with a BIO-SILICO N-38 sponge plug; Shin-Etsu Polymer Co., Ltd, Tokyo, Japan; breathable culture-plug) under the same conditions as the O2 preference test.Photoorganoheterotrophic (or photoorganoautotrophic) and chemoorganoheterotrophic (or chemoorganoautotrophic) growthThe utilization of organic compounds as carbon sources/organic electron donors was tested in Basal medium (Basal_YE without yeast extract, pH 6.0) supplemented with one of the following sources (l-1): 0.3 ml of glycerol, or 0.3 g of sucrose, d-glucose, d-ribose, maltose, l-leucine, l-isoleucine, l-valine, l-serine, l-lysine, taurine, yeast extract or gellan gum, 1 ml of vitamin B12 solution (2 mg/L). Utilization was assessed by measuring growth of the cultures during a 25-day incubation at 30 °C in continuous light (250 μmol m–2s–1) for photoorganoheterotrophy, or continuous dark for chemoorganoheterotrophy under aerobic (air/CO2 [90:10, v/v]) and anaerobic (N2/CO2 [90:10, v/v]) conditions.Photolithoautotrophic and chemolithoautotrophic growthCells were inoculated into amended PSB2 [46] as described below or Basal medium with 5 mM Na2S or Na2S2O3 or 1% H2 (v/v; in the gas phase) as electron donors, and cultivated in continuous light (250 μmol m–2s–1) for photolithoautotrophic growth, or continuous dark for chemolithoautotrophic growth under aerobic (air/CO2 [90:10, v/v]) and anaerobic (N2/CO2 [90:10, v/v]) conditions for 60 days at 30 °C. The amended PBS2 contained (L-1): 0.5 g NH4Cl, 1.0 g KH2PO4, 0.2 g NaCl, 0.4 g MgSO4.7H2O, 0.05 g CaCl2.2H2O, 4.2 g NaHCO3, 1 ml trace element SL-8 [refer to DSMZ745], and 1 ml vitamin B12 solution (2 mg/L), pH6.0.Fermentative or anaerobic growthAnaerobic growth was examined in continuous dark in 20% R2A broth (pH 6.0) with N2/CO2 (90:10, v/v) supplemented with 5 mM Na2SO4, NaNO3, or dimethyl sulfoxide (DMSO) as electron acceptors for 60 days at 30 °C.Pigment assaysCells grown in Basal_YE (pH6.0) with air/CO2 (90:10, v/v) at 30 °C for 14 days (exponential growth phase) under continuous light (250 μmol m–2 s–1) and continuous dark were used for the pigment assays. The absorption spectrum was determined in a cell suspension in 60% (w/v) sucrose and in a 100% methanol extract using a spectrophotometer (V-630; JASCO, Tokyo, Japan) at 350–1100 nm. The BChl a concentration was determined spectroscopically in 100% methanol [47]. Dry cell weight was measured after harvested cells were washed twice with Milli-Q water and dried at 80 °C for 3 days. The extract was also analyzed by HPLC (NEXERA X2; Shimadzu; Kyoto, Japan) equipped with a 4.6 × 250 mm COSMOSIL 5C18-AR (Nakarai Taque; Tokyo, Japan) with isocratic elution of 92.5% (v/v) methanol in water at a flow rate of 1.0 mL/min. BChl a was monitored at 766 nm using a diode-array spectrophotometer detector (SPD-M20A; Shimadzu; Kyoto, Japan).Observation of taxisTo study phototaxis in WC8-2, cells were grown in 20% R2A broth with air/CO2 (90:10, v/v) for 14 days under a 12/12 h light/dark regimen. Cultures were transferred to tissue culture flasks (175 cm2, canted neck, Iwaki, Shizuoka, Japan). Light sources were ultraviolet (UV) at 395 nm (Linkman, Fukui, Japan), blue at 470 nm (CREE, Durham, NC, USA), green at 502 nm (Linkman), red at 653 nm (LENOO, Shinpei, Taiwan), and near-infrared (NIR) at 880 nm (LENOO). The light-emitting device was constructed by assembling LEDs on a breadboard with a power supply. Cultivations were illuminated with each wavelength using a light-emitting device from the underside and incubated at 20 °C for 18 h. Cells aggregating toward a light source was taken to indicate phototaxis. As a control, culture vessels were wrapped in aluminum foil to block light. Images and time-lapse video were captured using an iPhone 6 S camera.Stable carbon isotope ratio mass spectrometry (IRMS)The WC8-2 cells were grown in Basal_YE (pH6.0) with air/13CO2 [90:10, v/v] and air/unlabeled CO2 [90:10, v/v] under continuous light (250 μmol m–2s–1) for photoheterotroph and continuous dark for chemoheterotroph at 30 °C for 14 days (exponential growth phase). Approximately 10 mg culture biomass was collected, washed in HCl overnight, rinsed three times with deionized water, and placed into tin capsules. Stable carbon isotope ratios (δ13C) were analyzed at Shoko Science (Saitama, Japan) using a Delta V Advantage (EA-IRMS; Thermo Fisher Scientific, Bremen, Germany). The standard for C isotope ratio analysis was Vienna PeeDee Belemnite (VPDB). The δ13C values of 13CO2-cultivated cells exceeded the optimum calibration range of the instrument, but were used in this study to provide conclusive evidence that inorganic carbon was incorporated into the biomass.Quantitative reverse transcription PCR (qRT-PCR)Total RNA extraction and cDNA synthesisTotal RNA was extracted from cells grown in Basal_YE (pH6.0) with air/CO2 [90:10, v/v] in continuous light (250 μmol m–2s–1) for photoheterotrophic conditions and in continuous dark for chemoheterotrophic conditions at 30 °C for 14 days (exponential growth phase) using the Total RNA Purification Kit (Norgen, Biotek Corp, Ontario, Canada). DNA was removed from the extracted nucleic acids using an RNase-Free DNase I Kit (Norgen, Biotek Corp) according to the manufacturer’s protocol. The absence of DNA in the RNA samples was confirmed by PCR without reverse transcriptase. cDNA was generated from 500 ng total RNA using a TaKaRa PrimeScript™ 1st strand cDNA Synthesis Kit (TaKaRa Bio) with random hexamers according to the manufacturer’s protocol.Primer design, specificity and efficiencyThe following three photosynthesis- and CO2-fixation-related genes in the WC8-2 genome were selected for qRT-PCR: bchM encoding an enzyme involved in BChl synthesis, pufL encoding the anoxygenic Type II photochemical reaction centers L-subunit, and cbbL encoding the large subunit of type IE RuBisCO. The RNA polymerase subunit beta (rpoB) was used as a housekeeping reference gene. Primers for qRT-PCR were designed with Primer3 (v. 0.4.0) [48] with the following criteria: product size ranging from 80 to 150 bp, optimum Tm of 60 °C and GC content about 50 to 55%. Standard RT-PCR confirmed that each primer set amplified only a single product with expected size (data not shown), and the product was also sequenced using Sanger sequencing at Macrogen Japan Corp to confirm the candidate products. Primer efficiency was calculated for qRT-PCR using the slope of the calibration curve based on a 20-, 40-, 80-, 160-fold dilution series of cDNA samples [49]. In addition, the specificity of the primers was determined by the confirmation of a single peak in the melting curve. All information about the primers is shown in Table S4.qRT-PCRqRT-PCR was performed using a MiniOpticon Real-Time PCR System (Bio-Rad, Marnes la Coquette, France). The reaction mixture contained 10 μL TB Green Premix Ex Taq II (Tli RNaseH Plus, Takara Bio), 0.8 μL 10 mM primer, 2 μL of a 20-, 40-, 80-, 160-fold dilution series of cDNA, and 6.4 μL water. qRT-PCR was performed using the following protocol: denaturation at 95 °C for 30 s; denaturation and amplification at 95 °C for 5 s and 60 °C for 30 s, respectively (40 cycles). Fluorescence was measured at the end of the amplification step, and amplified products were examined by melting curve analysis from 60 to 95 °C. Each reaction was performed in three independent cultivations. Relative gene expression fold change was calculated using the comparative Ct method (2−ΔΔCt) [49]. Normalized expressions were used for reaction in dark. The 2−ΔΔCt values ≤0.5 were defined as downregulated and values ≥2.0 as upregulated, with P  More

  • in

    Reproductive performance and sex ratio adjustment of the wild boar (Sus scrofa) in South Korea

    Estes, J. A. Predators and ecosystem management. Wildl. Soc. Bull. 24, 390–396 (1996).
    Google Scholar 
    Licht, D. S., Millspaugh, J. J., Kunkel, K. E., Kochanny, C. O. & Peterson, R. O. Using small populations of wolves for ecosystem restoration and stewardship. Bioscience 60, 147–153 (2010).Article 

    Google Scholar 
    Schwartz, C. C., Swenson, J. E. & Miller, S. D. Large carnivores, moose, and humans: A changing paradigm of predator management in the 21st century. Alces J. Devot. Biol. Manag. Moose 39, 41–63 (2003).
    Google Scholar 
    Valente, A. M., Acevedo, P., Figueiredo, A. M., Fonseca, C. & Torres, R. T. Overabundant wild ungulate populations in Europe: Management with consideration of socio-ecological consequences. Mammal Rev. 50, 353–366 (2020).Article 

    Google Scholar 
    Lee, S. D. & Miller-Rushing, A. J. Degradation, urbanization, and restoration: A review of the challenges and future of conservation on the Korean Peninsula. Biol. Cons. 176, 262–276 (2014).Article 

    Google Scholar 
    Kodera, Y. Habitat selection of Japanese wild boar in Iwami district, Shimane Prefecture, western Japan. Wildl. Conserv. Jpn. 6, 119–129 (2001).
    Google Scholar 
    Ohashi, H. et al. Differences in the activity pattern of the wild boar Sus scrofa related to human disturbance. Eur. J. Wildl. Res. 59, 167–177 (2013).Article 

    Google Scholar 
    Ministry of Environment Republic of Korea. Management Plan of Pest Wild Boars. (Sejong, 2010).National Institute of Biological Resources. Analysis of Hunting Effect on Pest Animals and its Monitoring Techniques. (Incheon, 2017).Lee, S. M. & Lee, E. J. Diet of the wild boar (Sus scrofa): Implications for management in forest-agricultural and urban environments in South Korea. PeerJ 7, e7835 (2019).Article 

    Google Scholar 
    Bieber, C. & Ruf, T. Population dynamics in wild boar Sus scrofa: ecology, elasticity of growth rate and implications for the management of pulsed resource consumers. J. Appl. Ecol. 42, 1203–1213 (2005).Article 

    Google Scholar 
    Brogi, R. et al. Capital-income breeding in wild boar: A comparison between two sexes. Sci. Rep. 11, 1–10 (2021).Article 

    Google Scholar 
    Frauendorf, M., Gethöffer, F., Siebert, U. & Keuling, O. The influence of environmental and physiological factors on the litter size of wild boar (Sus scrofa) in an agriculture dominated area in Germany. Sci. Total Environ. 541, 877–882 (2016).Article 
    ADS 
    CAS 

    Google Scholar 
    Massei, G., Genov, P. V. & Staines, B. W. Diet, food availability and reproduction of wild boar in a Mediterranean coastal area. Acta Theriol. 41, 307–320 (1996).Article 

    Google Scholar 
    Sabrina, S., Jean-Michel, G., Carole, T., Serge, B. & Eric, B. Pulsed resources and climate-induced variation in the reproductive traits of wild boar under high hunting pressure. J. Anim. Ecol. 78, 1278–1290 (2009).Article 

    Google Scholar 
    Fonseca, C. et al. Reproduction in the wild boar (Sus scrofa Linnaeus, 1758) populations of Portugal. Galemys 16, 53–65 (2004).
    Google Scholar 
    Morreti, M. Birth distribution, structure and dynamics of a hunted mountain populatin of wild boars (Sus scrofa L.), Ticino, Switzerland. J. Mt. Ecol. 3, 192–196 (1995).
    Google Scholar 
    Rosell, C., Navas, F. & Romero, S. Reproduction of wild boar in a cropland and coastal wetland area: Implications for management. Anim. Biodivers. Conserv. 35, 209–217 (2012).Article 

    Google Scholar 
    Cellina, S. Effects of supplemental feeding on the body condition and reproductive state of wild boar Sus scrofa in Luxembourg PhD Dissertation thesis, University of Sussex, (2008).Gamelon, M. et al. High hunting pressure selects for earlier birth date: Wild boar as a case study. Evol. Int. J. Org. Evol. 65, 3100–3112 (2011).Article 

    Google Scholar 
    Gethöffer, F., Sodeikat, G. & Pohlmeyer, K. Reproductive parameters of wild boar (Sus scrofa) in three different parts of Germany. Eur. J. Wildl. Res. 53, 287–297 (2007).Article 

    Google Scholar 
    Fonseca, C., Da Silva, A., Alves, J., Vingada, J. & Soares, A. Reproductive performance of wild boar females in Portugal. Eur. J. Wildl. Res. 57, 363–371 (2011).Article 

    Google Scholar 
    Gaillard, J.-M., Brandt, S. & Jullien, J.-M. Body weight effect on reproduction of young wild boar (Sus scrofa) females: A comparative analysis. Folia Zool. (Brno) 42, 204–212 (1993).
    Google Scholar 
    Poteaux, C. et al. Socio-genetic structure and mating system of a wild boar population. J. Zool. 278, 116–125 (2009).Article 

    Google Scholar 
    Spitz, F., Valet, G. & Lehr Brisbin, I. Jr. Variation in body mass of wild boars from southern France. J. Mammal. 79, 251–259 (1998).Article 

    Google Scholar 
    Trivers, R. L. & Willard, D. E. Natural selection of parental ability to vary the sex ratio of offspring. Science 179, 90–92 (1973).Article 
    ADS 
    CAS 

    Google Scholar 
    Clutton-Brock, T. H., Albon, S. & Guinness, F. Parental investment in male and female offspring in polygynous mammals. Nature 289, 487–489 (1981).Article 
    ADS 

    Google Scholar 
    Hewison, A. M. & Gaillard, J.-M. Successful sons or advantaged daughters? The Trivers-Willard model and sex-biased maternal investment in ungulates. Trends Ecol. Evol. 14, 229–234 (1999).Article 
    CAS 

    Google Scholar 
    Clutton-Brock, T. H. & Iason, G. R. Sex ratio variation in mammals. Q. Rev. Biol. 61, 339–374 (1986).Article 
    CAS 

    Google Scholar 
    Fernández-Llario, P. & Mateos-Quesada, P. Body size and reproductive parameters in the wild boar Sus scrofa. Acta Theriol. 43, 439–444 (1998).Article 

    Google Scholar 
    Meikle, D. B., Drickamer, L. C., Vessey, S. H., Rosenthal, T. L. & Fitzgerald, K. S. Maternal dominance rank and secondary sex ratio in domestic swine. Anim. Behav. 46, 79–85 (1993).Article 

    Google Scholar 
    Servanty, S., Gaillard, J.-M., Allainé, D., Brandt, S. & Baubet, E. Litter size and fetal sex ratio adjustment in a highly polytocous species: The wild boar. Behav. Ecol. 18, 427–432 (2007).Article 

    Google Scholar 
    Mendl, M., Zanella, A. J., Broom, D. M. & Whittemore, C. T. Maternal social status and birth sex ratio in domestic pigs: An analysis of mechanisms. Anim. Behav. 50, 1361–1370 (1995).Article 

    Google Scholar 
    Cameron, E. Z. Facultative adjustment of mammalian sex ratios in support of the Trivers-Willard hypothesis: Evidence for a mechanism. Proc. R. Soc. Lond. Ser. B Biol. Sci. 271, 1723–1728 (2004).Article 

    Google Scholar 
    Clutton-Brock, T., Albon, S. & Guinness, F. Maternal dominance, breeding success and birth sex ratios in red deer. Nature 308, 358–360 (1984).Article 
    ADS 

    Google Scholar 
    Maillard, D. & Fournier, P. Timing and synchrony of births in the wild boar (Sus scrofa Linnaeus, 1758) in a Mediterranean habitat: The effect of food availability. Galemys 16, 67–74 (2004).
    Google Scholar 
    Bywater, K. A., Apollonio, M., Cappai, N. & Stephens, P. A. Litter size and latitude in a large mammal: the wild boar Sus scrofa. Mammal Rev. 40, 212–220 (2010).
    Google Scholar 
    Orłowska, L., Rembacz, W. & Florek, C. Carcass weight, condition and reproduction of wild boars harvested in north-western Poland. Pest Manag. Sci. 69, 367–370 (2013).Article 

    Google Scholar 
    Carranza, J. Sexual selection for male body mass and the evolution of litter size in mammals. Am. Nat. 148, 81–100 (1996).Article 

    Google Scholar 
    FernáNdez-Llario, P., Carranza, J. & Mateos-Quesada, P. Sex allocation in a polygynous mammal with large litters: The wild boar. Anim. Behav. 58, 1079–1084 (1999).Article 

    Google Scholar 
    McBride, G. The” teat order” and communication in young pigs. Animal Behaviour (1963).McBride, G., James, J. & Hodgens, N. Social behaviour of domestic animals. IV. Growing pigs. Anim. Sci. 6, 129–139 (1964).Article 

    Google Scholar 
    McBride, G., James, J. & Wyeth, G. Social behaviour of domestic animals VII. Variation in weaning weight in pigs. Anim. Sci. 7, 67–74 (1965).Article 

    Google Scholar 
    Geochang County. Geochang Statistical yearbook. (Geochang, 2015).Seoul Metropolitan Government. Seoul Statistical Yearbook. (Seoul, 2017).Animal and Plant Quarantine Agency and Ministry of Food and Drug Safety. Institutional Animal Care and Use Committee Standard Operation Guideline. (Gimcheon, 2020).Magnell, O. & Carter, R. The chronology of tooth development in wild boar – A guide to age determination of linear enamel hypoplasia in prehistoric and medieval pigs. Verterrinarija Ir Zootechnika. T. 40, 43–48 (2007).
    Google Scholar 
    Oroian, T. E., Oroian, R. G., Pasca, I., Oroian, E. & Covrig, I. Methods of age estimation by dentition in Sus scrofa ferus sp. Bull. UASVM Anim. Sci. Biotechnol. 67, 1–2 (2010).
    Google Scholar 
    Vericad, R. Fetal age, conception and birth period estimation on wild boar (Sus scrofa) in West Pyrenees. in Actas del XV Congresso Int. Fauna Cinegética y Silvestre. (Trujillo, 1983).Rosell, C., Fernández-Llario, P. & Herrero, J. The wild boar (Sus scrofa Linnaeus, 1758). Galemys 13, 1–25 (2001).
    Google Scholar 
    R core team. R: A language and environment for statistical computing v. 3.6.0 (Austria, 2019). More

  • in

    Grazing pressure on drylands

    Maestre and colleagues collected data using a standardized field survey at 98 sites across 25 countries and 6 continents, fitted linear mixed models to data from all sites and grazing pressure levels, and then applied a multimodel inference procedure to select the set of best-fitting models. The authors found interactions between grazing and biodiversity in almost half of the best-fitting models, where increasing grazing pressure had positive effects on ecosystem services in colder sites with high plant species richness. However, increases in grazing pressure at warmer sites with high rainfall seasonality and low plant species richness interacted with soil properties to either increase or reduce the delivery of multiple ecosystem services. The authors’ findings highlight how increasing herbivore richness could enhance ecosystem service delivery across contrasting environmental and biodiversity conditions, enhancing soil carbon storage and reducing the negative impacts of increased grazing pressure. More

  • in

    Thermal physiology integrated species distribution model predicts profound habitat fragmentation for estuarine fish with ocean warming

    Reygondeau, G. & Beaugrand, G. Future climate-driven shifts in distribution of Calanus finmarchicus. Glob. Change Biol. 17, 756–766 (2011).Article 
    ADS 

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

    Google Scholar 
    Bosso, L. et al. The rise and fall of an alien: Why the successful colonizer Littorina saxatilis failed to invade the Mediterranean Sea. Biol. Invasions 24, 3169–3187 (2022).Article 

    Google Scholar 
    Guisan, A. & Zimmermann, N. E. Predictive habitat distribution models in ecology. Ecol. Model. 135, 147–186 (2000).Article 

    Google Scholar 
    Guisan, A. & Thuiller, W. Predicting species distribution: offering more than simple habitat models. Ecol. Lett. 8, 993–1009 (2005).Article 

    Google Scholar 
    Kaschner, K., Watson, R., Trites, A. W. & Pauly, D. Mapping world-wide distributions of marine mammal species using a relative environmental suitability (RES) model. Mar. Ecol. Prog. Ser. 316, 285–310 (2006).Article 
    ADS 

    Google Scholar 
    Pearson, R. G. & Dawson, T. P. Predicting the impacts of climate change on the distribution of species: are bioclimate envelope models useful?. Glob. Ecol. Biogeogr. 12, 361–371 (2003).Article 

    Google Scholar 
    Buckley, L. B. Linking traits to energetics and population dynamics to predict lizard ranges in changing environments. Am. Nat. 171, E1–E19 (2008).Article 

    Google Scholar 
    Kolbe, J. J., Kearney, M. & Shine, R. Modeling the consequences of thermal trait variation for the cane toad invasion of Australia. Ecol. Appl. 20, 2273–2285 (2010).Article 

    Google Scholar 
    Sanford, E. & Kelly, M. W. Local adaptation in marine invertebrates. Ann. Rev. Mar. Sci. 3, 509–535 (2011).Article 

    Google Scholar 
    Somero, G. N., Lockwood, B. L. & Tomanek, L. Biochemical Adaptation: Response to Environmental Challenges, From Life’s Origins to the Anthropocene (Sinauer Associates, 2017).
    Google Scholar 
    Kuo, E. S. & Sanford, E. Geographic variation in the upper thermal limits of an intertidal snail: Implications for climate envelope models. Mar. Ecol. Prog. Ser. 388, 137–146 (2009).Article 
    ADS 

    Google Scholar 
    Smeraldo, S. et al. Ignoring seasonal changes in the ecological niche of non-migratory species may lead to biases in potential distribution models: lessons from bats. Biodivers. Conserv. 27, 2425–2441 (2018).Article 

    Google Scholar 
    Gamliel, I. et al. Incorporating physiology into species distribution models moderates the projected impact of warming on selected Mediterranean marine species. Ecography 43, 1090–1106 (2020).Article 

    Google Scholar 
    Kearney, M. R., Wintle, B. A. & Porter, W. P. Correlative and mechanistic models of species distribution provide congruent forecasts under climate change. Conserv. Lett. 3, 203–213 (2010).Article 

    Google Scholar 
    Buckley, L. B., Waaser, S. A., MacLean, H. J. & Fox, R. Does including physiology improve species distribution model predictions of responses to recent climate change?. Ecology 92, 2214–2221 (2011).Article 

    Google Scholar 
    Fry, F. E. J. Effects of the environment on animal activity. Pub. Ontario Fish. Lab. No. 68. Toronto Studies Biol. Ser. 55, 1–52 (1947).
    Google Scholar 
    Brett, J. R. Energetic responses of salmon to temperature. A study of some thermal relations in the physiology and freshwater ecology of sockeye salmon (Oncorhynchus nerkd). Am Zoologist 11, 99–113 (1971).Article 

    Google Scholar 
    Pörtner, H. O. & Knust, R. Climate change affects marine fishes through the oxygen limitation of thermal tolerance. Science 315, 95–97 (2007).Article 
    ADS 

    Google Scholar 
    Pörtner, H. O. & Farrell, A. P. Physiology and climate change. Science 322, 690–692 (2008).Article 

    Google Scholar 
    Eliason, E. J. et al. Differences in thermal tolerance among sockeye salmon populations. Science 332, 109–112 (2011).Article 
    ADS 
    CAS 

    Google Scholar 
    Donelson, J. M., Munday, P. L., McCormick, M. I. & Pitcher, C. R. Rapid transgenerational acclimation of a tropical reef fish to climate change. Nat. Clim. Change 2, 30–32 (2012).Article 
    ADS 

    Google Scholar 
    Pörtner, H. Climate change and temperature-dependent biogeography: Oxygen limitation of thermal tolerance in animals. Naturwissenschaften 88, 137–146 (2001).Article 
    ADS 

    Google Scholar 
    Pörtner, H.-O. Oxygen-and capacity-limitation of thermal tolerance: A matrix for integrating climate-related stressor effects in marine ecosystems. J. Exp. Biol. 213, 881–893 (2010).Article 

    Google Scholar 
    Clark, T. D., Sandblom, E. & Jutfelt, F. Response to Farrell and to Pörtner and Giomi. J. Exp. Biol. 216, 4495–4497 (2013).Article 

    Google Scholar 
    Farrell, A. P. Aerobic scope and its optimum temperature: Clarifying their usefulness and limitations: Correspondence on J. Exp. Biol. 216, 2771–2782. J. Exp. Biol. 216, 4493–4494 (2013).Article 

    Google Scholar 
    Dillon, M. E., Wang, G. & Huey, R. B. Global metabolic impacts of recent climate warming. Nature 467, 704–706 (2010).Article 
    ADS 
    CAS 

    Google Scholar 
    Deutsch, C., Ferrel, A., Seibel, B., Pörtner, H.-O. & Huey, R. B. Climate change tightens a metabolic constraint on marine habitats. Science 348, 1132–1135 (2015).Article 
    ADS 
    CAS 

    Google Scholar 
    Gillooly, J. F., Brown, J. H., West, G. B., Savage, V. M. & Charnov, E. L. Effects of size and temperature on metabolic rate. Science 293, 2248–2251 (2001).Article 
    ADS 
    CAS 

    Google Scholar 
    Brown, J. H., Gillooly, J. F., Allen, A. P., Savage, V. M. & West, G. B. Toward a metabolic theory of ecology. Ecology 85, 1771–1789 (2004).Article 

    Google Scholar 
    Clarke, A. Is there a universal temperature dependence of metabolism?. Funct. Ecol. 18, 252–256 (2004).Article 

    Google Scholar 
    Clarke, A. & Fraser, K. P. P. Why does metabolism scale with temperature?. Funct. Ecol. 18, 243–251 (2004).Article 

    Google Scholar 
    Fangue, N. A., Hofmeister, M. & Schulte, P. M. Intraspecific variation in thermal tolerance and heat shock protein gene expression in common killifish, Fundulus heteroclitus. J. Exp. Biol. 209, 2859–2872 (2006).Article 
    CAS 

    Google Scholar 
    Dhillon, R. S. & Schulte, P. M. Intraspecific variation in the thermal plasticity of mitochondria in killifish. J. Exp. Biol. 214, 3639–3648 (2011).Article 
    CAS 

    Google Scholar 
    Fangue, N. A., Podrabsky, J. E., Crawshaw, L. I. & Schulte, P. M. Countergradient variation in temperature preference in populations of killifish Fundulus heteroclitus. Physiol. Biochem. Zool. 82, 776–786 (2009).Article 

    Google Scholar 
    Healy, T. M. & Schulte, P. M. Thermal acclimation is not necessary to maintain a wide thermal breadth of aerobic scope in the common killifish (Fundulus heteroclitus). Physiol. Biochem. Zool. 85, 107–119 (2012).Article 
    CAS 

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

    Google Scholar 
    Norin, T., Malte, H. & Clark, T. D. Aerobic scope does not predict the performance of a tropical eurythermal fish at elevated temperatures. J. Exp. Biol. 217, 244–251 (2014).
    Google Scholar 
    Payne, N. L. et al. Temperature dependence of fish performance in the wild: links with species biogeography and physiological thermal tolerance. Funct. Ecol. 30, 903–912 (2016).Article 

    Google Scholar 
    Raffel, T. R. et al. Disease and thermal acclimation in a more variable and unpredictable climate. Nat. Clim. Change 3, 146–151 (2013).Article 
    ADS 

    Google Scholar 
    Sinclair, B. J. et al. Can we predict ectotherm responses to climate change using thermal performance curves and body temperatures?. Ecol. Lett. 19, 1372–1385 (2016).Article 

    Google Scholar 
    Dahlke, F. T. et al. Northern cod species face spawning habitat losses if global warming exceeds 1.5°C. Sci. Adv. 4, 8821 (2018).Article 
    ADS 

    Google Scholar 
    Pörtner, H.-O. & Giomi, F. Nothing in experimental biology makes sense except in the light of ecology and evolution: Correspondence on J. Exp. Biol. 2771-2782. J. Exp. Biol. 216, 4494–4495 (2013).Article 

    Google Scholar 
    Pörtner, H.-O. How and how not to investigate the oxygen and capacity limitation of thermal tolerance (OCLTT) and aerobic scope: Remarks on the article by Gräns et al. J. Exp. Biol. 217, 4432–4433 (2014).Article 

    Google Scholar 
    Kleiber, M. Body size and metabolism. Hilgardia 6, 315–353 (1932).Article 
    CAS 

    Google Scholar 
    Killen, S. S., Atkinson, D. & Glazier, D. S. The intraspecific scaling of metabolic rate with body mass in fishes depends on lifestyle and temperature. Ecol. Lett. 13, 184–193 (2010).Article 

    Google Scholar 
    Norin, T. & Gamperl, A. K. Metabolic scaling of individuals vs. populations: Evidence for variation in scaling exponents at different hierarchical levels. Funct. Ecol. 32, 379–388 (2018).Article 

    Google Scholar 
    Jayasundara, N., Kozal, J. S., Arnold, M. C., Chan, S. S. L. & Giulio, R. T. D. High-throughput tissue bioenergetics analysis reveals identical metabolic allometric scaling for teleost hearts and whole organisms. PLoS ONE 10, e0137710 (2015).Article 

    Google Scholar 
    Kinnison, M. T., Unwin, M. J. & Quinn, T. P. Migratory costs and contemporary evolution of reproductive allocation in male chinook salmon. J. Evol. Biol. 16, 1257–1269 (2003).Article 
    CAS 

    Google Scholar 
    Clarke, A. & Johnston, N. M. Scaling of metabolic rate with body mass and temperature in teleost fish. J. Anim. Ecol. 68, 893–905 (1999).Article 

    Google Scholar 
    Duvernell, D. D., Lindmeier, J. B., Faust, K. E. & Whitehead, A. Relative influences of historical and contemporary forces shaping the distribution of genetic variation in the Atlantic killifish, Fundulus heteroclitus. Mol. Ecol. 17, 1344–1360 (2008).Article 

    Google Scholar 
    Navarro-Racines, C., Tarapues, J., Thornton, P., Jarvis, A. & Ramirez-Villegas, J. High-resolution and bias-corrected CMIP5 projections for climate change impact assessments. Sci. Data 7, 1–14 (2020).Article 

    Google Scholar 
    Franke, R. Scattered data interpolation: Tests of some methods. Math. Comput. 38, 181–200 (1982).MathSciNet 
    MATH 

    Google Scholar 
    Levitus, S. et al. World ocean heat content and thermosteric sea level change (0–2000 m), 1955–2010. Geophys. Res. Lett. 39, 15. https://doi.org/10.1029/2012GL051106 (2012).Article 

    Google Scholar 
    Kaschner, K. et al. AquaMaps: Predicted Range Maps for Aquatic Species (Worldwide Web Electronic Publication, 2019).
    Google Scholar 
    Jayasundara, N. Ecological significance of mitochondrial toxicants. Toxicology 391, 64–74 (2017).Article 
    CAS 

    Google Scholar 
    Beers, J. M. & Jayasundara, N. Antarctic notothenioid fish: what are the future consequences of ‘losses’ and ‘gains’ acquired during long-term evolution at cold and stable temperatures?. J. Exp. Biol. 218, 1834–1845 (2015).Article 

    Google Scholar 
    Lear, K. O. et al. Thermal performance responses in free-ranging elasmobranchs depend on habitat use and body size. Oecologia 191, 829–842 (2019).Article 
    ADS 

    Google Scholar 
    Good, S. et al. The current configuration of the OSTIA system for operational production of foundation sea surface temperature and ice concentration analyses. Remote Sens. 12, 720 (2020).Article 
    ADS 

    Google Scholar 
    Stocker, T. Climate Change 2013: The Physical Science Basis: Working Group I Contribution to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change (Cambridge University Press, 2014).
    Google Scholar 
    Ready, J. et al. Predicting the distributions of marine organisms at the global scale. Ecol. Model. 221, 467–478 (2010).Article 

    Google Scholar 
    Pawlowicz, R. M_Map: A Mapping Package for MATLAB, Version 1.4 m. Computer Software, UBC EOAS. https://www.eoas.ubc.ca/rich/map.html (2020).Schulzweida, U., Kornblueh, L. & Quast, R. CDO User’s Guide. Climate Data Operators, Version 1, (2006).Nychka, D., Furrer, R., Paige, J. & Sain, S. Fields: Tools for Spatial Data. R Package Version 11.6. (2017).Chen, Z., Farrell, A. P., Matala, A. & Narum, S. R. Mechanisms of thermal adaptation and evolutionary potential of conspecific populations to changing environments. Mol. Ecol. 27, 659–674 (2018).Article 

    Google Scholar 
    da Silva, C. R. B., Riginos, C. & Wilson, R. S. An intertidal fish shows thermal acclimation despite living in a rapidly fluctuating environment. J. Comp. Physiol. B. 189, 385–398 (2019).Article 

    Google Scholar 
    Slesinger, E. et al. The effect of ocean warming on black sea bass (Centropristis striata) aerobic scope and hypoxia tolerance. PLoS ONE 14, e0218390 (2019).Article 
    CAS 

    Google Scholar 
    Moffett, E. R., Fryxell, D. C., Palkovacs, E. P., Kinnison, M. T. & Simon, K. S. Local adaptation reduces the metabolic cost of environmental warming. Ecology 99, 2318–2326 (2018).Article 

    Google Scholar 
    Turker, H. The effect of water temperature on standard and routine metabolic rate in two different sizes of Nile tilapia. Kafkas Universitesi Veteriner Fakultesi Dergisi 17, 575–580 (2011).
    Google Scholar 
    Hvas, M., Folkedal, O., Imsland, A. & Oppedal, F. The effect of thermal acclimation on aerobic scope and critical swimming speed in Atlantic salmon, Salmo salar. J. Exp. Biol. 220, 2757–2764 (2017).
    Google Scholar 
    Ohlberger, J., Mehner, T., Staaks, G. & Hölker, F. Intraspecific temperature dependence of the scaling of metabolic rate with body mass in fishes and its ecological implications. Oikos 121, 245–251 (2012).Article 

    Google Scholar 
    Kunz, K. L. et al. New encounters in Arctic waters: A comparison of metabolism and performance of polar cod (Boreogadus saida) and Atlantic cod (Gadus morhua) under ocean acidification and warming. Polar Biol. 39, 1137–1153 (2016).Article 

    Google Scholar 
    Norin, T., Bailey, J. A. & Gamperl, A. K. Thermal biology and swimming performance of Atlantic cod (Gadus morhua) and haddock (Melanogrammus aeglefinus). PeerJ 7, e7784 (2019).Article 

    Google Scholar 
    Nowell, L. B. et al. Swimming energetics and thermal ecology of adult bonefish (Albula vulpes): A combined laboratory and field study in Eleuthera, The Bahamas. Environ. Biol. Fishes 98, 2133–2146 (2015).Article 

    Google Scholar 
    Pang, X., Yuan, X.-Z., Cao, Z.-D., Zhang, Y.-G. & Fu, S.-J. The effect of temperature on repeat swimming performance in juvenile qingbo (Spinibarbus sinensis). Fish Physiol. Biochem. 41, 19–29 (2015).Article 
    CAS 

    Google Scholar 
    Schwieterman, G. D. et al. Metabolic Rates and Hypoxia Tolerences of clearnose skate (Rostaraja eglanteria), summer flounder (Paralichthys dentatus), and thorny skate (Amblyraja radiata). Biology 8, 56 (2019).Article 
    CAS 

    Google Scholar 
    Xie, H. et al. Effects of acute temperature change and temperature acclimation on the respiratory metabolism of the snakehead. Turk. J. Fish. Aquat. Sci. 17, 535–542 (2017).Article 

    Google Scholar  More

  • in

    Building a living shoreline to help combat climate change

    I’m a conservation land manager at the Port of San Diego in California. My team and I aim to manage the tidelands around San Diego Bay, an area of more than 4,850 hectares, three-quarters of which is covered by water at high tide. At least 60% of the bay’s shoreline is ‘hardened’ — that is, it is edged with either a solid seawall or rip rap, piles of artificial boulders.To prevent erosion of the adjacent natural shoreline and restore wetlands, we’re participating in the San Diego Bay Native Oyster Living Shoreline project. As part of that, in December 2021, we placed 360 reef balls — depicted in this photograph from September this year — along 260 metres of shoreline to form the foundation of a native-oyster reef in the south bay. Here, I’m looking for oysters that have settled and are growing on the spheres.The reef balls are made out of ‘baycrete’, a concrete mixture made with local sand and the shells of farmed oysters. These attract wild oysters, which come to live there. We’re targeting the native Olympia oysters (Ostrea lurida), which can filter up to 190 litres of water per day. And sediment should accumulate behind the reef balls, encouraging the growth of eelgrass (Zostera marina). The grass is the foundation of the bay’s food chain.In a couple of years, native oysters will cover the reef balls, forming an artificial reef offshore. This reef will cause storm waves to break farther from the shoreline, protecting the adjacent salt marsh. Just inland from this area is a wetlands habitat refuge for the endangered California least tern (Sternula antillarum browni), and many birds are already hopping onto the reef balls and eating what’s living there.Living shorelines are an important part of sequestering carbon to combat climate change — both eelgrass and oysters store a lot of carbon. The reef balls are win–win–win. I often joke that we’re trying to save the planet one acre (0.4 hectares) at a time. More

  • in

    Mild shading promotes sesquiterpenoid synthesis and accumulation in Atractylodes lancea by regulating photosynthesis and phytohormones

    Mild shading facilitates sesquiterpenoid accumulation and growth in Atractylodes lancea rhizomeTo determine a concrete shading value for the production of high-quality and high-yielding AR, we examined the major compounds, including the sesquiterpenoids hinesol (Hin), β-eudesmol (Edu), and atractylone (Atl), and the polyacetylene atractylodin (Atd), as well as the biomass of AR at different growth stages (Fig. 1A–C) under various light intensities. The sum of these four volatile oils as the total volatile oil content was subsequently analyzed. The results revealed that the accumulation of volatile oils was significantly different (p  More

  • in

    High-resolution tracking of hyrax social interactions highlights nighttime drivers of animal sociality

    Siegel, J. M. Do all animals sleep? Trends Neurosci. 31, 208–213 (2008).Article 
    CAS 

    Google Scholar 
    Lima, S. L., Rattenborg, N. C., Lesku, J. A. & Amlaner, C. J. Sleeping under the risk of predation. Anim. Behav. 70, 723–736 (2005).Article 

    Google Scholar 
    Tougeron, K. & Abram, P. K. An Ecological Perspective on Sleep Disruption. Am. Nat. 190, 55–66 (2017).Article 

    Google Scholar 
    Lesku, J. A., Aulsebrook, A. E., Kelly, M. L. & Tisdale, R. K. Evolution of Sleep and Adaptive Sleeplessness. Handbook of Behavioral Neuroscience vol. 30 (Elsevier B.V., 2019).Smeltzer, E. A. et al. Social sleepers: The effects of social status on sleep in terrestrial mammals. Horm. Behav. 143, 105181 (2022).Article 
    CAS 

    Google Scholar 
    Chu, H. S., Oh, J. & Lee, K. The Relationship between Living Arrangements and Sleep Quality in Older Adults: Gender Differences. Int. J. Environ. Res. Public Health 19, 3893 (2022).Karamihalev, S., Flachskamm, C., Eren, N., Kimura, M. & Chen, A. Social context and dominance status contribute to sleep patterns and quality in groups of freely-moving mice. Sci. Rep. 9, 1–9 (2019).Article 
    CAS 

    Google Scholar 
    Capellini, I., Barton, R. A., McNamara, P., Preston, B. T. & Nunn, C. L. Phylogenetic analysis of the ecology and evolution of mammalian sleep. Evolution 62, 1764–1776 (2008).Article 

    Google Scholar 
    Ogawa, H., Idani, G., Moore, J., Pintea, L. & Hernandez-Aguilar, A. Sleeping Parties and nest distribution of chimpanzees in the Savanna woodland, Ugalla, Tanzania. Int. J. Primatol. 28, 1397–1412 (2007).Article 

    Google Scholar 
    Mulavwa, M. N. et al. Nest groups of wild bonobos at Wamba: Selection of vegetation and tree species and relationships between nest group size and party size. Am. J. Primatol. 72, 575–586 (2010).
    Google Scholar 
    Matsuda, I., Tuuga, A. & Higashi, S. Effects of water level on sleeping-site selection and inter-group association in proboscis monkeys: Why do they sleep alone inland on flooded days? Ecol. Res. 25, 475–482 (2010).Article 

    Google Scholar 
    Schreier, A. L. & Swedell, L. Ecology and sociality in a multilevel society: Ecological determinants of spatial cohesion in hamadryas baboons. Am. J. Phys. Anthropol. 148, 580–588 (2012).Article 

    Google Scholar 
    Kummer, H. & Kurt, F. Social units of free-living population of hamadryas baboons. Folia Primotol. 1, 4–19 (1963).Ogawa, H. & Takahashi, H. Triadic positions of Tibetan macaques huddling at a sleeping site. Int. J. Primatol. 24, 591–606 (2002).Article 

    Google Scholar 
    Snyder-Mackler, N., Beehner, J. C. & Bergman, T. J. Defining Higher Levels in the Multilevel Societies of Geladas (Theropithecus gelada). Int. J. Primatol. 33, 1054–1068 (2012).Article 

    Google Scholar 
    Mochida, K. & Nishikawa, M. Sleep duration is affected by social relationships among sleeping partners in wild Japanese macaques. Behav. Process. 103, 102–104 (2014).Article 

    Google Scholar 
    Di Bitetti, M. S., Vidal, E. M. L., Baldovino, M. C. & Benesovsky, V. Sleeping site preferences in tufted capuchin monkeys (Cebus apella nigritus). Am. J. Primatol. 50, 257 (2000).Article 

    Google Scholar 
    Takahashi, H. Huddling relationships in night sleeping groups among wild Japanese macaques in Kinkazan Island during winter. Primates 38, 57–68 (1997).Article 

    Google Scholar 
    Park, O., Barden, A. & Williams, E. Studies in Nocturnal Ecology, IX. Further Analysis of Activity of Panama Rain Forest Animals. Ecology 21, 122 (1940).Article 

    Google Scholar 
    Gaston, K. J. Nighttime ecology: The “nocturnal problem” revisited. Am. Nat. 193, 481–502 (2019).Article 

    Google Scholar 
    Börger, L. et al. Biologging Special Feature. J. Anim. Ecol. 89, 6–15 (2020).Article 

    Google Scholar 
    Krause, J. et al. Reality mining of animal social systems. Trends Ecol. Evol. 28, 541–551 (2013).Article 

    Google Scholar 
    Zeus, V. M., Puechmaille, S. J. & Kerth, G. Conspecific and heterospecific social groups affect each other’s resource use: a study on roost sharing among bat colonies. Anim. Behav. 123, 329–338 (2017).Article 

    Google Scholar 
    Wey, T. W., Burger, J. R., Ebensperger, L. A. & Hayes, L. D. Reproductive correlates of social network variation in plurally breeding degus (Octodon degus). Anim. Behav. 85, 1407–1414 (2013).Article 

    Google Scholar 
    Hirsch, B. T., Prange, S., Hauver, S. A. & Gehrt, S. D. Genetic relatedness does not predict racoon social network structure. Anim. Behav. 85, 463–470 (2013).Article 

    Google Scholar 
    Robitaille, A. L., Webber, Q. M. R., Turner, J. W. & Wal Eric, V. The problem and promise of scale in multilayer animal social networks. Curr. Zool. 67, 113–123 (2021).Article 

    Google Scholar 
    Smith, J. E. et al. Split between two worlds: Automated sensing reveals links between above- and belowground social networks in a free-living mammal. Philos. Trans. R. Soc. B Biol. Sci. 373, 20170249 (2018).Silk, M. J. et al. Seasonal variation in daily patterns of social contacts in the European badger Meles meles. Ecol. Evol. 7, 9006–9015 (2017).Article 

    Google Scholar 
    Gaynor, K. M., Hojnowski, C. E., Carter, N. H. & Brashares, J. S. The influence of human disturbance on wildlife nocturnality. Science 360, 1232–1235 (2018).Article 
    CAS 

    Google Scholar 
    Barry, R. E. & Mundy, P. J. Seasonal variation in the degree of heterospecific association of two syntopic hyraxes (Heterohyrax brucei and Procavia capensis) exhibiting synchronous parturition. Behav. Ecol. Sociobiol. 52, 177–181 (2002).Article 

    Google Scholar 
    Barocas, A., Ilany, A., Koren, L., Kam, M. & Geffen, E. Variance in centrality within rock hyrax social networks predicts adult longevity. PLoS ONE 6, 1–8 (2011).Article 

    Google Scholar 
    Ilany, A., Barocas, A., Koren, L., Kam, M. & Geffen, E. Structural balance in the social networks of a wild mammal. Anim. Behav. 85, 1397–1405 (2013).Article 

    Google Scholar 
    Gravett, N., Bhagwandin, A., Lyamin, O. I., Siegel, M. & Manger, P. R. Sleep in the Rock Hyrax, Procavia capensis. Brain Behav. Evol. 79, 155–169 (2012).Coe, M. J. Notes on the habits of the mount kenya hyrax (Procavia johnstoni mackinderi thomas). Proc. Zool. Soc. Lond. 138, 638–644 (1961).
    Google Scholar 
    Viblanc, V. A., Pasquaretta, C., Sueur, C., Boonstra, R. & Dobson, F. S. Aggression in Columbian ground squirrels: relationships with age, kinship, energy allocation, and fitness. Behav. Ecol. 27, arw098 (2016).Article 

    Google Scholar 
    Wolf, J. B. W., Mawdsley, D., Trillmich, F. & James, R. Social structure in a colonial mammal: unravelling hidden structural layers and their foundations by network analysis. Anim. Behav. 74, 1293–1302 (2007).Article 

    Google Scholar 
    Podgórski, T., Lusseau, D., Scandura, M., Sönnichsen, L. & Jȩdrzejewska, B. Long-lasting, kin-directed female interactions in a spatially structured wild boar social network. PLoS ONE 9, 1–11 (2014).Article 

    Google Scholar 
    Druce, D. J. et al. Scale-dependent foraging costs: Habitat use by rock hyraxes (Procavia capensis) determined using giving-up densities. Oikos 115, 513–525 (2006).Article 

    Google Scholar 
    Goll, Y. et al. Sex-associated and context-dependent leadership in the rock hyrax. iScience 104063 https://doi.org/10.1016/j.isci.2022.104063 (2022).Kelley, J. L., Morrell, L. J., Inskip, C., Krause, J. & Croft, D. P. Predation risk shapes social networks in fission-fusion populations. PLoS One 6, e24280 (2011).Article 
    CAS 

    Google Scholar 
    Brown, K. J. Seasonal variation in the thermal biology of the rock hyrax (Procavia capensis) (Document N° 10413/10124) [Master Dissertation, University of KwaZulu-Natal]. ResearchSpace Digital Library for UKZN scholarly research. http://hdl.handle.net/10413/10124.Bar Ziv, E. et al. Individual, social, and sexual niche traits affect copulation success in a polygynandrous mating system. Behav. Ecol. Sociobiol. 70, 901–912 (2016).Article 

    Google Scholar 
    McDonald, G. C., Spurgin, L. G., Fairfield, E. A., Richardson, D. S. & Pizzari, T. Differential female sociality is linked with the fine-scale structure of sexual interactions in replicate groups of red junglefowl, Gallus gallus. Proc. R. Soc. B Biol. Sci. 286, 20191734 (2019).Stanley, C. R., Liddiard Williams, H. & Preziosi, R. F. Female clustering in cockroach aggregations—A case of social niche construction? Ethology 124, 706–718 (2018).Article 

    Google Scholar 
    Pilastro, A., Benetton, S. & Bisazza, A. Female aggregation and male competition reduce costs of sexual harassment in the mosquitofish Gambusia holbrooki. Anim. Behav. 65, 1161–1167 (2003).Article 

    Google Scholar 
    Schoepf, I. & Schradin, C. Better off alone! Reproductive competition and ecological constraints determine sociality in the African striped mouse (Rhabdomys pumilio). J. Anim. Ecol. 81, 649–656 (2012).Article 

    Google Scholar 
    Brent, L. J. N., MacLarnon, A., Platt, M. L. & Semple, S. Seasonal changes in the structure of rhesus macaque social networks. Behav. Ecol. Sociobiol. 67, 349–359 (2013).Article 

    Google Scholar 
    Sundaresan, S. R., Fischhoff, I. R., Dushoff, J. & Rubenstein, D. I. Network metrics reveal differences in social organization between two fission-fusion species, Grevy’s zebra and onager. Oecologia 151, 140–149 (2007).Article 

    Google Scholar 
    Hasenjager, M. J. & Dugatkin, L. A. Fear of predation shapes social network structure and the acquisition of foraging information in guppy shoals. Proc. R. Soc. B Biol. Sci. 284, 20172020 (2017).Heathcote, R. J. P., Darden, S. K., Franks, D. W., Ramnarine, I. W. & Croft, D. P. Fear of predation drives stable and differentiated social relationships in guppies. Sci. Rep. 7, 1–10 (2017).Article 

    Google Scholar 
    Dunbar, R. I. M. Social structure as a strategy to mitigate the costs of group living: a comparison of gelada and guereza monkeys. Anim. Behav. 136, 53–64 (2018).Article 
    CAS 

    Google Scholar 
    Sutcliffe, A., Dunbar, R., Binder, J. & Arrow, H. Relationships and the social brain: Integrating psychological and evolutionary perspectives. Br. J. Psychol. 103, 149–168 (2012).Article 

    Google Scholar 
    Brown, M. R. Comparing the Fission-Fusion Dynamics of Spider Monkeys (Ateles geoffroyi) From Day to Night. https://doi.org/10.11575/PRISM/25371 (2014).Fanson, K. V., Fanson, B. G. & Brown, J. S. Using path analysis to explore vigilance behavior in the rock hyrax (Procavia capensis). J. Mammal. 92, 78–85 (2011).Article 

    Google Scholar 
    Santema, P. & Clutton-Brock, T. Meerkat helpers increase sentinel behaviour and bipedal vigilance in the presence of pups. Anim. Behav. 85, 655–661 (2013).Article 

    Google Scholar 
    Wright, J., Berg, E., De Kort, S. R., Khazin, V. & Maklakov, A. A. Cooperative sentinel behaviour in the Arabian babbler. Anim. Behav. 62, 973–979 (2001).Article 

    Google Scholar 
    Moscovice, L. R., Sueur, C. & Aureli, F. How socio-ecological factors influence the differentiation of social relationships: An integrated conceptual framework. Biol. Lett. 16, 20200384 (2020).Kotler, B. P., Brown, J. S. & Knight, M. H. Habitat and patch use by hyraxes: There’s no place like home? Ecol. Lett. 2, 82–88 (1999).Article 

    Google Scholar 
    Margolis, E. Dietary composition of the wolf Canis lupus in the Ein Gedi area according to analysis of their droppings (in Hebrew). In Proceedings of 45th Meeting of the Israel Zoological Society, Isr. J. Ecol. Evol. 55, 157–180 (2008).Firth, J. A. & Sheldon, B. C. Social carry-over effects underpin trans-seasonally linked structure in a wild bird population. Ecol. Lett. 19, 1324–1332 (2016).Article 

    Google Scholar 
    Olds, N. & Shoshani, J. Procavia capensis. Mammalian Species 171, 1–7 (2016).Fourie, L. J. & Perrin, M. R. Social behaviour and spatial relationships of the rock hyrax. South 17, 91–98 (1987).Montiglio, P.-O., Ferrari, C. & Réale, D. Social niche specialization under constraints: Personality, social interactions and environmental heterogeneity. Philos. Trans. R. Soc. B Biol. Sci. 368, 20120343 (2013).Article 

    Google Scholar 
    Dunbar, R. I. M. Time: a hidden constraint on the behavioural ecology of baboons. Behav. Ecol. Sociobiol. 31, 35–49 (1992).Article 

    Google Scholar 
    Dunbar, R. I. M., Korstjens, A. H. & Lehmann, J. Time as an ecological constraint. Biol. Rev. 84, 413–429 (2009).Article 
    CAS 

    Google Scholar 
    Zahavi, A. Arabian babbler. In Cooperative Breeding in Birds (eds. Staceyp, B. & Koenigw, D.) 103-130 (Cambridge University Press, 1990).Smith, J. E. et al. Greetings promote cooperation and reinforce social bonds among spotted hyenas. Anim. Behav. 81, 401–415 (2011).Article 

    Google Scholar 
    Aureli, F. & Schaffner, C. M. Aggression and conflict management at fusion in spider monkeys. Biol. Lett. 3, 147–149 (2007).Article 

    Google Scholar 
    Deag, J. M. The diurnal patterns of behaviour of the wild Barbary macaque Macaca sylvanus. J. Zool. 206, 403–413 (1985).Article 

    Google Scholar 
    Canteloup, C., Cera, M. B., Barrett, B. J. & van de Waal, E. Processing of novel food reveals payoff and rank-biased social learning in a wild primate. Sci. Rep. 11, 1–13 (2021).Article 

    Google Scholar 
    Dragić, N., Keynan, O. & Ilany, A. Multilayer social networks reveal the social complexity of a cooperatively breeding bird. iScience 24, 103336 (2021).Kulahci, I. G., Ghazanfar, A. A. & Rubenstein, D. I. Knowledgeable Lemurs Become More Central in Social Networks. Curr. Biol. 28, 1306–1310.e2 (2018).Article 
    CAS 

    Google Scholar 
    Schino, G. Grooming and agonistic support: A meta-analysis of primate reciprocal altruism. Behav. Ecol. 18, 115–120 (2007).Article 

    Google Scholar 
    Kutsukake, N. & Clutton-Brock, T. H. Social functions of allogrooming in cooperatively breeding meerkats. Anim. Behav. 72, 1059–1068 (2006).Article 

    Google Scholar 
    Schweinfurth, M. K., Stieger, B. & Taborsky, M. Experimental evidence for reciprocity in allogrooming among wild-type Norway rats. Sci. Rep. 7, 1–8 (2017).Article 
    CAS 

    Google Scholar 
    Nandini, S., Keerthipriya, P. & Vidya, T. N. C. Group size differences may mask underlying similarities in social structure: A comparison of female elephant societies. Behav. Ecol. 29, 145–159 (2018).Article 

    Google Scholar 
    Hamede, R. K., Bashford, J., McCallum, H. & Jones, M. Contact networks in a wild Tasmanian devil (Sarcophilus harrisii) population: using social network analysis to reveal seasonal variability in social behaviour and its implications for transmission of devil facial tumour disease. Ecol. Lett. 12, 1147–1157 (2009).Article 

    Google Scholar 
    Henkel, S., Heistermann, M. & Fischer, J. Infants as costly social tools in male Barbary macaque networks. Anim. Behav. 79, 1199–1204 (2010).Article 

    Google Scholar 
    Prehn, S. G. et al. Seasonal variation and stability across years in a social network of wild giraffe. Anim. Behav. 157, 95–104 (2019).Article 

    Google Scholar 
    Borgeaud, C., Sosa, S., Sueur, C. & Bshary, R. The influence of demographic variation on social network stability in wild vervet monkeys. Anim. Behav. 134, 155–165 (2017).Article 

    Google Scholar 
    Kerth, G., Perony, N. & Schweitzer, F. Bats are able to maintain long-term social relationships despite the high fission-fusion dynamics of their groups. Proc. R. Soc. B 278, 2761–2767 (2011).Article 

    Google Scholar 
    Silk, J. B. et al. The benefits of social capital: Close social bonds among female baboons enhance offspring survival. Proc. R. Soc. B Biol. Sci. 276, 3099–3104 (2009).Article 

    Google Scholar 
    Riehl, C. & Strong, M. J. Stable social relationships between unrelated females increase individual fitness in a cooperative bird. Proc. R. Soc. B Biol. Sci. 285, 20180130 (2018).Shizuka, D. & Johnson, A. E. How demographic processes shape animal social networks. Behav. Ecol. 31, 1–11 (2020).Article 

    Google Scholar 
    Sick, C. et al. Evidence for varying social strategies across the day in chacma baboons. Biol. Lett. 10, 3–6 (2014).Article 

    Google Scholar 
    Barrett, L., Peter Henzi, S. & Lusseau, D. Taking sociality seriously: The structure of multi-dimensional social networks as a source of information for individuals. Philos. Trans. R. Soc. B Biol. Sci. 367, 2108–2118 (2012).Article 

    Google Scholar 
    Henzi, S. P., Lusseau, D., Weingrill, T., Van Schaik, C. P. & Barrett, L. Cyclicity in the structure of female baboon social networks. Behav. Ecol. Sociobiol. 63, 1015–1021 (2009).Article 

    Google Scholar 
    Ripperger, S. P. & Carter, G. G. Social foraging in vampire bats is predicted by long-term cooperative relationships. PLoS Biol. 19, 1–17 (2021).Article 

    Google Scholar 
    Wittemyer, G., Douglas-Hamilton, I. & Getz, W. M. The socioecology of elephants: analysis of the processes creating multitiered social structures. Anim. Behav. 69, 1357–1371 (2005).Article 

    Google Scholar 
    Wittemyer, G., Getz, W. M., Vollrath, F. & Douglas-Hamilton, I. Social dominance, seasonal movements, and spatial segregation in African elephants: A contribution to conservation behavior. Behav. Ecol. Sociobiol. 61, 1919–1931 (2007).Article 

    Google Scholar 
    Gelardi, V., Fagot, J., Barrat, A. & Claidière, N. Detecting social (in)stability in primates from their temporal co-presence network. Anim. Behav. 157, 239–254 (2019).Article 

    Google Scholar 
    Hobson, E. A., Ferdinand, V., Kolchinsky, A. & Garland, J. Rethinking animal social complexity measures with the help of complex systems concepts. Anim. Behav. 155, 287–296 (2019).Article 

    Google Scholar 
    Kappeler, P. M. A framework for studying social complexity. Behav. Ecol. Sociobiol. 73, 13 (2019).Ballerini, M. et al. Interaction ruling animal collective behavior depends on topological rather than metric distance: Evidence from a field study. Proc. Natl Acad. Sci. USA 105, 1232–1237 (2008).Article 
    CAS 

    Google Scholar 
    Bonabeau, E., Theraulaz, G., Deneubourg, J.-L., Aron, S. & Camazine, S. Self-organization in social insects. Trends Ecol. Evol. 12, 188–193 (1997).Article 
    CAS 

    Google Scholar 
    Wickramasinghe, A. & Muthukumarana, S. Assessing the impact of the density and sparsity of the network on community detection using a Gaussian mixture random partition graph generator. Int. J. Inf. Technol. 14, 607–618 (2022).
    Google Scholar 
    Motalebi, N., Stevens, N. T. & Steiner, S. H. Hurdle Blockmodels for Sparse Network Modeling. Am. Stat. 75, 383–393 (2021).Article 

    Google Scholar 
    Gokcekus, S., Cole, E. F., Sheldon, B. C. & Firth, J. A. Exploring the causes and consequences of cooperative behaviour in wild animal populations using a social network approach. Biol. Rev. 96, 2355–2372 (2021).Article 

    Google Scholar 
    Koren, L., Mokady, O. & Geffen, E. Social status and cortisol levels in singing rock hyraxes. Horm. Behav. 54, 212–216 (2008).Article 
    CAS 

    Google Scholar 
    Boyland, N. K., James, R., Mlynski, D. T., Madden, J. R. & Croft, D. P. Spatial proximity loggers for recording animal social networks: Consequences of inter-logger variation in performance. Behav. Ecol. Sociobiol. 67, 1877–1890 (2013).Article 

    Google Scholar 
    Drewe, J. A. et al. Performance of proximity loggers in recording Intra- and Inter-species interactions: A laboratory and field-based validation study. PLoS ONE 7, e39068 (2012).Hoppitt, W. & Farine, D. Association Indices For Quantifying Social Relationships: How To Deal With Missing Observations Of Individuals Or Groups. Anim. Behav. 136, 227–238 (2018).Article 

    Google Scholar 
    Farine, D. R. Animal social network inference and permutations for ecologists in R using asnipe. Methods Ecol. Evol. 4, 1187–1194 (2013).Bejder, L., Fletcher, D. & Bräger, S. A method for testing association patterns of social animals. Anim. Behav. 56, 719–725 (1998).Article 
    CAS 

    Google Scholar 
    Kalinka, A. T. & Tomancak, P. linkcomm: An R package for the generation, visualization, and analysis of link communities in networks of arbitrary size and type. Bioinformatics 27, 2011–2012 (2011).Article 
    CAS 

    Google Scholar 
    R Core Team, R. R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. https://www.R-project.org/ (2020).Wild, F. lsa: Latent Semantic Analysis. R package version 0.73.2. https://CRAN.R-project.org/package=lsa (2020).Han, J., Kamber, M. & Pei, J. Getting to Know Your Data. An R Companion Third Ed. Fundam. Polit. Sci. Res. https://doi.org/10.1016/B978-0-12-381479-1.00002-2 (2021).Benjamini, Y. Controlling the false discovery rate – A practical and powerful approach to multiple testing. J. R. Stat. Soc. Ser. B 57, 289–300 (1995).
    Google Scholar 
    Csardi, G. & Nepusz, T. The igraph software package for complex network research. InterJournal, Complex Syst. 1695, 1–9 (2006).Dai, H., Leeder, J. S. & Cui, Y. A modified generalized fisher method for combining probabilities from dependent tests. Front. Genet. 20, 2–7 (2014).
    Google Scholar  More