More stories

  • in

    Inversions maintain differences between migratory phenotypes of a songbird

    The research in this study was performed in agreement with permission M45-14 issued by Malmö/Lund Ethical Committee for Animal Research, Sweden, which granted capture and blood sampling of wild birdsSamplesNine willow warblers, determined to be males (based on a wing length > 69 mm), were caught opportunistically with mist nets during the time of autumn migration in September 2016 at Krankesjön, 15 km East of Lund, Southern Sweden. While most of the individuals were phenotypically similar to willow warblers breeding in Southern Scandinavia, some were slightly larger and had a greyer plumage, which is more commonly seen in Northern Scandinavia12. The set of samples thus potentially contained willow warblers of each of the two major migratory phenotypes. Blood from each bird was collected through a puncture of the brachial vein and was stored in two vails containing SET buffer and 70% ethanol, respectively. An aliquot of the blood was used for DNA extraction with a phenol-chloroform protocol. From the extracted DNA, we genotyped the samples for two loci located on chromosomes 1 and 5, respectively (NBEA and FADS2)45,46, and for a bi-allelic marker within the divergent region on chromosome 3 (AFLP-ww1)47. Based on the genotyping results we selected two samples that were homozygous northern or homozygous southern for all three loci, respectively. We also included a sample from a chiffchaff Phylloscopus collybita (female) for de novo genome sequencing of a closely related outgroup species, as well as an additional willow warbler (DD81063, male) to confirm breakpoint differences with linked read sequencing. Both of these birds were opportunistically caught at the same site as above during autumn migration in 2019, and collection of blood followed the same approach as for the other birds.Optical mapsDNA from the northern and southern willow warbler was extracted from blood stored in ethanol using a Plug Lysis protocol (v.30026D; Bionano Genomics, CA, USA). The blood was first separated from the ethanol through gentle centrifugation and embedded in molten 2% agarose plugs (DNA plug kit; Bio-Rad, CA, USA). The solidified plugs were submerged in Lysis Buffer solution (Bionano Genomics) and 66.8 µl per ml Buffer Puregene Proteinase K (Qiagen,MD, USA) for 2 h at 50 °C. The plugs were subsequently washed in 1× Wash buffer (Bio-Rad DNA plug kit) followed by TE buffer. In the following step, the plugs were treated with RNase (Qiagen, 20 µl in 1 ml TE buffer) for 1 h at 37 °C, followed by another washing step using the same buffers as in the previous step. Next, the plugs were melted for 2 min at 70 °C and treated with GELase (Epicenter, WI, USA) for 45 min at 43 °C. The DNA was then purified from digested agarose using drop dialysis against TE buffer on a 0.1 µm dialysis membrane (MF-Millipore, Merck KGaA, Germany) for 2.5 h.Optical maps for each of the two samples were produced using Bionano Genomic’s commercial Irys system48. BspQ1 was determined to be the most suitable nicking enzyme after using the software LabelDensityCalculator v.1.3.0 and Knickers v.1.5.5 to analyze a previous short-read assembly13. Bionano Genomic’s IrysPrep Labeling-NLRS protocol (v.30024) was used for the NLRS reaction. For this step, DNA was treated with Nt.BspQ1 (NEB, MA, USA) to create single-stranded nicks in a molecule-specific pattern. These were then labeled with Bionano Genomic’s (CA, USA) labeling mix (NLRS kit), aided by Taq Polymerase (NEB), and repaired using Bionano Genomics’s repair mix (NLRS kit), in the presence of Thermopol Rxn buffer, NAD+, and Taq DNA Ligase (NEB). Finally, the DNA backbone was stained using DNA stain from Bionano Genomics’s NLRS kit. Each sample was then loaded on two IrysChips (Bionano Genomics) each, and the DNA with stained BspQ1 nicks was visualized using an Irys instrument, following Bionano Genomics’s Irys user guide (v.30047). This resulted in 200 and 182 Gb of data for the northern and southern sample, respectively.Genome maps were assembled de novo using Bionano Genomic’s in house software IrysView v.2.5.1, with noise parameter set to “autonoise” and using a human arguments xml file. The genome map was then further refined by re-assembling all data, but using the first assembly version as a reference. The final assemblies were both 1.3 Gb in total size, with an average coverage of 92.3 and 96.4×, and N50 of 0.93 and 0.95 Mb, for the northern and southern sample, respectively.Linked read sequencingFor the southern sample and sample DD81063, DNA for chromium sequencing (10× Genomics, CA, USA) was extracted from blood stored in SET buffer using a MagAttract HMW DNAkit (Qiagen) at Scilifelab, Stockholm, Sweden. For the northern sample the extraction for bionano optical maps was used. The libraries of the northern and southern sample were each sequenced on a separate lane of a HiSeqX (Illumina, CA, USA) and the DD81063 sample was sequenced on a NovaSeq6000 (Illumina). For all samples sequencing was performed using a 2 × 150 bp setup.Northern willow warbler de novo assemblyLibrary preparation for long read sequencing was done on DNA previously extracted for the optical map and followed Pacific Bioscience’s (CA, USA) standard protocol for 10–20 kb libraries. No shearing was performed prior to the library construction, but the library was size selected using the BluePippin pulse field size selection system (Sage Science, MA, USA), with a size cut-off >25 kb. The library was sequenced on eight SMRT cells on a Sequel platform (Pacific Biosciences). The sequencing yielded 63.66 Gbp of data comprised of 4,690,365 subreads with a mean length of 13,573 bp (range: 50–170,531 bp).The Pacbio reads were assembled de novo in HGAP449 in the SMRT Link package with default settings except for specifying an expected genome size of 1.2 Gbp and setting the polishing algorithm to “Arrow”. We ran Falcon unzip50 on the assembly to obtain partially phased primary contigs and fully phased haplotigs. Within the software, Arrow was used to polish the assembly using reads assigned to each haplotype. We evaluated two unzipped assemblies based on 30× or 40× coverage of seed reads in the preassembly step in HGAP4. A lower coverage threshold will lead to longer reads in the initial assembly step, which may increase the contiguity of the assembly, but will on the other hand, limit the number of reads that can be used in the phasing and polishing step. Although the unzipped assemblies were very similar, the 40× version was chosen for downstream analyses as it was slightly more contiguous and contained a higher number of single-copy bird orthologues as determined by BUSCO version 3.0.251.The assembly was further polished with Pilon 1.2252 with Illumina chromium reads from the same sample. The Illumina reads were mapped to the assembly using bwa version 0.7.17-r118853 and duplicated reads were marked using picardtools 2.10.3 (http://broadinstitute.github.io/picard). Pilon was run by only correcting indels and in total the software made 1,043,827 insertions and 275,457 deletions, respectively, of which the vast majority (94%) were single basepair changes. The Illumina polishing had a pronounced effect on the number of single-copy bird orthologues that could be detected in the primary contigs (Supplementary Table 1).For further assembly steps, we extracted the Illumina-polished primary Pacbio contigs (N = 2737, N50 of 2.1 Mb and a length of 1.29 Gb). These contigs showed an unexpectedly high level of duplicated single-copy orthologues (7.4%), which suggested partial or complete overlap between some contigs. As a first step to reduce the redundancy and increase the contiguity of the assembly, we hybridized the primary contigs to the optical map of the same sample using bionano solve version 3.2.2 (BioNano Genomics) with default settings except for specifying aggressive scaffolding parameters. The hybrid scaffolding resulted in 19 cuts to the bionano maps and 259 cuts to the Pacbio contigs and created 363 super-scaffolds. Most of the gaps between the contigs in the super-scaffolds were estimated to be negative (i.e., some overlap between sequences). However, in the hybrid assembly, sequences on either side of these gaps were not collapsed and thus formed false segmental duplications. To remedy this problem we extracted 304 sets of overlapping contigs (“supercontigs”) and used GAP5 in the staden package 2.0.0.b1154 to find potential joins between the contig ends. Using this approach, we merged contigs at 558 (87%) of the putative overlaps. The mean alignment length in the overlaps was 111 kb (range: 0.259–661 kb) with a mean sequence divergence of 3.28% (range: 0.31–15.55%). The highest divergence was caused by the presence of large indels. By trimming off one or both ends of the contigs at the gaps (mean 23 kb, range: 0.6–60 kb), we were able to close 23 further gaps. For the remainder of gaps, GAP5 failed to find potential joins between contigs or the ends supposed to be joined were considered to have too high divergence. The new assembly, including supercontigs consisted of 2401 contigs with an N50 of 6.5 Mb and had a considerably lower amount of duplicated single-copy genes (4.6% vs 7.4%).To further reduce the redundancy, we used the purge haplotig pipeline55 (downloaded 2019-02-15) to remove contigs that could be mapped over most of their length to larger contigs and that showed limited diploid coverage. We first estimated coverage by mapping the Pacbio subreads used for the de novo assembly with minimap2 version2.13-r86056 using default settings for Pacbio reads (-x map-pb). To minimize the loss of repetitive sequences that could be separated and scaffolded by the bionano optical map, we used the first bionano hybrid assembly (363 superscaffolds and 1500 cut and unscaffolded contigs) as a reference for mapping. From the mapped data we detected a clear haploid and diploid peak and set a threshold of diploid coverage above 34× and below 85×. Any scaffold where less than 80% of its positions had diploid coverage was considered a putative haplotig and was mapped to other scaffolds using minimap2 within the software. We removed 1209 scaffolds (mean size: 107,655 bp, range: 598–495,788 bp) with a coverage to the best hit of at least 70% (mean: 97.4%). Using this approach, we specifically excluded contigs that could not be incorporated in superscaffolds. However, we also removed three contigs that each entirely made up short superscaffolds that could be uniquely assigned to larger superscaffolds and that had a high degree of haploid coverage. At this stage, we also removed five additional contigs shorter than 1000 bp that were the result of cutting the assembly with the bionano optical map. This led to an assembly with 1187 contigs, a length of 1.1 Gbp and a N50 of 7.9 Mb. The filtered assembly showed a large reduction in single-copy orthologue bird genes (1.3 vs 4.6%).To provide an intermediate level of scaffolding to the optical map, we mapped the 10× chromium reads of the same sample to the assembly using bwa and used arcs version 1.0.557 and LINKS version 1.8.658 for scaffolding. Arcs was run with default settings except for enabling gap size estimation (–dist_est) and LINKS was run by setting the number of supporting links to at least 5 (-l = 5) and the maximum link ratio between the two best contig pairs to 0.3 (-a = 0.3). The scaffolding resulted in 739 scaffolds with a N50 of 16.4 Mb and a length 1.12 Gb.As a final scaffolding step, we hybridized the 10× chromium-Pacbio scaffolds to the bionano optical map using the same settings as before. The hybrid scaffolding made 23 cuts to the optical map, 122 cuts to the scaffolds and resulted in 497 scaffolds with an N50 of 16.8 Mb. Two contigs representing the divergent region on chromosome 1 had been scaffolded together by arcs but were separated and not re-scaffolded with other sequences in the bionano hybrid assembly. Since the mismatched end of the optical map was short, located at a large gap, and the gene order is the same as seen in other bird genomes, we decided to keep the scaffold generated by arcs.For this round of hybrid scaffolding, there were 52 gaps that were estimated to be negative. Using the same approach as when creating supercontigs, we were able to close 10 of these gaps. We additionally closed gaps using PBJelly59 from PBSuite 15.8.24 with default settings except for specifying –spanOnly –capturedOnly”. The software filled 97 gaps, extended one end of 12 gaps, extended both ends of 18 gaps and overfilled 28 gaps (extended both ends but detected no overlap despite the extension is larger than the predicted gap).We further checked for potential misjoins between scaffolds that originate from different chromosomes. To this end, we used SatsumaSynteny 2.060 to produce whole-genome alignments between the assembly and the genomes of chicken (version GRCg6a) and zebra finch (version taeGut3.2.4), both downloaded from Ensembl (www.ensembl.org). Using this approach, we detected a scaffold that showed good alignments to both chromosomes 10 and 23 in both of the other species. We considered this join unlikely and decided to split the scaffold.Next, we performed a second round of polishing with the 10× chromium Illumina data from the same sample. For this round, since we had fewer than 500 scaffolds, we used the longranger 2.1.14 align pipeline61 to map reads in a barcode-aware way. Pilon was then run with the same settings as before and resulted in the correction of 417,032 indels, of which 78.7% were single-basepair changes. The second round of polishing considerably increased the number of single-copy bird orthologues that could be identified in the assembly (Supplementary Table 1).The mitochondrial genome was not found in the original Pacbio genome assembly. We obtained this genome by adding the complete mitochondrial sequence from a previous short-read assembly13. We then used bwa to map the 10× chromium reads from the northern sample to the assembly and extracted alignments on the mitochondrial sequence. Next, freebayes was used with a haploid setting to detect differences present in the aligned reads. The raw variant file was filtered with vcftools for sites with a quality less than 30 and for two intervals with excessive read coverage (possibly reads from unassembled NUMTs). The filtered variant file contained 11 substitutions and three indels, and was used with bcftools version 1.1462 to create a new mitochondrial reference.For the extraction and removal of sequences in the different assembly steps we used kentUtils 370 (https://github.com/ucscGenomeBrowser/kent). Summary statistics for each assembly (e.g., N50) were calculated using the assemblathon_stats.pl script63.Southern willow warbler and chiffchaff de novo assembliesThe southern willow warbler and the chiffchaff were each sequenced on two lanes on a Sequel II (Pacific Biosciences) using a high-fidelity (HiFi) setup. Sequencing libraries for the southern willow warbler was prepared from a previous extraction used for optical maps (see above), whereas for the chiffchaff, DNA was extracted from blood using a Nanobind extraction kit (Circulomics, MD, USA). The southern willow sample yielded 2,576,876 HiFi reads with a mean length 19,303 bp and representing 49.7 Gbp. The chiffchaff sample yielded 2,612,165 HiFi reads with a mean length of 19,829 bp and representing 51.8 Gbp.The HiFi reads were assembled de novo using hifiasm version 0.15.5-r35064 with default settings and primary contigs were selected for downstream analyses. For the chiffchaff hifiasm assembly, we removed the first 6 Mb part of a contig overlapping with another contig and removed a short interval at the end of a contig containing adaptor sequences. For the southern willow warbler, the primary contigs (N = 540, Supplementary Table 1) were hybridized to the optical map of the same sample using the same pipeline as for the northern sample. Although we had access to chromium data from the same sample, we did not include it to perform an intermediate scaffolding step (as we did for the northern willow warbler assembly) because the long-read assembly was already highly contiguous. The hybridization step made 39 cuts to the contigs and 20 cuts to the optical maps, resulting in an assembly with 111 superscaffolds and 439 non-scaffolded contigs. We decided to ignore an optical map-supported fusion of contigs that mapped to separate chromosomes in other bird species, as this fusion was made in a large repetitive region. We further excluded a 45 bp sequence resulting from the hybrid assembly cutting and masked four short intervals containing adaptor sequences. The assembly of the mitochondrion in the southern assembly followed the same pipeline as used for the northern assembly (see above). In this case, 10 substitutions and two indels were added to the mitochondrial sequence from the previous short-read assembly based on alignments of linked reads from the southern sample.Repeat annotationWe used Repeatmodeler version 1.0.865 for de novo identification of repeats in the southern assembly. The repeats detected by repeatmodeler were combined with 1,023 bird-specific repeats into a custom library. Next, we used repeatmasker version 4.0.766 with the custom library and by using a more sensitive search (-s flag) to annotate repeats in the genome. Bedtools v2.29.267, together with the annotated repeats, was used to create a softmasked version of the southern assembly, which was used in the gene annotation step. The same repeat library was also used to annotate repeats in the de novo assembly of the northern sample. For the chiffchaff assembly we used the same annotation approach as for the southern willow warbler, but included a species-specific library generated with repeatmodeler, and also included a tandem-repeat associated sequence associated with the divergent regions on chromosomes 1 and 3 from the willow warbler library. Intervals with tandem repeats in divergent regions were also analyzed with tandem repeats finder version 4.0.968 using default settings except for specifying a maximum period size of 2000 bp.Duplicated intervals within divergent scaffolds were identified with Minimap2 and subsequently aligned with EMBOSS Stretcher 6.6.0 (https://www.ebi.ac.uk/Tools/psa/emboss_stretcher/).RNA sequencingWe used total RNA extracted from whole brain from six samples used in an earlier study quantifying differential expression in migratory and breeding willow warblers69 (Supplementary Table 3). The quality of the RNA was checked with a Bioanalyzer version 2100 (Agilent, CA, USA). All of the extractions had a RNA Integrity Number (RIN) of at least > 7.10. RNA libraries for sequencing were prepared using a TruSeq Stranded mRNA Sample prep kit with 96 dual indexes (Illumina) according to the instructions of the manufacturer with the exception of automating the protocols using an NGS workstation (Agilent) and using purification steps as described in Lundin et al70. and Borgström et al71. The raw RNA data was trimmed using cutadapt version 1.872 within Trim Galore version 0.4.0 (https://github.com/FelixKrueger/TrimGalore) with default settings.We used Stringtie version 1.3.373 to create transcripts from the RNAseq data. These transcripts were not used directly in the generation of gene models, but used in the manual curation step as potential alternative transcripts. For the software, we first mapped the reads with Hisat2 version 2.1.074 using default settings for stranded sequence libraries and downstream transcript analyses.Gene annotationWe used Augustus version 3.2.375 to create gene models using hints provided from RNAseq data and protein data from other bird species. For the RNAseq data, we mapped the trimmed reads to the assembly using STAR version 2.7.9a76. Accessory scripts in the Augustus package were used to filter the alignments for paired and uniquely mapped reads and for extracting intron hints. We additionally generated coverage wig files for each strand from the filtered alignment file using the software stranded-coverage (https://github.com/pmenzel/stranded-coverage) and used these as input for the august wig2hints.pl to generate exonpart hints.For homology evidence, we downloaded a set of bird proteins from NCBI (https://www.ncbi.nlm.nih.gov/). This data set included 49,673 proteins from chicken, 41,214 proteins from zebra finch and 38,619 proteins from great tit. We also downloaded an additional dataset from Uniprot (www.uniprot.org) that consisted of 3175 manually reviewed bird proteins and 204 and 12,263 bird proteins that were not manually reviewed but supported by protein or transcript data, respectively. The protein data was mapped to the genome using exonerate version 2.4.077. We used the script align2hints.pl from braker 2.1.678 to generate CDSpart, intron, start and stop hints from the data.Augustus was run with species-specific parameters (see training Augustus below) and with default settings except for specifying “softmasking=true”, “–alternatives-from-evidence=true”, “–UTR = on”, “–gff3=on” and “–allow_hinted_splicesites=atac”. In the extrinsic configuration file, we changed the malus for introns from 0.34 to 0.001, which increases the penalty for predicted introns that are not supported by the extrinsic data (RNAseq and protein hints). The prediction resulted in 28,491 genes and 35,389 transcripts.The Augustus-derived gene models were assigned names based on overlap with synteny-transferred zebra finch genes. For this purpose, we used SatsumaSynteny with default settings to obtain whole-genome alignments between our assembly and the zebra finch genome version bTaeGut1.4.pri79. Based on the alignment, we used kraken80 (downloaded 2020-04-14) to transfer the zebra finch genome annotations (NCBI Release 106) to the willow warbler assembly. We then extracted the CDS from the Augustus gene models and the kraken genes and used bedtools intersect to quantify the overlap. The gene models were also searched against the longest translation of each of the chicken, zebra finch and great tit Parus major genes used as evidence for the gene prediction step and against 86,131 swissprot vertebrate proteins using blastp 2.5.0+81 with an E value threshold of 1e−5. Gene models that were not annotated through synteny were assigned a gene name based on the blast results. Protein domains in the gene models were annotated with interproscan v 5.30–69.082. To reduce the number of false positive predictions we removed 5697 genes that were not supported by synteny to zebra finch genes, showed no significant similarity to vertebrate proteins or did not contain any annotated protein domains.We used Webapollo 2.6.583 to manually curate gene models in the previously identified divergent chromosome regions and in other regions where differences were present. In the curation step, we specifically validated the support for the coding sequence and the UTR and also removed genes that were likely to be pseudogenes based on a truncated coding sequence compared to homologous genes in other vertebrates, had no support from synteny in other bird species and/or that were located in repeat-rich regions.Training AugustusWe used a previous repeat-masked short-read assembly13 and the trimmed RNAseq data used in this study to obtain species-specific parameters for Augustus. The RNAseq data was assembled into transcripts using Trinity version 2.0.284 to create a de novo and a genome-guided assembly that together were comprised of 1,929,396 transcripts. The genome-guided transcript assembly was based on RNAseq mapped to the genome using GSNAP version 2016-07-1185 with default settings. We used PASA version 2.0.286 to create high-quality transcripts, which were imported into Webapollo. To assess the completeness of the transcripts, we compared them to synteny-transferred models from the chicken genome using Kraken. We selected 1249 transcripts that appeared complete, were not overlapping with other genes and showed less than 80% amino acid similarity to another gene in the training set. From this set, we excluded 21 genes that were giving initial training errors, which gave us a training set of 1228 genes. This gene set was randomly split into 1028 training genes and 200 genes used for testing. For training, we used the optimize_augustus.pl script with default settings except for the flag –UTR = on.Whole-genome resequencing and variant callingWe used the whole-genome resequencing data from nine samples of each migratory phenotype provided in Lundberg et al13. and sequenced an additional two high-coverage samples from each migratory phenotype (Supplementary Table 4). Sequencing libraries for the new samples were prepared with a TruSeq DNA PCR-Free kit (Illumina) with a targeted insert size of 670 bp or with a Truseq DNA nano (Illumina) with a targeted insert size of 350 bp. All of the new samples were sequenced on a HiSeqX (Illumina). The raw reads were trimmed with trimmomatic 0.3687 with the parameters “ILLUMINACLIP:TruSeq3-PE-2.fa:2:30:10 LEADING:3 TRAILING:3 SLIDINGWINDOW:4:15 MINLEN:30”.Quality-trimmed reads were mapped to the southern assembly using bwa mem with default settings except for specifying -M flag to ensure compatibility with the downstream duplicate removal steps and converted into binary alignment map (bam) files using samtools. For samples sequenced across multiple lanes, reads from each lane were mapped independently and the resulting bam files were merged with samtools. Read duplicates were removed with the markduplicates tool provided in picardtools.From the aligned whole-genome resequencing data set, we called variants with freebayes v1.1.0 using default settings and parallelizing the analyses of separate scaffolds using GNU parallel88. Vcflib version 2017-04-0489 was used to filter the raw set of variants for sites with quality score >30 and for alternate alleles that were supported by at least one read on each strand (SAF  > 0 & SAR  > 0) and had at least one read balanced to the right and the left (RPL  > 0 & RPR  > 0). Next, we used vcftools 0.1.1690 to filter genotypes with a coverage of at least 5x and removed sites a maximum of four genotypes missing in each of the populations. The variants were also filtered for collapsed repeats by removing sites with a mean coverage of more than twice the median mean coverage (30×). We next used vcflib to decompose haplotype calls and complex alleles into indels and SNPs and removed any variants that were overlapping with annotated repeats. This gave us a final of 51 million variants of which 45 million were bi-allelic SNPs. We used vcftools to calculate FST91 for each variant and for bi-allelic SNPs in non-overlapping windows of 10 kb. As many rare variants segregate in the willow warbler populations, which may downwardly bias differentiation estimates92, we focused on variants with a minor allele frequency of at least 0.1.Coverage for each resequenced sample was calculated in non-overlapping 1 kb windows using bedtools and only included properly paired reads with a mapping quality of at least 1. The raw coverage values for each sample were normalized by its median coverage across all windows.Structural variant callingWe used a combination of delly 0.9.193 and GraphTyper 2.7.494 to call structural variants in the resequenced samples. To identify a set of high confidence variants, we first mapped the long reads from the northern willow warbler to the southern assembly using minimap 2.22-r110156 with default settings for Pacbio reads and from the alignments called variants using delly. Next, GraphTyper was used to genotype the resequenced samples for the delly variants in the scaffolds containing the divergent chromosome regions. The raw set of variants were filtered to contain only sites with a “PASS” flag and, for each variant, the aggregated genotype, which is the genotype model out of breakpoint alignments and coverage that has the highest genotyping quality, was chosen for downstream analyses. Genetic differentiation (FST) was calculated in vcftools and variants with FST ≥ 0.7 between homozygotes in each divergent chromosome region were extracted and checked for overlap with genes and gene features using bedtools. To get more reliable differentiation estimates, we only included sites where at least 80% of the southern and northern homozygotes had genotypes.Inversion genotypes for resequenced samplesThe resequenced samples were assigned a genotype of southern and northern haplotypes for each of the divergent regions based on a multidimensional scaling (MDS)-based clustering in invclust95 of SNP array genotypes in Lundberg et al.13. To obtain genotypes of the SNPs included on the array in the resequenced samples, we mapped the SNP array probe sequences to the northern assembly using gmap and from the alignments extracted the positions of the focal SNPs. Next, we used freebayes to genotype the resequenced samples for these positions and plink version 1.996 to combine the genotypes with the genotypes from the SNP array. In the genotyping step, we also included mapped 10× chromium libraries for the northern and southern reference samples and the additional willow warbler sample. From the combined dataset, we extracted genotypes for SNPs located in each of the divergent regions and used invclust to assign each sample a genotype of inverted and non-inverted haplotypes. The inverted and non-inverted haplotypes were recoded as southern or northern haplotypes based on their frequency in each subspecies.Breakpoint analysesWe used MUMmer 4.0.0rc197 to align the genomes of the southern and northern willow warblers, and the southern willow warbler genome to the genomes of the chiffchaff, zebra finch (3.2.4) and collared flycatcher FicAlb (1.5)98.To provide further evidence of breakpoints, we mapped the 10× chromium reads of each sample to both the northern and the southern assembly and called structural variants using the longranger wgs pipeline. For the southern genome, we selected the 499 largest scaffolds and concatenated the rest into a single scaffold to make it compatible with the software. We also checked for differences in linked read molecule coverage between the samples. For this purpose, the raw reads of each sample were first processed with longranger basic for quality trimming and barcode processing. The trimmed reads were mapped to the assemblies using bwa mem using a -C flag to extract the barcode information of each read and alignments converted into bam files using samtools. To estimate coverage of barcodes, we first used the tigmint-molecule script from tigmint 1.1.299 to obtain positional information of barcodes (molecules) in each divergent region. The software was run with default settings except for only using reads with a mapping quality of at least 1 and only to report molecules that were estimated to be at least 10 kb. We next used bedtools to count the number of overlapping molecules in 1 kb windows.We explored differences between optical maps by using the runSV.py script in bionano solve with the southern optical map as a query and the northern assembly as target and the reciprocal analysis with the northern optical map as a query and the southern assembly as a target. We also used the bionano solve hybrid assembly pipeline to visualize differences between the optical maps and the genome assemblies at breakpoint regions.Functional annotation of differencesWe used bedtools to quantify the distance between breakpoint intervals and annotated genes. To provide a functional annotation of the SNPs and short indels, we selected variants that showed a FST ≥ 0.7 between southern and northern homozygotes for each of the region and used these as input to Snpeff 5.0.0e100 together with the annotation and reference genome. We used Snpsift 5.0.0e101 to select variants that were predicted to have a moderate to high effect on genes. Gene ontology terms for the genes were extracted from orthologous genes in other bird genomes in ensembl (www.ensembl.org) or through domain searches of the proteins with interproscan.Age estimation and demographic analyses of divergent regionsIn order to estimate the timing of the inversion events, we used high-coverage resequencing data from two southern samples, two northern samples and, as an outgroup, one dusky warbler Phylloscopus fuscatus (Supplementary Table 4). The willow warbler samples were chosen so that they were either homozygous southern or northern for all of three divergent regions. The dusky warbler library was prepared using a TruSeq Nano DNA library prep kit for Neoprep (Illumina) according to the instructions of the manufacturer and sequenced on a HiSeq X (Illumina). Quality-trimming of the raw reads and mapping of the trimmed reads to the northern reference genome followed the same approach as used for the willow warbler resequencing samples (see above).Variants were called using freebayes and the raw set of variants were filtered using gIMble’s preprocess module (v0.6.0). Sample-specific callable sites were identified using gIMble preprocess and were defined as those with a minimum coverage of 8× and a maximum of 0.75 standard deviations above the mean coverage. Genic and repetitive regions of the genome were removed from the callable sites in order to limit downstream analyses to intergenic regions.Summary statistics of genetic variation (π and dxy) within the divergent regions were calculated using gIMble. Following this, net divergence (da) between northern and southern samples was calculated as dnorth–south − (πnorth + πsouth)/2. To convert the net divergence into years we used the germline mutation rate (4.6 × 10−9) estimated in the collared flycatcher21. Relative node depth (RND) using the dusky warbler (DW) as an outgroup was calculated as dnorth–south/(dDW-north + dDW-south)/2. For each divergent region, a blockwise site frequency spectrum (bSFS) was generated with gIMble using blocks of 64 bp in length. This length refers to the number of callable sites within a block, while the physical length of blocks was allowed to vary due to missing data but was limited to 128 bp. Downstream analyses that relied on a bSFS used a kmax of 2, meaning that only marginal probabilities were calculated for mutation counts >2. The composite likelihood (CL) of a model, given the bSFS of one of the divergent regions, was optimized using the Nelder-Mead algorithm with the maximum number of iterations set to 1000. Within the software we evaluated three different population models. The first model was a strict isolation model (SI), with parameters ancestral effective population size, effective population sizes for southern and northern willow warblers and divergence time. The second model was an isolation with migration model (IM1) that also included a migration rate from northern to southern samples, and the third model (IM2) instead had a migration rate from southern to northern willow warblers.Simulations were carried out by msprime 0.7.4102 through gIMble. The recombination rates used for these simulations were chromosome-specific estimates from a high-density recombination map of the collared flycatcher98 and were 2.04, 1.95, and 2.63 cM/Mb for chromosomes 1, 3, and 5, respectively. A total of 100 replicates were simulated for the optimized SI parameters of each region. These simulated bSFSs were then optimized under both an SI model as well as the best fitting IM model for that region. The improvement in CL between these models was used as a null distribution for testing whether improvements in CL observed for the real data were greater than expected given a history of no migration. For each parameter, we calculated 95% CI as Maximum Composite Likelihood (MCL) estimate ± 1.96 * standard deviation of simulations (Supplementary Table 7). As a result, our estimates of uncertainty are affected by the recombination rates that we assumed for simulations. We also used the results of simulations to quantify the potential bias in MCL estimates due to intra-block recombination (Supplementary Table 7). However, we did not attempt to correct for this bias as it is relatively small (e.g., the MCL divergence times are estimated to be biased upwards by 7, 24, and 10%) and our estimation of the bias itself is largely dependent on the recombination rates we assumed.MSMC224 was used to explore genome-wide changes in Ne through time. As input to the software, we used the callable intergenic bed file and filtered vcf file mentioned above, with the addition of further filtering the bed file to only include autosomal scaffolds ≥500 kb and excluding the divergent regions. The input files for MSMC2, i.e., an unphased set of heterozygous sites for each sample, were generated using the generate_multihetsep.py script from msmc-tools. MSMC2 was run with a starting ρ/μ of 1 for 30 expectation-maximum iterations. For both the demographic modeling and MSMC2, we used the collared flycatcher germline mutation rate21 and a generation time of 1.7 years11 to convert divergence times into years.To infer the effects of demographic events and selection, we also calculated several genetic summary statistics. To this end, we first imputed missing genotypes and inferred haplotypes for the filtered set of variants using beagle version 5.4103. From the full set of samples, we selected 10 and seven samples that were homozygous southern or northern for the three divergent regions, respectively, as determined from the MDS analysis (see above), and extracted bi-allelic SNPs. To identify ancestral and derived alleles, we extracted genotypes for the focal SNP positions from the aligned chiffchaff and dusky warblers reads using bcftools 1.1462 with the mpileup command. As a conservative approach, we considered any site with the presence of both the reference and alternate allele as heterozygous (regardless of their frequencies) and only included sites where the coverage was at least one-third of the mean coverage among all sites for each outgroup species. We next used a customized script to extract the sites from the original vcf files, and, if necessary, switch the reference and alternate allele and swap the genotypes accordingly. With the polarized genotype data, we used PopGenome 2.7.5104 to calculate Fay and Wu’s H and vcftools to get counts for the derived allele. We further used selscan 1.3.0105 to calculate XP-nsl106 between the southern and northern samples, Sweepfinder2107 to calculate a composite likelihood ratio (CLR) between a model where a selective sweep has had an effect on the allele frequency and a model based on the genome-wide allele frequency spectrum and used vcftools to calculate nucleotide diversity, Tajima’s D and linkage disequilibrium (D’).The use of the southern assembly as a reference could potentially lead to a mapping bias for reads from southern samples, particularly in regions of higher divergence between the subspecies. This, in turn, could have an effect on genetic summary statistics and demographic modeling estimates. To explore the effect of reference bias, we therefore also mapped the resequencing data to the northern assembly, performed variant calling and calculated nucleotide diversity and Tajima’s D in 10 kb windows. For the northern assembly, we also used the same demographic modeling as used for the southern assembly. Contrasting average genetic summary statistics and demographic parameter estimates, we found negligible differences between the two genome assemblies (Supplementary Table 10).Reporting summaryFurther information on research design is available in the Nature Portfolio Reporting Summary linked to this article. More

  • in

    Inferring genetic structure when there is little: population genetics versus genomics of the threatened bat Miniopterus schreibersii across Europe

    Charlesworth, B. & Charlesworth, D. Population genetics from 1966 to 2016. Heredity 118, 2–9 (2017).CAS 

    Google Scholar 
    Orsini, L., Vanoverbeke, J., Swillen, I., Mergeay, J. & Meester, L. Drivers of population genetic differentiation in the wild: Isolation by dispersal limitation, isolation by adaptation and isolation by colonization. Mol. Ecol. 22, 5983–5999 (2013).
    Google Scholar 
    Vendrami, D. L. J. et al. RAD sequencing resolves fine-scale population structure in a benthic invertebrate: Implications for understanding phenotypic plasticity. R. Soc. Open Sci. 4, 160548 (2017).ADS 

    Google Scholar 
    Dufresnes, C., Rodrigues, N. & Savary, R. Slow and steady wins the race: Contrasted phylogeographic signatures in two Alpine amphibians. Integr. Zool. 17, 181–190 (2021).
    Google Scholar 
    Frankham, R. Genetics and extinction. Biol. Conserv. 126, 131–140 (2005).
    Google Scholar 
    Schwartz, M. K., Luikart, G. & Waples, R. S. Genetic monitoring as a promising tool for conservation and management. Trends in Ecol. Evol. 22, 25–33 (2007).
    Google Scholar 
    Ottewell, K. M., Bickerton, D. C., Byrne, M. & Lowe, A. J. Bridging the gap: A genetic assessment framework for population-level threatened plant conservation prioritization and decision-making. Divers. Distrib. 22, 174–188 (2016).
    Google Scholar 
    Frankham, R., Bradshaw, C. J. A. & Brook, B. W. Genetics in conservation management: Revised recommendations for the 50/500 rules, red list criteria and population viability analyses. Biol. Conserv. 170, 56–63 (2014).
    Google Scholar 
    Hohenlohe, P. A., Funk, C. W. & Rajora, O. P. Population genomics for wildlife conservation and management. Mol. Ecol. 30, 62–82 (2020).
    Google Scholar 
    Angelone, S. & Holderegger, R. Population genetics suggests effectiveness of habitat connectivity measures for the European tree frog in Switzerland. J. Appl. Ecol. 46, 879–887 (2009).
    Google Scholar 
    Griffiths, S. M., Taylor-Cox, E. D., Behringer, D. C., Butler, M. J. IV. & Preziosi, R. F. Using genetics to inform restoration and predict resilience in declining populations of a keystone marine sponge. Biodivers. Conserv. 29, 1383–1410 (2020).
    Google Scholar 
    Moritz, C. Conservation units and translocations: Strategies for conserving evolutionary processes. Hereditas 130, 217–228 (1999).
    Google Scholar 
    Bohonak, A. J. Dispersal, gene flow, and population structure. Q. Rev. Biol. 74, 21–45 (1999).CAS 

    Google Scholar 
    Arguedas, N. & Parker, P. G. Seasonal migration and genetic population structure in house wrens. Condor 102, 517–528 (2000).
    Google Scholar 
    Quillfeldt, P. et al. Does genetic structure reflect differences in non-breeding movements? A case study in small, highly mobile seabirds. BMC Evol. Biol. 17, 160 (2017).
    Google Scholar 
    Charlesworth, B., Sniegowski, P. & Stephan, W. The evolutionary dynamics of repetitive DNA in eukaryotes. Nature 371, 215–220 (1994).ADS 
    CAS 

    Google Scholar 
    Schlötterer, C. Evolutionary dynamics of microsatellite DNA. Chromosoma 109, 365–371 (2000).
    Google Scholar 
    Ellegren, H. Microsatellites: Simple sequences with complex evolution. Nat. Rev. Genet. 5, 435–445 (2004).CAS 

    Google Scholar 
    Hodel, R. G. J. et al. The report of my death was an exaggeration: A review for researchers using microsatellites in the 21st century. Appl. Plant Sci. 4, 1600025 (2016).
    Google Scholar 
    Dufresnes, C. & Litvinchuk, S. N. Diversity, distribution and molecular species delimitation in frogs and toads from the Eastern Palearctic. Zool. J. Linn. Soc. 195, 695–760 (2022).
    Google Scholar 
    Galtier, N., Nabholz, B., Glémin, S. & Hurst, G. D. D. Mitochondrial DNA as a marker of molecular diversity: A reappraisal. Mol. Evol. 18, 4541–4550 (2009).CAS 

    Google Scholar 
    Zink, R. M. & Barrowclough, G. Mitochondrial DNA under siege in avian phylogeography. Mol. Ecol. 17, 2107–2121 (2008).CAS 

    Google Scholar 
    Toews, D. P. L. & Brelsford, A. The biogeography of mitochondrial and nuclear discordance in animals. Mol. Ecol. 21, 3907–3930 (2012).CAS 

    Google Scholar 
    Bonnet, T., Leblois, R., Rousset, F. & Crochet, P.-A. A reassessment of explanations for discordant introgressions of mitochondrial and nuclear genomes. Evolution 71, 2140–2218 (2017).
    Google Scholar 
    Davey, J. W. & Blaxter, M. L. RADSeq: Next-generation population genetics. Brief Funct. Genomics 9, 416–423 (2010).CAS 

    Google Scholar 
    Lexer, C. et al. ‘Next generation’ biogeography: Towards understanding the drivers of species diversification and persistence. J. Biogeogr. 40, 1013–1022 (2013).
    Google Scholar 
    Baird, N. A. et al. Rapid SNP discovery and genetic mapping using sequenced RAD markers. PLoS ONE 3, e3376 (2008).ADS 

    Google Scholar 
    Peterson, B. K., Weber, J. N., Kay, E. H., Fisher, H. S. & Hoekstra, H. E. Double Digest RADseq: An inexpensive method for de novo SNP discovery and genotyping in model and non-model species. PLoS ONE 7, e37135 (2012).ADS 
    CAS 

    Google Scholar 
    Dufresnes, C. et al. Phylogeography of a cryptic speciation continuum in Eurasian spadefoot toads (Pelobates). Mol. Ecol. 28, 3257–3270 (2019).
    Google Scholar 
    Sunde, J., Yıldırım, Y., Tibblin, P. & Forsman, A. Comparing the performance of microsatellites and RADseq in population genetic studies: Analysis of data for pike (Esox Lucius) and a synthesis of previous studies. Front. Genet. 11, 218 (2020).
    Google Scholar 
    Moussy, C. et al. Migration and dispersal patterns of bats and their influence on genetic structure. Mammal Rev. 43, 183–195 (2013).
    Google Scholar 
    Berthier, P., Excoffier, L. & Ruedi, M. Recurrent replacement of mtDNA and cryptic hybridization between two sibling bat species Myotis myotis and Myotis blythii. Proc. R. Soc. B: Biol. Sci. 273, 3101–3109 (2007).
    Google Scholar 
    Wright, P. G. R. et al. Hydrogen isotopes reveal evidence of migration of Miniopterus schreibersii in Europe. BMC Ecol. 20, 52 (2020).CAS 

    Google Scholar 
    Schnetter, W. Beringungsergebnisse an der Langflügelfledermaus (Miniopterus schreibersi Kühl) im Kaiserstuhl. Bonn. Zool. Beitr. 11, 150–165 (1960).
    Google Scholar 
    Rodrigues, L. Miniopterus schreibersii. In The Atlas of European Mammals (eds Mitchell-Jones, A. J. et al.) 154–155 (Academic Press, 1999).
    Google Scholar 
    Rodrigues, L., Ramos Pereira, M. J., Rainho, A. & Palmeirim, J. M. Behavioral determinants of gene flow in the bat Miniopterus schreibersii. Behav. Ecol. Sociobiol. 64, 835–843 (2010).
    Google Scholar 
    Rodrigues, L. & Palmeirim, J. M. Migratory behaviour of Miniopterus schreibersii (Chiroptera): When, where, and why do cave bats migrate in a Mediterranean region?. J. Zool. 274, 116–125 (2008).
    Google Scholar 
    Ramos Pereira, M. J., Salgueiro, P., Rodrigues, L., Coelho, M. M. & Palmeirim, J. M. Population structure of a cave-dwelling bat, Miniopterus schreibersii: Does it reflect history and social organization?. J. Hered. 100, 533–544 (2009).
    Google Scholar 
    Bilgin, R. et al. Circum-Mediterranean phylogeography of a bat coupled with past environmental niche modeling: A new paradigm for the recolonization of Europe?. Mol. Phylogenet. Evol. 99, 323–336 (2016).
    Google Scholar 
    Gürün, K. et al. A continent-scale study of the social structure and phylogeography of the bent-wing bat, Miniopterus schreibersii (Mammalia: Chiroptera), using new microsatellite data. J. Mammal. 100, 1865–1878 (2019).
    Google Scholar 
    Gazaryan, S., Bücs, S., Çoraman, E. Miniopterus schreibersii (errata version published in 2021). The IUCN Red List of Threatened Species 2020: e.T81633057A195856522 (2020).Miller-Butterworth, C. M., Jacobs, D. S. & Harley, E. H. Isolation and characterization of highly polymorphic microsatellite loci in Schreibers’ long-fingered bat, Miniopterus schreibersii (Chiroptera: Vespertilionidae). Mol. Ecol. Notes 2, 139–141 (2002).CAS 

    Google Scholar 
    Wood, R., Weyeneth, N. & Appleton, B. Development and characterisation of 20 microsatellite loci isolated from the large bent-wing bat, Miniopterus schreibersii (Chiroptera: Miniopteridae) and their cross-taxa utility in the family Miniopteridae. Mol. Ecol. Resour. 11, 675–685 (2011).
    Google Scholar 
    Witsenburg, F. et al. How a haemosporidian parasite of bats gets around: The genetic structure of a parasite, vector and host compared. Mol. Ecol. 24, 926–940 (2015).CAS 

    Google Scholar 
    Van Oosterhout, C., Hutchinson, W. F., Wills, D. P. M. & Shipley, P. micro-checker: Software for identifying and correcting genotyping errors in microsatellite data. Mol. Ecol. Notes 4, 535–538 (2004).
    Google Scholar 
    Parchman, T. L. et al. Genome wide association mapping of an adaptive trait in lodgepole pine. Mol. Ecol. 21, 2991–3005 (2012).CAS 

    Google Scholar 
    Catchen, J. M., Amores, A., Hohenlohe, P., Cresko, W. & Postlethwait, J. H. Stacks: Building and genotyping loci de novo from short-read sequences. G3 1, 171–182 (2011).CAS 

    Google Scholar 
    Weir, B. S. & Cockerham, C. C. Estimating F-statistics for the analyses of population structure. Evolution 38, 1358–1370 (1984).CAS 

    Google Scholar 
    Goudet, J. hierfstat, a package for r to compute and test hierarchical F-statistics. Mol. Ecol. Notes 5, 184–186 (2005).
    Google Scholar 
    Frankham, R., Ballou, J. D. & Briscoe, D. A. A Primer of Conservation Genetics (Cambridge University Press, 2004).
    Google Scholar 
    Weir, B. S. & Goudet, J. A unified characterization of population structure and relatedness. Genetics 206, 2085–2103 (2017).
    Google Scholar 
    Mantel, N. A. The detection of disease clustering and a generalized regression approach. Cancer Res. 27, 209–220 (1967).CAS 

    Google Scholar 
    Wright, S. Isolation by distance. Genetics 28, 114–138 (1943).CAS 

    Google Scholar 
    Cavalli-Sforza, L. L. & Edwards, A. W. F. Phylogenetic analysis: Model and estimation procedures. Am. J. Hum. Genet. 19, 233–257 (1967).CAS 

    Google Scholar 
    Paradis, E. & Schliep, K. ape 5.0: An environment for modern phylogenetics and evolutionary analyses in R. Bioinformatics 35, 526–528 (2019).CAS 

    Google Scholar 
    Goudet, J., Perrin, N. & Waser, P. Tests for sex-biased dispersal using bi-parentally inherited genetic markers. Mol. Ecol. 11, 1103–1114 (2002).CAS 

    Google Scholar 
    Frichot, E. & François, O. lea: An r package for landscape and ecological association studies. Methods Ecol. Evol. 6, 925–929 (2015).
    Google Scholar 
    Yannic, G. et al. High connectivity in a long-lived High-Arctic seabird, the ivory gull Pagophila eburnea. Polar Biol. 39, 221–236 (2016).
    Google Scholar 
    Cumer, T. et al. Landscape and climatic variations of the Quaternary shaped multiple secondary contacts among barn owls (Tyto alba) of the Western Palearctic. Mol. Biol. Evol. 39, msab343 (2022).CAS 

    Google Scholar 
    Boston, E. S. M., Montgomery, W. I., Hynes, R. & Prodöhl, P. A. New insights on postglacial colonization in western Europe: The phylogeography of the Leisler’s bat (Nyctalus leisleri). Proc. R. Soc. B: Biol. Sci. 282, 20142605 (2015).
    Google Scholar 
    Razgour, O. et al. The shaping of genetic variation in edge-of-range populations under past and future climate change. Ecol. Lett. 16, 1258–1266 (2013).
    Google Scholar 
    Petit, E., Balloux, F. & Goudet, J. Sex-biased dispersal in a migratory bat: A characterization using sex-specific demographic parameters. Evolution 55, 635–640 (2001).CAS 

    Google Scholar 
    Moussy, C. et al. Population genetic structure of serotine bats (Eptesicus serotinus) across Europe and implications for the potential spread of bat rabies (European bat lyssavirus EBLV-1). Heredity 115, 83–92 (2015).CAS 

    Google Scholar 
    Rossiter, S. J., Benda, P., Dietz, C., Zhang, S. & Jones, G. Rangewide phylogeography in the greater horseshoe bat inferred from microsatellites: Implications for population history, taxonomy and conservation. Mol. Ecol. 16, 4699–4714 (2007).CAS 

    Google Scholar 
    Dool, S. E. et al. Phylogeography and postglacial recolonization of Europe by Rhinolophus hipposideros: Evidence from multiple genetic markers. Mol. Ecol. 22, 4055–4070 (2013).CAS 

    Google Scholar 
    Kerth, G. et al. Communally breeding Bechstein’s bats have a stable social system that is independent from the postglacial history and location of the populations. Mol. Ecol. 17, 2368–2381 (2008).CAS 

    Google Scholar 
    Garrick, R. C., Banusiewicz, J. D., Burgess, S., Hyseni, C. & Symula, R. E. Extending phylogeography to account for lineage fusion. J. Biogeogr. 46, 268–278 (2019).
    Google Scholar 
    Burri, R. et al. The genetic basis of color-related local adaptation in a ring-like colonization around the Mediterranean. Evolution 70, 140–153 (2016).
    Google Scholar 
    Taberlet, P., Fumagalli, L., Wust-Saucy, A.-G. & Cosson, J.-F. Comparative phylogeography and postglacial colonization routes in Europe. Mol. Ecol. 7, 453–464 (1998).CAS 

    Google Scholar 
    Hewitt, G. M. Post-glacial re-colonization of European biota. Biol. J. Linn. Soc. 68, 87–112 (1999).
    Google Scholar 
    Gómez, A. & Lunt, D. H. Refugia within refugia: Patterns of phylogeographic concordance in the Iberian Peninsula. In Phylogeography of Southern European Refugia (eds Weiss, S. & Ferrand, N.) 155–188 (Springer, 2007).
    Google Scholar 
    Vonhof, M. J., Russell, A. L. & Miller-Butterworth, M. Range-wide genetic analysis of little brown bat (Myotis lucifugus) populations: Estimating the risk of spread of white-nose syndrome. PLoS ONE 10, e0128713 (2015).
    Google Scholar 
    Auteri, G. G. & Knowles, L. L. Decimated little brown bats show potential for adaptive change. Sci. Rep. 10, 3023 (2020).ADS 
    CAS 

    Google Scholar 
    Gignoux-Wolfsohn, S. A. et al. Genomic signatures of selection in bats surviving white-nose syndrome. Mol. Ecol. 30, 5643–5657 (2021).
    Google Scholar 
    Rivers, N. M., Butlin, R. K. & Altringham, J. D. Autumn swarming behaviour of Natterer’s bats in the UK: Population size, catchment area and dispersal. Biol. Conserv. 127, 215–226 (2006).
    Google Scholar 
    Reis, N. R., Fregonezi, M. N., Peracchi, A. L. & Rossaneis, B. K. Metapopulation in bats of Southern Brazil. Braz. J. Biol. 72, 605–609 (2012).CAS 

    Google Scholar 
    Humphrey, S. R. & Oli, M. K. Population dynamics and site fidelity of the cave bat, Myotis velifer, Oklahoma. J. Mammal. 96, 946–956 (2015).
    Google Scholar 
    Jeffries, D. L. et al. Comparing RADseq and microsatellites to infer complex phylogeographic patterns, an empirical perspective in the Crucian carp, Carassius carassius. L. Mol. Ecol. 25, 2997–3018 (2016).
    Google Scholar 
    Hodel, R. G. J. et al. Adding loci improves phylogeographic resolution in red mangroves despite increased missing data: Comparing microsatellites and RAD-Seq and investigating loci filtering. Sci. Rep. 7, 17598 (2017).ADS 

    Google Scholar 
    Lemopoulos, A. et al. Comparing RADseq and microsatellites for estimating genetic diversity and relatedness—Implications for brown trout conservation. Ecol. Evol. 9, 2106–2120 (2019).
    Google Scholar 
    Zimmerman, S. J., Aldridge, C. L. & Oyler-McCance, S. J. An empirical comparison of population genetic analyses using microsatellite and SNP data for a species of conservation concern. BMC Genom. 21, 382 (2020).CAS 

    Google Scholar 
    Hale, M. L., Burg, T. M. & Steeves, T. E. Sampling for microsatellite-based population genetic studies: 25 to 30 individuals per population is enough to accurately estimate allele frequencies. PLoS ONE 7, e45170 (2012).ADS 
    CAS 

    Google Scholar 
    Quetglas, J., Gonzalez, F. & Paz, O. Estudian la extraña mortandad de miles de murcielago de cuevas. Quercus 203, 50 (2003).
    Google Scholar 
    Negredo, A. et al. Discovery of an ebolavirus-like filovirus in Europe. PLoS Pathog. 7, e1002304 (2011).CAS 

    Google Scholar 
    Reed, D. H. & Frankham, R. Correlation between fitness and genetic diversity. Conserv. Biol. 17, 230–237 (2003).
    Google Scholar 
    Alcalde, J. T., Artácoz, A. & Meijide, F. Recuperación de la colonia de Miniopterus schreibersii de la cueva de Cueva de Ágreda (Soria). Barbastella 5, 32–35 (2012).
    Google Scholar 
    Kemenesi, G. et al. Re-emergence of Lloviu virus in Miniopterus schreibersii bats, Hungary, 2016. Emerg. Microbes Infect. 7, 66 (2018).
    Google Scholar 
    Kemenesi, et al. Isolation of infectious Lloviu virus from Schreiber’s bats in Hungary. Nat. Commun. 13, 1706 (2022).ADS 
    CAS 

    Google Scholar 
    Stoffel, C. et al. Genetic consequences of population expansions and contractions in the common hippopotamus (Hippopotamus amphibius) since the late Pleistocene. Mol. Ecol. 24, 2507–2520 (2015).
    Google Scholar  More

  • in

    Green roofs and pollinators, useful green spots for some wild bee species (Hymenoptera: Anthophila), but not so much for hoverflies (Diptera: Syrphidae)

    Seto, K. C., Güneralp, B. & Hutyra, L. R. Global forecasts of urban expansion to 2030 and direct impacts on biodiversity and carbon pools. Proc. Natl. Acad. Sci. USA 109, 16083–16088. https://doi.org/10.1073/pnas.1211658109 (2012).Article 
    ADS 

    Google Scholar 
    Faeth, S. H., Bang, C. & Saari, S. Urban biodiversity: Patterns and mechanisms. Ann. N. Y. Acad. Sci. 1223, 69–81. https://doi.org/10.1111/j.1749-6632.2010.05925.x (2011).Article 
    ADS 

    Google Scholar 
    Elmqvist, T., Zipperer, W. & Güneralp, B. Urbanisation, habitat loss, biodiversity decline: Solution pathways to break the cycle. In Routledge Handbook of Urbanisation and Global Environmental Change (eds Seta, K. et al.) 139–151 (Routledge, 2016).
    Google Scholar 
    Dirzo, R. et al. Defaunation in the Anthropocene. Science 345, 401–406 (2014).Article 
    ADS 
    CAS 

    Google Scholar 
    Hallmann, C. A. et al. More than 75 percent decline over 27 years in total flying insect biomass in protected areas. PLoS One 12, e0185809. https://doi.org/10.1371/journal.pone.0185809 (2017).Article 
    CAS 

    Google Scholar 
    Wagner, D., Grames, E. M., Forister, M. L., Berenbaum, M. R. & Stopak, D. Insect decline in the Anthropocene: Death by a thousand cuts. Biological sciences 118, e2023989118. https://doi.org/10.1073/pnas.2023989118 (2021).Article 
    CAS 

    Google Scholar 
    Goulson, D., Nicholls, E., Botias, C. & Rotheray, E. L. Bee declines driven by combined stress from parasites, pesticides, and lack of flowers. Science 347, 6229. https://doi.org/10.1126/science.1255957 (2015).Article 
    CAS 

    Google Scholar 
    Ollerton, J. (2021) Pollinators & pollination: nature and society. Pelagic publishing.IPBES (2016). The assessment report of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services on pollinators, pollination and food production. potts, S.G., Imperatriz-Fonseca, V.L and Ngo, H.T. (eds). Secretariat of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services, Bonn, Germany. 552 pages.Mallinger, R. E. & Gratton, C. Species richness of wild bees, but not the use of managed honeybees, increases fruit set of a pollinator dependent crop. J. Appl. Ecol. 52, 323–330 (2015).Article 

    Google Scholar 
    Kremen, C., Williams, N. M. & Thorp, R. W. Crop pollination from native bees at risk from agricultural intensification. Proc. Natl. Acad. Sci. U.S.A. 99, 16812–16816 (2002).Article 
    ADS 
    CAS 

    Google Scholar 
    Winfree, R., Fox, J. W., Williams, N. M., Reilly, J. R. & Cariveau, D. P. Abundance of common species, not species richness, drives delivery of a real-world ecosystem service. Ecol. Lett. 18, 626–635 (2015).Article 

    Google Scholar 
    Soroye, P., Newbold, T. & Kerr, J. Climate change contributes to widespread declines among bumble bees across continents. Science 367, 685–688 (2020).Article 
    ADS 
    CAS 

    Google Scholar 
    Matteson, K. C., Ascher, J. S. & Langellotto, G. A. Bee richness and abundance in New York City urban gardens. Ann. Entomol. Soc. Am. 101(1), 140–150. https://doi.org/10.1603/0013-8746(2008)101[140:BRAAIN]2.0.CO;2 (2008).Article 

    Google Scholar 
    Carré, G. et al. Landscape context and habitat type as drivers of bee diversity in European annual crops. Agr. Ecosyst. Environ. 133(1–2), 40–47. https://doi.org/10.1016/j.agee.2009.05.001 (2009).Article 

    Google Scholar 
    Goulson, D., Lye, G. C. & Darvill, B. Decline and conservation of bumble bees. Ann. Rev. Entomol. 53, 191–208. https://doi.org/10.1146/annurev.ento.53.103106.093454 (2008).Article 
    CAS 

    Google Scholar 
    Bates, A. J. et al. Changing bee and hoverfly pollinator assemblages along an urban-rural gradient. PLoS One 6(8), e23459. https://doi.org/10.1371/journal.pone.0023459 (2011).Article 
    ADS 
    CAS 

    Google Scholar 
    Deguines, N., Julliard, R., De Flores, M. & Fontaine, C. Functional homogenization of flower visitor communities with urbanisation. Ecol. Evol. 6(7), 1967–1976. https://doi.org/10.1002/ece3.2009 (2016).Article 

    Google Scholar 
    Larsson, M. Higher pollinator effectiveness by specialist than generalist flower-visitors of unspecialized Knautia arvensis (Dipsacaceae). Oecologia 146(3), 394–403. https://doi.org/10.1007/s00442-005-0217-y (2005).Article 
    ADS 

    Google Scholar 
    Pataki, D. E. et al. Coupling biogeochemical cycles in urban environments: Ecosystem services, green solutions, and misconceptions. Front. Ecol. Environ. 9, 27–36. https://doi.org/10.1890/090220 (2011).Article 

    Google Scholar 
    Mentens, J., Raes, D. & Hermy, M. Green roofs as a tool for solving rainwater runoff problems in the urbanized 21st century?. Landscape Urban Plann. 77, 217–226. https://doi.org/10.1016/j.landurbplan.2005.02.010 (2006).Article 

    Google Scholar 
    Oberndorfer, E. et al. Green roofs as urban ecosystems: Ecological structures, functions and services. Bioscience 57, 823–834. https://doi.org/10.1641/B571005 (2007).Article 

    Google Scholar 
    Braaker, S., Ghazoul, J., Obrist, M. K. & Moretti, M. Habitat connectivity shapes urban arthropod communities: The key role of green roofs. Ecology 95, 1010–1021. https://doi.org/10.1890/13-0705.1 (2014).Article 
    CAS 

    Google Scholar 
    Colla, S. R., Willis, E. & Packer, I. Can green roofs provide habitat for urban bees (Hymenoptera: Apidae)?. Cities and the Environment 2(1), 1–12 (2009).Article 

    Google Scholar 
    Tonietto, R., Fant, J., Ascher, J., Ellis, K. & Larkin, D. A comparison of bee communities of Chicago green roofs, parks and prairies. Landsc. Urban Plan. 103, 102–108 (2011).Article 

    Google Scholar 
    Ksiazek, K., Fant, J. & Skogen, K. An asssement of pollen limitation on Chicago green roofs. Landsc. Urban Plan. 107, 401–408 (2012).Article 

    Google Scholar 
    MacIvor, J. S. Building height matters: Nesting activity of bees and wasps on vegetated roofs. Israel J. Ecol. Evol. 62, 88–96. https://doi.org/10.1080/15659801.2015.1052635 (2015).Article 

    Google Scholar 
    Kratschmer, S., Kriechbaum, M. & Pachinger, B. Buzzing on top: Linking wild bee diversity, abundance and traits with green roof qualities. Urban Ecosyst. 21, 429–441 (2018).Article 

    Google Scholar 
    MacIvor, J. S., Ruttan, R. & Salehi, B. Exotics on exotics: Pollen analysis of urban bees visiting Sedum on a green roof. Urban Ecosyst. 18, 419–430 (2014).Article 

    Google Scholar 
    Matteson, K. C. & Langellotto, G. A. Determinates of inner city butterfly and bee species richness. Urban Ecosyst. 13, 333–347. https://doi.org/10.1007/s11252-010-0122-y (2010).Article 

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

    Google Scholar 
    Baldock, K.C.R, et al. (2015) Where is the UK’s pollinator biodiversity? The importance of urban areas for flower-visiting insects. Proc. R. Soc. B. https://doi.org/10.1098/rspb.2014.2849Theodorou, P. et al. Urban fragmentation leads to lower floral diversity, with knock-on impacts on bee biodiversity. Sci. Rep. 10, 21756. https://doi.org/10.1038/s41598-020-78736-x (2020).Article 
    ADS 
    CAS 

    Google Scholar 
    Lowenstein, D.M., Matteson, K.C., Xiao, I., Silva, A.M. and Minor, E.S (2014) Humans, bees, and pollination services in the city: The case of Chicago, IL (USA). Biodiversity Conservation 1–18. https://doi.org/10.1007/s10531-014-0752-0Winfree, R., Bartomeus, I. & Cariveau, D. Native pollinators in anthropogenic habitats. Annu. Rev. Ecol. Evol. Syst. 42, 1–22 (2011).Article 

    Google Scholar 
    Cariveau, D. P. & Winfree, R. Causes of variation in wild bee responses to anthropogenic drivers. Curr. Opin. Insect. Sci. 10, 104–109. https://doi.org/10.1016/j.cois.2015.05.004 (2015).Article 

    Google Scholar 
    Baldock, K. C. R. et al. A systems approach reveals urban pollinator hotspots and conservation opportunities. Nat. Ecol. Evol. 3, 363–373. https://doi.org/10.1038/s41559-018-0769-y (2019).Article 

    Google Scholar 
    Li, W. C. & Yeung, K. K. A. A comprehensive study of green roof performance from environmental perspective. Int. J. Sustain. Built Environ. 3, 127–134 (2021).Article 

    Google Scholar 
    Turner, M., Baker, W. L., Peterson, C. J. & Peet, R. K. Factors influencing succession: Lessons from large, infrequent natural disturbances. Ecosystems 1, 511–523. https://doi.org/10.1007/s100219900047 (1998).Article 

    Google Scholar 
    Molineux, C. J., Connop, S. P. & Gange, A. C. Manipulating soil microbial communities in extensive green roof substrates. Sci. Total Environ. 493, 632–638. https://doi.org/10.1016/j.scitotenv.2014.06.045 (2014).Article 
    ADS 
    CAS 

    Google Scholar 
    Macivor, S. & Ksiazek, K. Invertebrates on green roofs. Ecol. Stud. Anal. Synthes. 223, 333–355. https://doi.org/10.1007/978-3-319-14983-7_14 (2015).Article 

    Google Scholar 
    Madre, F., Vergnes, A., Machon, N. & Clergeau, P. A comparison of 3 types of green roof as habitats for arthropods. Ecol. Eng. 57, 109–117. https://doi.org/10.1016/j.ecoleng.2013.04.029 (2013).Article 

    Google Scholar 
    Lee, L. H. & Lin, J. C. Green roof performance towards good habitat for butterflies in the compact city. Int. J. Biol. 7, 103. https://doi.org/10.5539/ijb.v7n2p103 (2015).Article 
    CAS 

    Google Scholar 
    Preston, F. W. The canonical distribution of commonness and rarity: Part I. Ecology 43(2), 185–215. https://doi.org/10.2307/1931976 (1962).Article 

    Google Scholar 
    Orford, K. A., Murray, P. J., Vaughan, I. P. & Memmott, J. Modest enhancements to conventional grassland diversity improve the provision of pollination services. J. Appl. Ecol. 53, 906–915. https://doi.org/10.1111/1365-2664.12608 (2016).Article 

    Google Scholar 
    Brenneisen, S. The Natural Roof (NADA): Research Project Report on the Use of Extensive Green Roofs by Wild Bees (University of Wädenswil, 2005).
    Google Scholar 
    Jacobs, J., Berg, M., Beenaerts, N. & Artois, T. Biodiversity of Collembola on green roofs: A case study of three cities in Belgium. Ecol. Eng. 177, 106572. https://doi.org/10.1016/j.ecoleng.2022.106572 (2022).Article 

    Google Scholar 
    McKinney, M.L., Sisco, N.D. (2018) Systematic variation in roof spontaneous vegetation: residential “low rise” versus commercial “high rise” buildings. Urban Nature SI, 73–88.Rotheray, G.E., & Gilbert, S.F. (2011) The natural history of hoverflies. Tresaith, UK: Forrest TextBenvenuti, S. Wildflower green roofs for urban landscaping, ecological sustainability and biodiversity. Landsc. Urban Plan. 124, 151–161. https://doi.org/10.1016/j.landurbplan.2014.01.004 (2014).Article 

    Google Scholar 
    Schneider, F. Beitrag zur Kenntnis der Generationsverhaltnisse und Diapause rauberischer Schwebfliegen (Syrphldae, Dipt.). Mittl. Schweiz Ent Ges 21, 249–285 (1948).
    Google Scholar 
    Rader, R., Edwards, W., Westcott, D. A., Cunningham, S. A. & Howlett, B. G. Pollen transport differs among bees and flies in a human-modified landscape. Divers. Distrib. 17, 519–529. https://doi.org/10.1111/j.1472-4642.2011.00757.x (2011).Article 

    Google Scholar 
    Burgio, G. & Sommaggio, D. Syrphids as landscape bioindicators in Italian agroecosystems. Agr. Ecosyst. Environ. 120, 416–422 (2007).Article 

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

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

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

    Google Scholar 
    Schönrogge, K. et al. Host propagation permits extreme local adaptation in a social parasite of ants. Ecol. Lett. 9, 1032–1040 (2006).Article 

    Google Scholar 
    Schweiger, O. et al. Functional richness of local hoverfly communities (Diptera, Syrphidae) in response to land use across temperate Europe. Oikos 116, 461–472 (2007).Article 

    Google Scholar 
    KMI: Koninklijk Meteorologisch Instituut (2022) Analyse van het jaar 2020 en 2021. Available from https://www.meteobelgie.be/klimatologie/waarnemingen-en-analyses/jaar-2020/2274-jaa-2020 (2020) https://www.meteobelgie.be/klimatologie/waarnemingen-en-analyses/jaar-2021/2291-analyse-van-het-jaar-2021 (2021). Accessed on 12/05/2022.Shrestha, M. et al. Fluorescent pan traps affect the capture rate of insect orders in different ways. Insects 10(2), 40. https://doi.org/10.3390/insects10020040 (2019).Article 

    Google Scholar 
    Cooper, R., & Whitmore, R.C. (1990) Arthropod sampling methods in ornithology, Avian Foraging: theory, methodology, and applications. Studies in Avian Biology 13, Cooper Ornithological Society, California.Oberprieler, S. K., Andersen, A. & Braby, M. F. Invertebrate by-catch from vertebrate pitfall rraps can be useful for documenting patterns of invertebrate diversity. J. Insect. Conserv. 23(3), 547–554. https://doi.org/10.1007/s10841-019-00143-z (2019).Article 

    Google Scholar 
    Skvarla, M. J., Larson, J. L. & Dowling, A. P. G. Pitfalls and preservatives: A review. J. Entomol. Soc. Ontario 145, 15–43 (2014).
    Google Scholar 
    Michez, D., Rasmont, P., Terzo, M. and Vereecken, N.J. (2019) Bees of Europe. Hymenoptera of Europe 1. N.A.P Editions.Williams, P.H., et al. (2012): Unveiling cryptic species of the bumblebee subgenus Bombus s. str. worldwide with COI barcodes (Hymenoptera: Apidae). Syste. Biodiversity. https://doi.org/10.1080/14772000.2012.66457Falck, S., & Lewington, R (2020) Bijen veldgids voor Nederland en Vlaanderen. Tirion.Koster, A. (2022) De Nederlandse wilde bijen en hun planten. http://www.denederlandsebijen.nl/. Accessed on 21/4/2022.Speight, M.C.D. & Sarthou, J.P. (2013) StN keys for the identification of adult European Syrphidae (Diptera) 2013/Clés StN pour la détermination des adultes des Syrphidae Européens (Diptères) 2013. Syrph the Net, the database of European Syrphidae, Vol. 74, 133pp, Syrph the Net publications, Dublin.Roback, P., Legler, J. (2021) Beyond Multiple Linear Regression: Applied Generalized Linear Models and Multilevel Models in R. Taylor & Francis Group, LLC.R Core Team (2020) R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna, Austria.Oksanen, J., et al. (2014) Vegan: community ecology package. R Package 280.Bengtsson, H. (2017). matrixStats: Functions that Apply to Rows and Columns of Matrices (and to Vectors). R Package Version 0.52.2.Bates, D., Mächler, M., Bolker, B., & Walker, S. (2015) Fitting linear mixed-effects models using lme4. J. Stat. Softw. 67(1), 1–48. https://doi.org/10.18637/jss.v067.i01.Wickham, H., François, R., Henry, L. and Müller, K. (2022). dplyr: A Grammar of Data Manipulation. https://dplyr.tidyverse.org, https://github.com/tidyverse/dplyr.Venables, W.N., & Ripley, B.D. (2002) Modern Applied Statistics with S, 4th ed. Springer, New York. ISBN 0–387–95457–0. https://www.stats.ox.ac.uk/pub/MASS4/.Ricotta, C. & Moretti, M. CWM and Rao’s quadratic diversity: A unified framework for functional ecology. Oecologia 167(1), 181–188 (2011).Article 
    ADS 

    Google Scholar 
    Leclère, D. et al. Bending the curve of terrestrial biodiversity needs an integrated strategy. Nature 585(7826), 551–556. https://doi.org/10.1038/s41586-020-2705-y (2020).Article 
    ADS 
    CAS 

    Google Scholar 
    Drossart, M., et al. (2019) Belgian red list of Bees. Belgian Science Policy (BRAIN-be – (Belgian Research Action through Interdisciplinary Networks). Mons: Presse universitaire de l’Université de Mons.Fahrig, L. Why do several small patches hold more species than few large patches?. Glob. Ecol. Biogeogr. 29, 615–628. https://doi.org/10.1111/geb.13059 (2020).Article 

    Google Scholar 
    Ayers, A. C. & Rehan, S. M. Supporting bees in cities: how bees are influenced by local and landscape features. Insects 12, 128. https://doi.org/10.3390/insects12020128 (2021).Article 

    Google Scholar 
    Domínguez, M. V. S., González, E., Fabián, D., Salvo, A. & Fenoglio, M. S. Arthropod diversity and ecological processes on green roofs in a semi-rural area of Argentina: Similarity to neighbor ground habitats and landscape effects. Landscape and Urban Planning 199, (2020).Castagneyrol, B. & Jactel, H. Unravelling plant- animal diversity relationships: A meta-regression analysis. Ecology 93(9), 2115–2124 (2012).Article 

    Google Scholar 
    Harrison, T., Gibbs, J. & Winfree, R. Phylogenetic homogenization of bee communities across ecoregions. Glob. Ecol. Biogeogr. 27, 1457–1466. https://doi.org/10.1111/geb.12822 (2018).Article 

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

    Google Scholar 
    Martins, K. T., Gonzalez, A. & Lechowicz, M. J. Patterns of pollinator turnover and increasing diversity associated with urban habitats. Urban Ecosyst. 20, 1359–1371 (2017).Article 

    Google Scholar 
    Bucholz, S. & Egerer, M. Functional ecology of wild bees in cities: Towards a better understanding of trait-urbanisation relationships. Biodiver. Conserv. 29, 2779–2801 (2020).Article 

    Google Scholar 
    Hernandez, J. L., Frankie, G. W. & Thorp, R. W. Ecology of urban bees : A review of current knowledge and directions for future study. Cities Environ. 2, 1–15 (2009).Article 

    Google Scholar 
    Cane, J. H. Bees, pollination, and the challenges of sprawl. In Nature in fragments: The legacy of sprawl (eds Johnson, E. A. & Klemens, M. W.) 109–124 (Columbia University Press, 2005).Chapter 

    Google Scholar 
    Koch, K. Wilde bijensoorten in een stedelijke omgeving: Stad Antwerpen. Antenna 4, 8–12 (2014).
    Google Scholar 
    Soper, J. & Beggs, J. Assessing the impact of an introduced bee, Anthidium manicatum, on pollinator communities in New Zealand. NZ J. Bot. 51(3), 213–228. https://doi.org/10.1080/0028825X.2013.793202 (2013).Article 

    Google Scholar 
    Bennet, D.G., Kelly, D., & Clemens, J. (2018). Food plants and foraging distances for the native bee Lasioglossum sordidum in Christchurch Botanic Gardens. New Zealand J. Ecol. 42(1), 40–47. https://doi.org/10.20417/nzjecol.42.1Vanormelingen, P., Remer, M., & D’Haeseleer, J. (2021) Wilde bijen en bebouwing: meer verliezers dan winnaars? Themanummer bijen in de stad en dorp, Hymenovaria, maart 2021.Rader, R. et al. Alternative pollinator taxa are equally efficient but not as effective as the honey-bee in a mass flowering crop. J. Appl. Ecol. 46(5), 1080–1087. https://doi.org/10.1111/j.1365-2664.2009.01700.x (2009).Article 

    Google Scholar 
    Garantonakis, N. et al. Comparing the pollination services of honey bees and wild bees in a watermelon field. Sci. Hortic. 204, 138–144. https://doi.org/10.1016/j.scienta.2016.04.006 (2016).Article 

    Google Scholar 
    Foldesi, R., Howlett, B. G., Grass, I. & Batary, P. Larger pollinators deposit more pollen on stigmas across multiple plant species – A meta-analysis. J. Appl. Ecol. 58(4), 699–707. https://doi.org/10.1111/1365-2664.13798 (2021).Article 

    Google Scholar 
    Howlett, et al. (2011). Can insect body pollen counts be used to estimate pollen deposition on pak choi stigmas? New Zealand Plant Protection 64, 25–31. https://doi.org/10.30843/nzpp.2011.64.5951Nelson, W., Barry Donovan, L. E. & Howlett, B. Lasioglossum bees – the forgotten pollinators. J. Apic. Res. https://doi.org/10.1080/00218839.2022.2028966 (2022).Article 

    Google Scholar 
    Passaseo, A., Pétremand, G., Rochefort, S. & Castella, E. Pollinators emerging from extensive green roofs: Wild bees (Hymenoptera: Antophila) and hoverflies (Diptera: Syrphidae) in Geneva (Switzerland). Urban Ecosyst. 23, 1079–1086. https://doi.org/10.1007/s11252-020-00973-9 (2020).Article 

    Google Scholar 
    Odanaka, K. A. & Rehan, S. M. Impact indicators: Effects of land use management on functional trait and phylogenetic diversity of wild bees. Agric. Ecosyst. Environ. 286, 106663 (2019).Article 

    Google Scholar 
    Wilson, C. J. & Jamieson, M. A. The effects of urbanisation on bee communities depends on floral resource availability and bee functional traits. PLoS ONE 14(12), e0225852. https://doi.org/10.1371/journal.pone.0225852 (2019).Article 
    CAS 

    Google Scholar 
    Osborne, J. L. et al. Quantifying and comparing bumblebee nest densities in gardens and countryside habitats. J. Appl. Ecol. 45, 784–792. https://doi.org/10.1111/j.1365-2664.2007.01359.x (2007).Article 

    Google Scholar 
    Glaum, P., Simao, M. C., Vaidya, C., Fitch, G. & Lulinao, B. Big city Bombus: Using natural history and land-use history to find significant environmental drivers in bumble-bee declines in urban development. R Soc Open Sci. 4, 170156. https://doi.org/10.1098/rsos.170156 (2017) (PMID: 28573023).Article 
    ADS 

    Google Scholar 
    Rasmont, P. et al. Climatic risk and distribution atlas of European bumblebees. Biorisk 10, 1–246 (2015).Article 

    Google Scholar 
    Roger, N. et al. Impact of pollen resources drift on common bumblebees in NW Europe. Glob. Change Biol. 23, 68–76 (2017).Article 
    ADS 

    Google Scholar 
    Frankie, G. W. et al. Ecological patterns of bees and their host ornamental flowers in two northern California cities. J. Kansas Entomol. Soc. 78, 227–246 (2005).Article 

    Google Scholar 
    Lerman, S. B. & Milam, J. Bee fauna and floral abundance within lawn-dominated suburban yards in Springfield, MA. Ann. Entomol. Soc. Am. 109, 713–723 (2016).Article 
    CAS 

    Google Scholar 
    Braaker, S., Obrist, M. K., Ghazoul, J. & Moretti, M. Habitat connectivity and local conditions shape taxonomic and functional diversity of arthropods on green roofs. J. Anim. Ecol. 86, 521–531. https://doi.org/10.1111/1365-2656.12648 (2017).Article 

    Google Scholar 
    Passaseo, A., Rochefort, S., Pétremand, G., & Castella, E. (2021) Pollinators on green roofs: Diversity and trait analysis of wild bees (Hymenoptera: Anthophila) and Hoverflies (Diptera: Syrphidae) in an urban area (Geneva, Switzerland). Cities and the Environment (CATE) https://doi.org/10.15365/cate.2021.140201Hennig, E. & Ghazoul, J. Pollinating animals in the urban environment. Urban Ecosyst. 15, 149–166. https://doi.org/10.1007/s11252-011-0202-7 (2012).Article 

    Google Scholar 
    Mecke R. (1996) Die fauna begrünter dächer: Ökologische untersuchung verschiedener dachflächer im Hamburger stadtgebiet. University of Hamburg, Diploma dissertation.Bevk, D. The diversity of pollinators on green roofs. Acta Entomol. Slovenica 29(1), 5–14 (2021).
    Google Scholar 
    Speight, M.C.D. (2011) Species accounts of European Syrphidae (Diptera), Glasgow 2011. Syrph the Net, the database of European Syrphidae, vol. 65, 285 pp., Syrph the Net publications, Dublin.Wotton, K. R. et al. Mass seasonal migrations of hoverflies provide extensive pollination and crop protection services. Curr. Biol. 29, 2167–2173 (2019).Article 
    CAS 

    Google Scholar 
    Boyer, K. J., Fragoso, F. P., Mabin, M. E. D. & Brunet, J. Netting and pan traps fail to identify the pollinator guild of an agricultural crop. Nat. Res. Sci. Rep. 10, 13819. https://doi.org/10.1038/s41598-020-70518-9 (2020).Article 
    CAS 

    Google Scholar  More

  • in

    A predictive timeline of wildlife population collapse

    Ceballos, G. et al. Accelerated modern human-induced species losses: entering the sixth mass extinction. Sci. Adv. 1, e1400253 (2015).Article 

    Google Scholar 
    Dereniowska, M. & Meinard, Y. The unknownness of biodiversity: its value and ethical significance for conservation action. Biol. Conserv. 260, 109199 (2021).Article 

    Google Scholar 
    Maron, M. et al. Towards a threat assessment framework for ecosystem services. Trends Ecol. Evol. 32, 240–248 (2017).Article 

    Google Scholar 
    Tilman, D. et al. Future threats to biodiversity and pathways to their prevention. Nature 546, 73–81 (2017).Article 
    CAS 

    Google Scholar 
    Taborsky, B. et al. Towards an evolutionary theory of stress responses. Trends Ecol. Evol. 36, 39–48 (2021).Article 

    Google Scholar 
    van de Leemput, I. A., Dakos, V., Scheffer, M. & van Nes, E. H. Slow recovery from local disturbances as an indicator for loss of ecosystem resilience. Ecosystems 21, 141–152 (2018).Article 

    Google Scholar 
    Fagan, W. F. & Holmes, E. E. Quantifying the extinction vortex. Ecol. Lett. 9, 51–60 (2005).
    Google Scholar 
    Williams, N. F., McRae, L., Freeman, R., Capdevila, P. & Clements, C. F. Scaling the extinction vortex: body size as a predictor of population dynamics close to extinction events. Ecol. Evol. 11, 7069–7079 (2021).Article 

    Google Scholar 
    Clements, C. F. & Ozgul, A. Indicators of transitions in biological systems. Ecol. Lett. 21, 905–919 (2018).Article 

    Google Scholar 
    Shaffer, M. L. in Challenges in the Conservation of Biological Resources (eds. Decker, D. J., Krasny, M. E., Goff, G. R., Smith, C. R. & Gross, D. W.) 107–118 (Routledge, 2019).Scheffer, M. et al. Early-warning signals for critical transitions. Nature 461, 53–59 (2009).Article 
    CAS 

    Google Scholar 
    Gardner, T. A. et al. The cost-effectiveness of biodiversity surveys in tropical forests. Ecol. Lett. 11, 139–150 (2008).Article 

    Google Scholar 
    Coulson, T., Mace, G. M., Hudson, E. & Possingham, H. The use and abuse of population viability analysis. Trends Ecol. Evol. 16, 219–221 (2001).Article 
    CAS 

    Google Scholar 
    Clements, C. F., Drake, J. M., Griffiths, J. I. & Ozgul, A. Factors influencing the detectability of early warning signals of population collapse. Am. Nat. 186, 50–58 (2015).Article 

    Google Scholar 
    Patterson, A. C., Strang, A. G. & Abbott, K. C. When and where we can expect to see early warning signals in multispecies systems approaching tipping points: insights from theory. Am. Nat. 198, E12–E26 (2021).Article 

    Google Scholar 
    Vinton, A. C., Gascoigne, S. J. L., Sepil, I. & Salguero-Gómez, R. Plasticity’s role in adaptive evolution depends on environmental change components. Trends Ecol. Evol. https://doi.org/10.1016/j.tree.2022.08.008 (2022).Levin, S. A. The problem of pattern and scale in ecology: the Robert H. MacArthur Award lecture. Ecology 73, 1943–1967 (1992).Article 

    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 
    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 
    CAS 

    Google Scholar 
    Haberle, I., Marn, N., Geček, S. & Klanjšček, T. Dynamic energy budget of endemic and critically endangered bivalve Pinna nobilis: a mechanistic model for informed conservation. Ecol. Model. 434, 109207 (2020).Article 

    Google Scholar 
    Gislason, H., Daan, N., Rice, J. C. & Pope, J. G. Size, growth, temperature and the natural mortality of marine fish. Fish Fish. 11, 149–158 (2010).Article 

    Google Scholar 
    Jennings, S. & Blanchard, J. L. Fish abundance with no fishing: predictions based on macroecological theory. J. Anim. Ecol. 73, 632–642 (2004).Article 

    Google Scholar 
    Valderrama, D. & Fields, K. H. Flawed evidence supporting the metabolic theory of ecology may undermine goals of ecosystem-based fishery management: the case of invasive Indo-Pacific lionfish in the western Atlantic. ICES J. Mar. Sci. 74, 1256–1267 (2017).Article 

    Google Scholar 
    Marshall, D. J. & McQuaid, C. D. Warming reduces metabolic rate in marine snails: adaptation to fluctuating high temperatures challenges the metabolic theory of ecology. Proc. R. Soc. B 278, 281–288 (2011).Article 

    Google Scholar 
    Rombouts, I., Beaugrand, G., Ibaňez, F., Chiba, S. & Legendre, L. Marine copepod diversity patterns and the metabolic theory of ecology. Oecologia 166, 349–355 (2011).Article 

    Google Scholar 
    Allen, A. P. & Gillooly, J. F. The mechanistic basis of the metabolic theory of ecology. Oikos 116, 1073–1077 (2022).Article 

    Google Scholar 
    Lawton, J. H. From physiology to population dynamics and communities. Funct. Ecol. 5, 155–161 (1991).Article 

    Google Scholar 
    Ames, E. M. et al. Striving for population-level conservation: integrating physiology across the biological hierarchy. Conserv. Physiol. 8, coaa019 (2020).Article 

    Google Scholar 
    Berger-Tal, O. et al. Integrating animal behavior and conservation biology: a conceptual framework. Behav. Ecol. 22, 236–239 (2011).Article 

    Google Scholar 
    Baruah, G., Clements, C. F., Guillaume, F. & Ozgul, A. When do shifts in trait dynamics precede population declines? Am. Nat. 193, 633–644 (2019).Article 

    Google Scholar 
    Dakos, V. et al. Methods for detecting early warnings of critical transitions in time series illustrated using simulated ecological data. PLoS ONE 7, e41010 (2012).Article 
    CAS 

    Google Scholar 
    Ward, R. J., Griffiths, R. A., Wilkinson, J. W. & Cornish, N. Optimising monitoring efforts for secretive snakes: a comparison of occupancy and N-mixture models for assessment of population status. Sci. Rep. 7, 18074 (2017).Article 

    Google Scholar 
    Thompson, W. Sampling Rare or Elusive Species: Concepts, Designs, and Techniques for Estimating Population Parameters (Island Press, 2013).Clements, C. F., Blanchard, J. L., Nash, K. L., Hindell, M. A. & Ozgul, A. Body size shifts and early warning signals precede the historic collapse of whale stocks. Nat. Ecol. Evol. 1, 0188 (2017).Article 

    Google Scholar 
    Burant, J. B., Park, C., Betini, G. S. & Norris, D. R. Early warning indicators of population collapse in a seasonal environment. J. Anim. Ecol. 90, 1538–1549 (2021).Article 

    Google Scholar 
    Tuomainen, U. & Candolin, U. Behavioural responses to human-induced environmental change. Biol. Rev. 86, 640–657 (2011).Article 

    Google Scholar 
    Mazza, V., Dammhahn, M., Lösche, E. & Eccard, J. A. Small mammals in the big city: behavioural adjustments of non-commensal rodents to urban environments. Glob. Change Biol. 26, 6326–6337 (2020).Article 

    Google Scholar 
    Hendry, A. P., Farrugia, T. J. & Kinnison, M. T. Human influences on rates of phenotypic change in wild animal populations. Mol. Ecol. 17, 20–29 (2008).Article 

    Google Scholar 
    Speakman, J. R., Król, E. & Johnson, M. S. The functional significance of individual variation in basal metabolic rate. Physiol. Biochem. Zool. 77, 900–915 (2004).Article 

    Google Scholar 
    Péron, G. et al. Evidence of reduced individual heterogeneity in adult survival of long-lived species. Evolution 70, 2909–2914 (2016).Article 

    Google Scholar 
    Fleming, A. H., Clark, C. T., Calambokidis, J. & Barlow, J. Humpback whale diets respond to variance in ocean climate and ecosystem conditions in the California Current. Glob. Change Biol. 22, 1214–1224 (2016).Article 

    Google Scholar 
    Kirkwood, T. B. L., Rose, M. R., Harvey, P. H., Partridge, L. & Southwood, S. R. Evolution of senescence: late survival sacrificed for reproduction. Phil. Trans. R. Soc. Lond. B 332, 15–24 (1991).Article 
    CAS 

    Google Scholar 
    Mallela, A. & Hastings, A. The role of stochasticity in noise-induced tipping point cascades: a master equation approach. Bull. Math. Biol. 83, 53 (2021).Article 

    Google Scholar 
    Burthe, S. J. et al. Do early warning indicators consistently predict nonlinear change in long-term ecological data? J. Appl. Ecol. 53, 666–676 (2016).Article 

    Google Scholar 
    Vucetich, J. A. & Waite, T. A. Erosion of heterozygosity in fluctuating populations. Conserv. Biol. 13, 860–868 (1999).Article 

    Google Scholar 
    Kramer, A. M. & Drake, J. M. Experimental demonstration of population extinction due to a predator-driven Allee effect. J. Anim. Ecol. 79, 633–639 (2010).Article 

    Google Scholar 
    Oram, E. & Spitze, K. Depth selection by Daphnia pulex in response to Chaoborus kairomone. Freshw. Biol. 58, 409–415 (2013).Article 

    Google Scholar 
    Trites, A. W. & Donnelly, C. P. The decline of Steller sea lions Eumetopias jubatus in Alaska: a review of the nutritional stress hypothesis. Mammal. Rev. 33, 3–28 (2003).Article 

    Google Scholar 
    Sibly, R. M., Barker, D., Hone, J. & Pagel, M. On the stability of populations of mammals, birds, fish and insects. Ecol. Lett. 10, 970–976 (2007).Article 

    Google Scholar 
    Dakos, V. et al. Ecosystem tipping points in an evolving world. Nat. Ecol. Evol. 3, 355–362 (2019).Article 

    Google Scholar 
    Dingemanse, N. J., Kazem, A. J. N., Réale, D. & Wright, J. Behavioural reaction norms: animal personality meets individual plasticity. Trends Ecol. Evol. 25, 81–89 (2010).Article 

    Google Scholar 
    Tanner, R. L. & Dowd, W. W. Inter-individual physiological variation in responses to environmental variation and environmental change: integrating across traits and time. Comp. Biochem. Physiol. A 238, 110577 (2019).Article 
    CAS 

    Google Scholar 
    Patrick, S. C., Martin, J. G. A., Ummenhofer, C. C., Corbeau, A. & Weimerskirch, H. Albatrosses respond adaptively to climate variability by changing variance in a foraging trait. Glob. Change Biol. 27, 4564–4574 (2021).Article 
    CAS 

    Google Scholar 
    Fayet, A. L., Clucas, G. V., Anker‐Nilssen, T., Syposz, M. & Hansen, E. S. Local prey shortages drive foraging costs and breeding success in a declining seabird, the Atlantic puffin. J. Anim. Ecol. https://doi.org/10.1111/1365-2656.13442 (2021).Pierce, C. L. Predator avoidance, microhabitat shift, and risk-sensitive foraging in larval dragonflies. Oecologia 77, 81–90 (1988).Article 
    CAS 

    Google Scholar 
    Leibold, M. & Tessier, A. J. Contrasting patterns of body size for Daphnia species that segregate by habitat. Oecologia 86, 342–348 (1991).Article 

    Google Scholar 
    Charmantier, A. & Gienapp, P. Climate change and timing of avian breeding and migration: evolutionary versus plastic changes. Evol. Appl. 7, 15–28 (2014).Article 

    Google Scholar 
    Kopp, M. & Matuszewski, S. Rapid evolution of quantitative traits: theoretical perspectives. Evol. Appl. 7, 169–191 (2014).Article 

    Google Scholar 
    Williams, J. W., Ordonez, A. & Svenning, J.-C. A unifying framework for studying and managing climate-driven rates of ecological change. Nat. Ecol. Evol. 5, 17–26 (2021).Article 

    Google Scholar 
    Jaureguiberry, P. et al. The direct drivers of recent global anthropogenic biodiversity loss. Sci. Adv. 8, eabm9982 (2022).Article 

    Google Scholar 
    Chevin, L.-M., Collins, S. & Lefèvre, F. Phenotypic plasticity and evolutionary demographic responses to climate change: taking theory out to the field. Funct. Ecol. 27, 967–979 (2013).Article 

    Google Scholar 
    Ferriere, R. & Legendre, S. Eco-evolutionary feedbacks, adaptive dynamics and evolutionary rescue theory. Phil. Trans. R. Soc. B 368, 20120081 (2013).Article 

    Google Scholar 
    Rebecchi, L., Boschetti, C. & Nelson, D. R. Extreme-tolerance mechanisms in meiofaunal organisms: a case study with tardigrades, rotifers and nematodes. Hydrobiologia 847, 2779–2799 (2020).Article 

    Google Scholar 
    Hansson, B. & Westerberg, L. On the correlation between heterozygosity and fitness in natural populations. Mol. Ecol. 11, 2467–2474 (2002).Article 

    Google Scholar 
    Mammola, S., Carmona, C. P., Guillerme, T. & Cardoso, P. Concepts and applications in functional diversity. Funct. Ecol. 35, 1869–1885 (2021).Article 
    CAS 

    Google Scholar 
    McClanahan, T. R. et al. Highly variable taxa-specific coral bleaching responses to thermal stresses. Mar. Ecol. Prog. Ser. 648, 135–151 (2020).Article 

    Google Scholar 
    Reside, A. E. et al. Beyond the model: expert knowledge improves predictions of species’ fates under climate change. Ecol. Appl. 29, e01824 (2019).Article 

    Google Scholar 
    Desjonquères, C., Gifford, T. & Linke, S. Passive acoustic monitoring as a potential tool to survey animal and ecosystem processes in freshwater environments. Freshw. Biol. 65, 7–19 (2020).Article 

    Google Scholar 
    Sequeira, A. M. M. et al. A standardisation framework for bio-logging data to advance ecological research and conservation. Methods Ecol. Evol. 12, 996–1007 (2021).Article 

    Google Scholar 
    Shimada, T. et al. Optimising sample sizes for animal distribution analysis using tracking data. Methods Ecol. Evol. 12, 288–297 (2021).Article 

    Google Scholar 
    Wauchope, H. S. et al. Evaluating impact using time-series data. Trends Ecol. Evol. 36, 196–205 (2021).Article 

    Google Scholar 
    Krause, D. J., Hinke, J. T., Perryman, W. L., Goebel, M. E. & LeRoi, D. J. An accurate and adaptable photogrammetric approach for estimating the mass and body condition of pinnipeds using an unmanned aerial system. PLoS ONE 12, e0187465 (2017).Article 

    Google Scholar 
    Besson, M. et al. Towards the fully automated monitoring of ecological communities. Ecol. Lett. https://doi.org/10.1111/ele.14123 (2022).Article 

    Google Scholar 
    Cavender-Bares, J. et al. Integrating remote sensing with ecology and evolution to advance biodiversity conservation. Nat. Ecol. Evol. 6, 506–519 (2022).Article 

    Google Scholar 
    Ingram, D. J., Ferreira, G. B., Jones, K. E. & Mace, G. M. Targeting conservation actions at species threat response thresholds. Trends Ecol. Evol. 36, 216–226 (2021).Article 

    Google Scholar 
    Keith, S. A. et al. Synchronous behavioural shifts in reef fishes linked to mass coral bleaching. Nat. Clim. Change 8, 986–991 (2018).Article 

    Google Scholar 
    Drake, J. M. & Griffen, B. D. Early warning signals of extinction in deteriorating environments. Nature 467, 456–459 (2010).Article 
    CAS 

    Google Scholar 
    Enquist, B. J. et al. in Advances in Ecological Research Vol. 52 (eds Pawar, S. et al.) 249–318 (Academic Press, 2015).Wei, W. W. S. Multivariate Time Series Analysis and Applications (John Wiley & Sons, 2018).Holmes, E. E., Ward, E. J. & Wills, K. MARSS: multivariate autoregressive state-space models for analyzing time-series data. R J. 4, 11–19 (2012).Article 

    Google Scholar 
    Zhu, M., Yamakawa, T. & Sakai, T. Combined use of trawl fishery and research vessel survey data in a multivariate autoregressive state-space (MARSS) model to improve the accuracy of abundance index estimates. Fish. Sci. 84, 437–451 (2018).Article 
    CAS 

    Google Scholar 
    Lai, G., Chang, W.-C., Yang, Y. & Liu, H. Modeling long- and short-term temporal patterns with deep neural networks. In The 41st International ACM SIGIR Conference on Research & Development in Information Retrieval 95–104, https://doi.org/10.1145/3209978.3210006 (ACM, 2018).Bury, T. M. et al. Deep learning for early warning signals of tipping points. Proc. Natl Acad. Sci. USA 118, e2106140118 (2021).Article 
    CAS 

    Google Scholar 
    Lara-Benítez, P., Carranza-García, M. & Riquelme, J. C. An experimental review on deep learning architectures for time series forecasting. Int. J. Neural Syst. 31, 2130001 (2021).Article 

    Google Scholar 
    Guo, Q. et al. Application of deep learning in ecological resource research: theories, methods, and challenges. Sci. China Earth Sci. 63, 1457–1474 (2020).Article 

    Google Scholar 
    Rogers, T. L., Johnson, B. J. & Munch, S. B. Chaos is not rare in natural ecosystems. Nat. Ecol. Evol. 6, 1105–1111 (2022).Article 

    Google Scholar 
    Samplonius, J. M. et al. Phenological sensitivity to climate change is higher in resident than in migrant bird populations among European cavity breeders. Glob. Change Biol. 24, 3780–3790 (2018).Article 

    Google Scholar 
    Menzel, A. et al. Climate change fingerprints in recent European plant phenology. Glob. Change Biol. 26, 2599–2612 (2020).Article 

    Google Scholar 
    Koleček, J., Adamík, P. & Reif, J. Shifts in migration phenology under climate change: temperature vs. abundance effects in birds. Clim. Change 159, 177–194 (2020).Article 

    Google Scholar 
    Altermatt, F. et al. Big answers from small worlds: a user’s guide for protist microcosms as a model system in ecology and evolution. Methods Ecol. Evol. 6, 218–231 (2015).Article 

    Google Scholar 
    Beermann, A. J. et al. Multiple-stressor effects on stream macroinvertebrate communities: a mesocosm experiment manipulating salinity, fine sediment and flow velocity. Sci. Total Environ. 610–611, 961–971 (2018).Article 

    Google Scholar 
    Clements, C. F. & Ozgul, A. Including trait-based early warning signals helps predict population collapse. Nat. Commun. 7, 10984 (2016).Article 
    CAS 

    Google Scholar 
    Jacquet, C. & Altermatt, F. The ghost of disturbance past: long-term effects of pulse disturbances on community biomass and composition. Proc. R. Soc. B 287, 20200678 (2020).Article 

    Google Scholar 
    Greggor, A. L. et al. Research priorities from animal behaviour for maximising conservation progress. Trends Ecol. Evol. 31, 953–964 (2016).Article 

    Google Scholar 
    Couvillon, M. J., Schürch, R. & Ratnieks, F. L. W. Waggle dance distances as integrative indicators of seasonal foraging challenges. PLoS ONE 9, e93495 (2014).Article 

    Google Scholar 
    Hamilton, C. D., Lydersen, C., Ims, R. A. & Kovacs, K. M. Predictions replaced by facts: a keystone species’ behavioural responses to declining Arctic sea-ice. Biol. Lett. 11, 20150803 (2015).Article 

    Google Scholar 
    Holt, R. E. & Jørgensen, C. Climate change in fish: effects of respiratory constraints on optimal life history and behaviour. Biol. Lett. 11, 20141032 (2015).Article 

    Google Scholar 
    Gauzens, B. et al. Adaptive foraging behaviour increases vulnerability to climate change. Preprint at https://doi.org/10.1101/2021.05.05.442768 (2021).Lenda, M., Witek, M., Skórka, P., Moroń, D. & Woyciechowski, M. Invasive alien plants affect grassland ant communities, colony size and foraging behaviour. Biol. Invasions 15, 2403–2414 (2013).Article 

    Google Scholar 
    Hertel, A. G. et al. Don’t poke the bear: using tracking data to quantify behavioural syndromes in elusive wildlife. Anim. Behav. 147, 91–104 (2019).Article 

    Google Scholar 
    Tini, M. et al. Use of space and dispersal ability of a flagship saproxylic insect: a telemetric study of the stag beetle (Lucanus cervus) in a relict lowland forest. Insect Conserv. Divers. 11, 116–129 (2018).Article 

    Google Scholar 
    Kunc, H. P. & Schmidt, R. Species sensitivities to a global pollutant: a meta-analysis on acoustic signals in response to anthropogenic noise. Glob. Change Biol. 27, 675–688 (2021).Article 

    Google Scholar 
    Anestis, A., Lazou, A., Pörtner, H. O. & Michaelidis, B. Behavioral, metabolic, and molecular stress responses of marine bivalve Mytilus galloprovincialis during long-term acclimation at increasing ambient temperature. Am. J. Physiol. 293, R911–R921 (2007).CAS 

    Google Scholar 
    Pacherres, C. O., Schmidt, G. M. & Richter, C. Autotrophic and heterotrophic responses of the coral Porites lutea to large amplitude internal waves. J. Exp. Biol. 216, 4365–4374 (2013).
    Google Scholar 
    Ban, S. S., Graham, N. A. J. & Connolly, S. R. Evidence for multiple stressor interactions and effects on coral reefs. Glob. Change Biol. 20, 681–697 (2014).Article 

    Google Scholar 
    Singh, R., Prathibha, P. & Jain, M. Effect of temperature on life-history traits and mating calls of a field cricket, Acanthogryllus asiaticus. J. Therm. Biol. 93, 102740 (2020).Article 

    Google Scholar 
    Pellegrini, A. Y., Romeu, B., Ingram, S. N. & Daura-Jorge, F. G. Boat disturbance affects the acoustic behaviour of dolphins engaged in a rare foraging cooperation with fishers. Anim. Conserv. 24, 613–625 (2021).Article 

    Google Scholar 
    McMahan, M. D. & Grabowski, J. H. Nonconsumptive effects of a range-expanding predator on juvenile lobster (Homarus americanus) population dynamics. Ecosphere 10, e02867 (2019).Article 

    Google Scholar 
    Vilhunen, S., Hirvonen, H. & Laakkonen, M. V.-M. Less is more: social learning of predator recognition requires a low demonstrator to observer ratio in Arctic charr (Salvelinus alpinus). Behav. Ecol. Sociobiol. 57, 275–282 (2005).Article 

    Google Scholar 
    Ortega, Z., Mencía, A. & Pérez-Mellado, V. Rapid acquisition of antipredatory responses to new predators by an insular lizard. Behav. Ecol. Sociobiol. 71, 1 (2017).Article 

    Google Scholar 
    Fox, R. J., Donelson, J. M., Schunter, C., Ravasi, T. & Gaitán-Espitia, J. D. Beyond buying time: the role of plasticity in phenotypic adaptation to rapid environmental change. Phil. Trans. R. Soc. B 374, 20180174 (2019).Article 

    Google Scholar 
    Pigeon, G., Ezard, T. H. G., Festa-Bianchet, M., Coltman, D. W. & Pelletier, F. Fluctuating effects of genetic and plastic changes in body mass on population dynamics in a large herbivore. Ecology 98, 2456–2467 (2017).Article 

    Google Scholar 
    Lomolino, M. V. & Perault, D. R. Body size variation of mammals in a fragmented, temperate rainforest. Conserv. Biol. 21, 1059–1069 (2007).Article 

    Google Scholar 
    Gardner, J. L., Peters, A., Kearney, M. R., Joseph, L. & Heinsohn, R. Declining body size: a third universal response to warming? Trends Ecol. Evol. 26, 285–291 (2011).Article 

    Google Scholar 
    Sheridan, J. A. & Bickford, D. Shrinking body size as an ecological response to climate change. Nat. Clim. Change 1, 401–406 (2011).Article 

    Google Scholar 
    Thoral, E. et al. Changes in foraging mode caused by a decline in prey size have major bioenergetic consequences for a small pelagic fish. J. Anim. Ecol. 90, 2289–2301 (2021).Article 

    Google Scholar 
    Stirling, I. & Derocher, A. E. Effects of climate warming on polar bears: a review of the evidence. Glob. Change Biol. 18, 2694–2706 (2012).Article 

    Google Scholar 
    Spanbauer, T. L. et al. Body size distributions signal a regime shift in a lake ecosystem. Proc. R. Soc. B 283, 20160249 (2016).Article 

    Google Scholar 
    Bjorndal, K. A. et al. Ecological regime shift drives declining growth rates of sea turtles throughout the West Atlantic. Glob. Change Biol. 23, 4556–4568 (2017).Article 

    Google Scholar 
    Eshun-Wilson, F., Wolf, R., Andersen, T., Hessen, D. O. & Sperfeld, E. UV radiation affects antipredatory defense traits in Daphnia pulex. Ecol. Evol. 10, 14082–14097 (2020).Article 

    Google Scholar 
    Zhang, H., Hollander, J. & Hansson, L.-A. Bi-directional plasticity: rotifer prey adjust spine length to different predator regimes. Sci. Rep. 7, 10254 (2017).Article 

    Google Scholar 
    Simbula, G., Vignoli, L., Carretero, M. A. & Kaliontzopoulou, A. Fluctuating asymmetry as biomarker of pesticides exposure in the Italian wall lizards (Podarcis siculus). Zoology 147, 125928 (2021).Article 

    Google Scholar 
    Leary, R. F. & Allendorf, F. W. Fluctuating asymmetry as an indicator of stress: implications for conservation biology. Trends Ecol. Evol. 4, 214–217 (1989).Article 
    CAS 

    Google Scholar 
    Gavrilchuk, K. et al. Trophic niche partitioning among sympatric baleen whale species following the collapse of groundfish stocks in the Northwest Atlantic. Mar. Ecol. Prog. Ser. 497, 285–301 (2014).Article 

    Google Scholar 
    Kershaw, J. L. et al. Declining reproductive success in the Gulf of St. Lawrence’s humpback whales (Megaptera novaeangliae) reflects ecosystem shifts on their feeding grounds. Glob. Change Biol. 27, 1027–1041 (2021).Article 
    CAS 

    Google Scholar 
    Rode, K. D., Amstrup, S. C. & Regehr, E. V. Reduced body size and cub recruitment in polar bears associated with sea ice decline. Ecol. Appl. 20, 768–782 (2010).Article 

    Google Scholar 
    Obbard, M. E. et al. Re-assessing abundance of Southern Hudson Bay polar bears by aerial survey: effects of climate change at the southern edge of the range. Arct. Sci. 4, 634–655 (2018).Article 

    Google Scholar 
    Hutchings, J. A. The cod that got away. Nature 428, 899–900 (2004).Article 
    CAS 

    Google Scholar 
    Zhang, F. Early warning signals of population productivity regime shifts in global fisheries. Ecol. Indic. 115, 106371 (2020).Article 

    Google Scholar 
    Fulton, G. R. The Bramble Cay melomys: the first mammalian extinction due to human-induced climate change. Pac. Conserv. Biol. 23, 1–3 (2017).Article 

    Google Scholar  More

  • in

    Ecologically unequal exchanges driven by EU consumption

    Rockström, J. et al. A safe operation space for humanity. Nature 461, 472–475 (2009).Article 

    Google Scholar 
    Chancel, L., Piketty, T., Saez, E. & Zucman, G. World Inequality Report 2022 (Belknap Press, 2022).Ivanova, D. et al. Environmental impact assessment of household consumption. J. Ind. Ecol. 20, 526–536 (2016).Article 
    CAS 

    Google Scholar 
    Steen-Olsen, K., Weinzettel, J., Cranston, G., Ercin, A. E. & Hertwich, E. G. Carbon, land, and water footprint accounts for the European Union: consumption, production, and displacements through international trade. Environ. Sci. Technol. 46, 10883–10891 (2012).Article 
    CAS 

    Google Scholar 
    Tukker, A. et al. Environmental and resource footprints in a global context: Europe’s structural deficit in resource endowments. Glob. Environ. Change 40, 171–181 (2016).Article 

    Google Scholar 
    Bruckner, B., Hubacek, K., Shan, Y., Zhong, H. & Feng, K. Impacts of poverty alleviation on national and global carbon emissions. Nat. Sustain. 5, 311–320 (2022).Article 

    Google Scholar 
    Hubacek, K. et al. Global carbon inequality. Energy, Ecol. Environ. 2, 361–369 (2017).Article 

    Google Scholar 
    Yu, Y., Feng, K. & Hubacek, K. Tele-connecting local consumption to global land use. Glob. Environ. Change 23, 1178–1186 (2013).Article 

    Google Scholar 
    Wilting, H. C., Schipper, A. M., Bakkenes, M., Meijer, J. R. & Huijbregts, M. A. J. Quantifying biodiversity losses due to human consumption: a global-scale footprint analysis. Environ. Sci. Technol. 51, 3298–3306 (2017).Article 
    CAS 

    Google Scholar 
    Lucas, P. L., Wilting, H. C., Hof, A. F. & Van Vuuren, D. P. Allocating planetary boundaries to large economies: distributional consequences of alternative perspectives on distributive fairness. Glob. Environ. Change 60, 102017 (2020).Article 

    Google Scholar 
    Beylot, A. et al. Assessing the environmental impacts of EU consumption at macro-scale. J. Clean. Prod. 216, 382–393 (2019).Article 

    Google Scholar 
    Koslowski, M., Moran, D. D., Tisserant, A., Verones, F. & Wood, R. Quantifying Europe’s biodiversity footprints and the role of urbanization and income. Glob. Sustain. 3, e1 (2020).Lutter, S., Pfister, S., Giljum, S., Wieland, H. & Mutel, C. Spatially explicit assessment of water embodied in European trade: a product-level multi-regional input-output analysis. Glob. Environ. Change 38, 171–182 (2016).Article 

    Google Scholar 
    Stadler, K. et al. EXIOBASE 3 (3.8.1) [Data set]. Zenodo https://doi.org/10.5281/ZENODO.4588235 (2021).Roadmap to a Resource Efficient Europe (European Commission, 2011).Steinmann, Z. J. N. et al. Headline environmental indicators revisited with the global multi-regional input–output database EXIOBASE. J. Ind. Ecol. 22, 565–573 (2018).Article 

    Google Scholar 
    Ivanova, D. et al. Mapping the carbon footprint of EU regions. Environ. Res. Lett. 12, 054013 (2017).Wiedmann, T. O. et al. The material footprint of nations. Proc. Natl Acad. Sci. USA 112, 6271–6276 (2015).Article 
    CAS 

    Google Scholar 
    Lenzen, M. et al. Implementing the material footprint to measure progress towards Sustainable Development Goals 8 and 12. Nat. Sustain. 5, 157–166 (2022).Dorninger, C. et al. The effect of industrialization and globalization on domestic land-use: a global resource footprint perspective. Glob. Environ. Change 69, 102311 (2021).Article 

    Google Scholar 
    Mekonnen, M. M. & Gerbens-Leenes, W. The water footprint of food. Water 12, 12 (2020).Article 

    Google Scholar 
    Prell, C. & Feng, K. Unequal carbon exchanges: the environmental and economic impacts of iconic U.S. consumption items. J. Ind. Ecol. 20, 537–546 (2016).Article 

    Google Scholar 
    Prell, C., Feng, K., Sun, L., Geores, M. & Hubacek, K. The economic gains and environmental losses of US consumption: a world-systems and input-output approach. Soc. Forces 93, 405–428 (2014).Article 

    Google Scholar 
    Prell, C. Wealth and pollution inequalities of global trade: a network and input-output approach. Soc. Sci. J. 53, 111–121 (2016).Article 

    Google Scholar 
    World Economic Outlook (October 2022) (International Monetary Fund, 2022); https://www.imf.org/external/datamapper/datasets/WEOWilting, H. C., Schipper, A. M., Ivanova, O., Ivanova, D. & Huijbregts, M. A. J. Subnational greenhouse gas and land-based biodiversity footprints in the European Union. J. Ind. Ecol. 25, 79–94 (2021). https://doi.org/10.1111/jiec.13042Cabernard, L. & Pfister, S. A highly resolved MRIO database for analyzing environmental footprints and Green Economy Progress. Sci. Total Environ. 755, 142587 (2021).Jakob, M., Ward, H. & Steckel, J. C. Sharing responsibility for trade-related emissions based on economic benefits. Glob. Environ. Chang. 66, 102207 (2021).Article 

    Google Scholar 
    Wood, R. et al. The structure, drivers and policy implications of the European carbon footprint. Clim. Policy 20, S39–S57 (2020).Article 

    Google Scholar 
    Wood, R. et al. Growth in environmental footprints and environmental impacts embodied in trade: resource efficiency indicators from EXIOBASE3. J. Ind. Ecol. 22, 553–564 (2018).Article 

    Google Scholar 
    Hubacek, K., Chen, X., Feng, K., Wiedmann, T. & Shan, Y. Evidence of decoupling consumption-based CO2 emissions from economic growth. Adv. Appl. Energy 4, 100074 (2021).Article 

    Google Scholar 
    Wiedmann, T. & Lenzen, M. Environmental and social footprints of international trade. Nat. Geosci. 11, 314–321 (2018).Article 
    CAS 

    Google Scholar 
    Dorninger, C. et al. Global patterns of ecologically unequal exchange: Implications for sustainability in the 21st century. Ecol. Econ. 179, 106824 (2021).Article 

    Google Scholar 
    Hickel, J., Dorninger, C., Wieland, H. & Suwandi, I. Imperialist appropriation in the world economy: drain from the global South through unequal exchange, 1990–2015. Glob. Environ. Change 73, 102467 (2022).Poore, J. & Nemecek, T. Reducing food’s environmental impacts through producers and consumers. Science 360, 987–992 (2018).Article 
    CAS 

    Google Scholar 
    Ivanova, D. et al. Quantifying the potential for climate change mitigation of consumption options. Environ. Res. Lett. 15, 093001 (2020).Springmann, M. et al. Options for keeping the food system within environmental limits. Nature 562, 519–525 (2018).Article 
    CAS 

    Google Scholar 
    Ivanova, D. & Wood, R. The unequal distribution of household carbon footprints in Europe and its link to sustainability. Glob. Sustain. 3, e18 (2020).Hickel, J., O’Neill, D. W., Fanning, A. L. & Zoomkawala, H. National responsibility for ecological breakdown: a fair-shares assessment of resource use, 1970–2017. Lancet Planet. Heal. 6, e342–e349 (2022).Article 

    Google Scholar 
    Otto, I. M., Kim, K. M., Dubrovsky, N. & Lucht, W. Shift the focus from the super-poor to the super-rich. Nat. Clim. Change 9, 82–84 (2019).Article 

    Google Scholar 
    Wiedmann, T., Lenzen, M., Keyßer, L. T. & Steinberger, J. K. Scientists’ warning on affluence. Nat. Commun. 11, 3107 (2020).Nielsen, K. S., Nicholas, K. A., Creutzig, F., Dietz, T. & Stern, P. C. The role of high-socioeconomic-status people in locking in or rapidly reducing energy-driven greenhouse gas emissions. Nat. Energy 6, 1011–1016 (2021).Article 

    Google Scholar 
    Jakob, M. Why carbon leakage matters and what can be done against it. One Earth 4, 609–614 (2021).Article 

    Google Scholar 
    Lave, L. B. Using input–output analysis to estimate economy-wide discharges. Environ. Sci. Technol. 29, 420A–426A (1995).Article 
    CAS 

    Google Scholar 
    Wiedmann, T. A review of recent multi-region input–output models used for consumption-based emission and resource accounting. Ecol. Econ. 69, 211–222 (2009).Article 

    Google Scholar 
    Ewing, B. R. et al. Integrating ecological and water footprint accounting in a multi-regional input–output framework. Ecol. Indic. 23, 1–8 (2012).Article 

    Google Scholar 
    Brizga, J., Feng, K. & Hubacek, K. Household carbon footprints in the Baltic States: a global multi-regional input–output analysis from 1995 to 2011. Appl. Energy 189, 780–788 (2017).Hertwich, E. G. & Peters, G. P. Carbon footprint of nations: a global, trade-linked analysis. Environ. Sci. Technol. 43, 6414–6420 (2009).Article 
    CAS 

    Google Scholar 
    Zhong, H., Feng, K., Sun, L., Cheng, L. & Hubacek, K. Household carbon and energy inequality in Latin American and Caribbean countries. J. Environ. Manag. 273, 110979 (2020).Article 

    Google Scholar 
    Stadler, K. et al. EXIOBASE 3: developing a time series of detailed environmentally extended multi-regional input–output tables. J. Ind. Ecol. 22, 502–515 (2018).Article 

    Google Scholar 
    Hardadi, G., Buchholz, A. & Pauliuk, S. Implications of the distribution of German household environmental footprints across income groups for integrating environmental and social policy design. J. Ind. Ecol. 25, 95–113 (2021).Zhang, Q. et al. Transboundary health impacts of transported global air pollution and international trade. Nature 543, 705–709 (2017).Article 
    CAS 

    Google Scholar 
    Hoekstra, A. Y., Mekonnen, M. M., Chapagain, A. K., Mathews, R. E. & Richter, B. D. Global monthly water scarcity: blue water footprints versus blue water availability. PLoS ONE 7, e32688 (2012).Article 
    CAS 

    Google Scholar 
    IPCC Climate Change 2007: The Physical Science Basis (eds Solomon, S. et al.) (Cambridge Univ. Press, 2007).Schmidt, S. et al. Understanding GHG emissions from Swedish consumption—current challenges in reaching the generational goal. J. Clean. Prod. 212, 428–437 (2019).Article 

    Google Scholar 
    Huijbregts, M. A. J. Priority Assessment of Toxic Substances in the Frame of LCA. Development and Application of the Multi-Media Fate, Exposure and Effect Model USES-LCA (Interfaculty Department of Envrionmental Science, 1999).Huijbregts, M. A. J. Priority Assessment of Toxic Substances in the Frame of LCA. Time Horizon Dependency in Toxicity Potentials Calculated with the Multi-Media Fate, Exposure and Effects Model USES-LCA (Institute for Biodiversity and Ecosystem Dynamics, 2000).International Reference Life Cycle Data System (ILCD) Handbook (Publications Office EU, 2011).Verones, F., Moran, D., Stadler, K., Kanemoto, K. & Wood, R. Resource footprints an d their ecosystem consequences. Sci. Rep. 7, 40743 (2017).Chaudhary, A., Pfister, S. & Hellweg, S. Spatially explicit analysis of biodiversity loss due to global agriculture, pasture and forest land use from a producer and consumer perspective. Environ. Sci. Technol. 50, 3928–3936 (2016).Article 
    CAS 

    Google Scholar 
    Chaudhary, A., Verones, F., De Baan, L. & Hellweg, S. Quantifying land use impacts on biodiversity: combining species-area models and vulnerability indicators. Environ. Sci. Technol. 49, 9987–9995 (2015).Article 
    CAS 

    Google Scholar 
    Marquardt, S. G. et al. Consumption-based biodiversity footprints—do different indicators yield different results? Ecol. Indic. 103, 461–470 (2019).Article 

    Google Scholar 
    World Development Indicators DataBank (World Bank, 2022); https://databank.worldbank.org/source/world-development-indicatorsWorld Population Prospects 2022 (United Nations, 2022); https://population.un.org/wpp/Natural Earth Vector (Natural Earth, 2022); https://www.naturalearthdata.com/Lahti, L., Huovari, J., Kainu, M. & Biecek, P. Retrieval and analysis of eurostat open data with the Eurostat package. R J. 9, 385–392 (2017).Castellani, V., Beylot, A. & Sala, S. Environmental impacts of household consumption in Europe: comparing process-based LCA and environmentally extended input-output analysis. J. Clean. Prod. 240, 117966 (2019).Article 

    Google Scholar  More

  • in

    Climate change threatens olive oil production in the Levant

    Liphschitz, N., Gophna, R., Hartman, M. & Biger, G. The beginning of olive (Olea europaea) cultivation in the Old World: a reassessment. J. Archaeol. Sci. 18, 441–453 (1991).Article 

    Google Scholar 
    Blondel, J. & Aronson, J. Biology and Wildlife of the Mediterranean Region (Oxford Univ. Press, 1999).Fall, P. L., Falconer, S. E. & Lines, L. Agricultural intensification and the secondary products revolution along the Jordan Rift. Hum. Ecol. 30, 445–482 (2002).Article 

    Google Scholar 
    Terral, J.-F. et al. Historical biogeography of olive domestication (Olea europaea L.) as revealed by geometrical morphometry applied to biological and archaeological material. J. Biogeogr. 31, 63–77 (2004).Article 

    Google Scholar 
    Chartzoulakis, K. Salinity and olive: growth, salt tolerance, photosynthesis and yield. Agric. Water Manag. 78, 108–121 (2005).Article 

    Google Scholar 
    Vossen, P. Olive oil: history, production, and characteristics of the world’s classic oils. HortScience 42, 1093–1100 (2007).Article 

    Google Scholar 
    Kaniewski, D. et al. Primary domestication and early uses of the emblematic olive tree: palaeobotanical, historical and molecular evidence from the Middle East. Biol. Rev. 87, 885–899 (2012).Article 

    Google Scholar 
    Langgut, D. et al. The origin and spread of olive cultivation in the Mediterranean Basin: the fossil pollen evidence. Holocene 29, 902–922 (2019).Article 

    Google Scholar 
    IPCC. AR5 Synthesis Report: Climate Change 2014 https://www.ipcc.ch/report/ar5/syr/ (IPCC, 2014).IPCC. IPCC WGII Sixth Assessment Report. Cross-Chapter Paper 4: Mediterranean Region https://www.ipcc.ch/report/sixth-assessment-report-working-group-ii/ (IPCC, 2022).Fischer, E. M. & Schär, C. Consistent geographical patterns of changes in high-impact European heatwaves. Nat. Geosci. 3, 398–403 (2010).Article 
    CAS 

    Google Scholar 
    Cramer, W. et al. Climate change and interconnected risks to sustainable development in the Mediterranean. Nat. Clim. Change 8, 972–980 (2018).Article 

    Google Scholar 
    Santos, J. A., Costa, R. & Fraga, H. Climate change impacts on thermal growing conditions of main fruit species in Portugal. Clim. Change 140, 273–286 (2017).Article 

    Google Scholar 
    Orlandi, F. et al. Impact of climate change on olive crop production in Italy. Atmosphere 11, 595 (2020).Article 

    Google Scholar 
    Rodríguez Sousa, A. A., Barandica, J. M., Aguilera, P. A. & Rescia, A. J. Examining potential environmental consequences of climate change and other driving forces on the sustainability of Spanish olive groves under a socio-ecological approach. Agriculture 10, 509 (2020).Article 

    Google Scholar 
    Besnard, G. et al. The complex history of the olive tree: from Late Quaternary diversification of Mediterranean lineages to primary domestication in the northern Levant. Proc. R. Soc. B 280, 20122833 (2013).Article 
    CAS 

    Google Scholar 
    Besnard, G., Terral, J. F. & Cornille, A. On the origins and domestication of the olive: a review and perspectives. Ann. Bot. 121, 385–403 (2018).Article 

    Google Scholar 
    Bartolini, G., Prevost, G., Messeri, C., Carignani, C. & Menini, U. G. Olive Germplasm: Cultivars and World-wide Collections (FAO, 1998).Zohary, D. & Spiegel-Roy, P. Beginnings of fruit growing in the Old World. Science 187, 319–327 (1975).Article 
    CAS 

    Google Scholar 
    Terral, J.-F. Wild and cultivated olive (Olea europaea L.): a new approach to an old problem using inorganic analyses of modern wood and archaeological charcoal. Rev. Palaeobot. Palynol. 91, 383–397 (1996).Article 

    Google Scholar 
    Carrión, Y., Ntinou, M. & Badal, E. Olea europaea L. in the North Mediterranean basin during the Pleniglacial and the Early–Middle Holocene. Quat. Sci. Rev. 29, 952–968 (2010).Article 

    Google Scholar 
    Zohary, M. Plants of the Bible (Cambridge Univ. Press, 1982).Galili, E., Weinstein-Evron, M. & Zohary, D. Appearance of olives in submerged Neolithic sites along the Carmel Coast. J. Isr. Plant Sci. 22, 95–97 (1989).
    Google Scholar 
    Galili, E., Stanley, D. J., Sharvit, J. & Weinstein-Evron, M. Evidence for earliest olive-oil production in submerged settlements off the Carmel Coast, Israel. J. Archaeol. Sci. 24, 1141–1150 (1997).Article 

    Google Scholar 
    Galili, E. et al. Early production of table olives at a mid-7th millennium BP submerged site off the Carmel Coast (Israel). Sci. Rep. 11, 2218 (2021).Article 
    CAS 

    Google Scholar 
    Fraga, H., Pinto, J. G., Viola, F. & Santos, J. A. Climate change projections for olive yields in the Mediterranean Basin. Int. J. Climatol. 40, 769–781 (2020).Article 

    Google Scholar 
    Ben Zaied, Y. & Zouabi, O. Impacts of climate change on Tunisian olive oil output. Clim. Change 139, 535–549 (2016).Article 

    Google Scholar 
    Brito, C., Dinis, L. T., Moutinho-Pereire, J. & Correia, C. M. Drought stress effects and olive tree acclimation under a changing climate. Plants 8, 232 (2019).Article 
    CAS 

    Google Scholar 
    Fraga, H., Moriondo, M., Leolini, L. & Santos, J. A. Mediterranean olive orchards under climate change: a review of future impacts and adaptation strategies. Agronomy 11, 56 (2021).Article 

    Google Scholar 
    Trærup, S. & Stephan, J. Technologies for adaptation to climate change. Examples from the agricultural and water sectors in Lebanon. Clim. Change 131, 435–449 (2015).Article 

    Google Scholar 
    Chalak, L. et al. Extent of the genetic diversity in Lebanese olive (Olea europaea L.) trees: a mixture of an ancient germplasm with recently introduced varieties. Genet. Resour. Crop. Evol. 62, 621–633 (2015).Article 

    Google Scholar 
    Bou-Zeid, E. & El-Fadel, M. Climate change and water resources in Lebanon and the Middle East. J. Water Resour. Plan. Manag. 128, 343–355 (2002).Article 

    Google Scholar 
    Ramadan, H. H., Beighley, R. E. & Ramamurthy, A. S. Sensitivity analysis of climate change impact on the hydrology of the Litani Basin in Lebanon. Int. J. Environ. Pollut. 52, 65–81 (2013).Article 
    CAS 

    Google Scholar 
    Saade, J., Atieh, M., Ghanimeh, S. & Golmohammadi, G. Modeling impact of climate change on surface water availability using SWAT model in a semi-arid basin: case of El Kalb River, Lebanon. Hydrology 8, 134 (2021).Article 

    Google Scholar 
    Halwani, J. & Halwani, B. in Climate Change in the Mediterranean and Middle Eastern Region (eds Filho, W. L. & Manolas, E.) 395–412 (Springer, 2022).Aubet, M.E. in Nomads of the Mediterranean: Trade and Contact in the Bronze and Iron Ages (eds Gilboa, A. & Yasur-Landau, A.) 14–30 (Brill, 2020).Bikai, P. M. The Pottery of Tyre (Aris & Phillips, 1979).Hajar, L., Khater, C. & Cheddadi, R. Vegetation changes during the late Pleistocene and Holocene in Lebanon: a pollen record from the Bekaa Valley. Holocene 18, 1089–1099 (2008).Article 

    Google Scholar 
    Hajar, L., Haïdar-Boustani, M., Khater, C. & Cheddadi, R. Environmental changes in Lebanon during the Holocene: man vs. climate impacts. J. Arid. Environ. 74, 746–755 (2010).Article 

    Google Scholar 
    Cheddadi, R. & Khater, C. Climate change since the last glacial period in Lebanon and the persistence of Mediterranean species. Quat. Sci. Rev. 150, 146–157 (2016).Article 

    Google Scholar 
    Ozturk, M. et al. An overview of olive cultivation in Turkey: botanical features, eco-physiology and phytochemical aspects. Agronomy 11, 295 (2021).Article 
    CAS 

    Google Scholar 
    Lionello, P., Congedi, L., Reale, M., Scarascia, L. & Tanzarella, A. Sensitivity of typical Mediterranean crops to past and future evolution of seasonal temperature and precipitation in Apulia. Reg. Environ. Change 14, 2025–2038 (2014).Article 

    Google Scholar 
    Arenas-Castro, S., Gonçalves, J. F., Moreno, M. & Villar, R. Projected climate changes are expected to decrease the suitability and production of olive varieties in southern Spain. Sci. Total Environ. 709, 136161 (2020).Article 
    CAS 

    Google Scholar 
    Mechri, B., Tekaya, M., Hammami, M. & Chehab, H. Effects of drought stress on phenolic accumulation in greenhouse-grown olive trees (Olea europaea). Biochem. Syst. Ecol. 92, 104112 (2020).Article 
    CAS 

    Google Scholar 
    Pedan, V., Popp, M., Rohn, S., Nyfeler, M. & Bongartz, A. Characterization of phenolic compounds and their contribution to sensory properties of olive oil. Molecules 24, 2041 (2019).Article 
    CAS 

    Google Scholar 
    Dias, M. C., Pinto, D. C. G. A., Figueiredo, C., Santos, C. & Silva, A. M. S. Phenolic and lipophilic metabolite adjustments in Olea europaea (olive) trees during drought stress and recovery. Phytochemistry 185, 112695 (2021).Article 
    CAS 

    Google Scholar 
    Peres, F. et al. Phenolic compounds of ‘Galega Vulgar’ and ‘Cobrançosa’ olive oils along early ripening stages. Food Chem. 211, 51–58 (2016).Article 
    CAS 

    Google Scholar 
    Tsimidou, M. Z. in Handbook of Olive Oil: Analysis and Properties (eds Aparicio, R. & Harwood, J.) 311–333 (Springer, 2013).Valente, S. et al. Modulation of phenolic and lipophilic compounds of olive fruits in response to combined drought and heat. Food Chem. 329, 127191 (2020).Article 
    CAS 

    Google Scholar 
    WCRP. World Research Climate Program https://www.wcrp-climate.org/wgcm-cmip/wgcm-cmip6 (WCRP, 2022).Rallo, L. et al. in Advances in Plant Breeding Strategies: Fruits (eds Al-Khayri, J. et al.) (Springer, 2018).Abou-Saaid, O. et al. Statistical approach to assess chill and heat requirements of olive tree based on flowering date and temperatures data: towards selection of adapted cultivars to global warming. Agronomy 12, 2975 (2022).Article 

    Google Scholar 
    Faegri, K. & Iversen, I. Textbook of Pollen Analysis 4th edn. (Wiley, 1989).Ferrara, G., Camposeo, S., Palasciano, M. & Godini, A. Production of total and stainable pollen grains in Olea europaea L. Grana 46, 85–90 (2007).Article 

    Google Scholar 
    Kaniewski, D. et al. Wild or cultivated Olea europaea L. in the eastern Mediterranean during the Middle–Late Holocene? A pollen-numerical approach. Holocene 19, 1039–1047 (2009).Article 

    Google Scholar 
    R Core Team. R: A Language and Environment for Statistical Computing https://www.R-project.org/ (R Foundation for Statistical Computing, 2020).Hammer, O. & Harper, D. Paleontological Data Analysis (Blackwell, 2006).Cheddadi, R. et al. Microrefugia, climate change, and conservation of Cedrus atlantica in the Rif Mountains, Morocco. Front. Ecol. Evol. 5, 114 (2017).Article 

    Google Scholar 
    Kaniewski, D. et al. Cold and dry outbreaks in the eastern Mediterranean 3200 years ago. Geology 47, 933–937 (2019).Article 

    Google Scholar 
    Kaniewski, D. et al. Recent anthropogenic climate change exceeds the rate and magnitude of natural Holocene variability on the Balearic Islands. Anthropocene 32, 100268 (2020).Article 

    Google Scholar 
    Kaniewski, D. et al. Coastal submersions in the north-eastern Adriatic during the last 5200 years. Glob. Planet. Change 204, 103570 (2021).Article 

    Google Scholar 
    Hijmans, R. J., Cameron, S. E., Parra, J. L., Jones, P. G. & Jarvis, A. Very high-resolution interpolated climate surfaces for global land areas. Int. J. Climatol. 25, 1965–1978 (2005).Article 

    Google Scholar 
    Akima, H. & Gebhardt, A. Akima: Interpolation of Irregularly and Regularly Spaced Data. R v.0.6-2 (R Foundation for Statistical Computing, 2016).Ooms, J. D., Debroy, S., Wickham, H. & Horner, J. RMySQL: Database Interface and ‘MySQL’ Driver for R. R v.0.10.18 (R Foundation for Statistical Computing, 2019).Harris, I., Jones, P. D., Osborn, T. J. & Lister, D. H. Updated high resolution grids of monthly climatic observations – the CRU TS3.10 Dataset. Int. J. Climatol. 34, 623–642 (2014).Article 

    Google Scholar  More

  • in

    Evaluating sea cucumbers as extractive species for benthic bioremediation in mussel farms

    Avdelas, L. et al. The decline of mussel aquaculture in the European Union: Causes, economic impacts and opportunities. Rev. Aquac. 13, 91–118. https://doi.org/10.1111/raq.12465 (2021).Article 

    Google Scholar 
    Tamburini, E., Turolla, E., Fano, E. A. & Castaldelli, G. Sustainability of Mussel (Mytilus galloprovincialis) farming in the Po River delta, northern Italy, based on a life cycle assessment approach. Sustainability 12, 3814. https://doi.org/10.3390/su12093814 (2020).Article 
    CAS 

    Google Scholar 
    Shumway, S. E. et al. Shellfish aquaculture-In praise of sustainable economies and environments. J. World Aquacult. Soc. 34, 8–10 (2003).
    Google Scholar 
    Musella, M. et al. Tissue-scale microbiota of the Mediterranean mussel (Mytilus galloprovincialis) and its relationship with the environment. Sci. Total Environ. 717, 137209. https://doi.org/10.1016/J.SCITOTENV.2020.137209 (2020).Article 
    ADS 
    CAS 

    Google Scholar 
    Peharda, M., Župan, I., Bavčević, L., Frankić, A. & Klanjšček, T. Growth and condition index of mussel Mytilus galloprovincialis in experimental integrated aquaculture. Aquac. Res. 38, 1714–1720. https://doi.org/10.1111/J.1365-2109.2007.01840.X (2007).Article 

    Google Scholar 
    Sarà, G., Zenone, A. & Tomasello, A. Growth of Mytilus galloprovincialis (Mollusca, bivalvia) close to fish farms: A case of integrated multi-trophic aquaculture within the Tyrrhenian sea. Hydrobiologia 636, 129–136. https://doi.org/10.1007/S10750-009-9942-2/TABLES/4 (2009).Article 

    Google Scholar 
    Danovaro, R., Gambi, C., Luna, G. M. & Mirto, S. Sustainable impact of mussel farming in the Adriatic Sea (Mediterranean Sea): Evidence from biochemical, microbial and meiofaunal indicators. Mar. Pollut. Bull. 49, 325–333. https://doi.org/10.1016/j.marpolbul.2004.02.038 (2004).Article 
    CAS 

    Google Scholar 
    Tancioni, L. et al. Anthropogenic threats to fish of interest in aquaculture: Gonad intersex in a wild population of thinlip grey mullet Liza ramada (Risso, 1827) from a polluted estuary in central Italy. Aquac. Res. 47(5), 1670–1674 (2016).Article 

    Google Scholar 
    Chary, K. et al. Integrated multi-trophic aquaculture of red drum (Sciaenops ocellatus) and sea cucumber (Holothuria scabra): Assessing bioremediation and life-cycle impacts. Aquaculture 516, 734621. https://doi.org/10.1016/j.aquaculture.2019.734621 (2020).Article 
    CAS 

    Google Scholar 
    Purcell, S. W., Williamson, D. H. & Ngaluafe, P. Chinese market prices of beche-de-mer: Implications for fisheries and aquaculture. Mar. Policy 91, 58–65. https://doi.org/10.1016/j.marpol.2018.02.005 (2018).Article 

    Google Scholar 
    Morroni, L. et al. Sea cucumber Holothuria polii (Delle Chiaje, 1823) as new model for embryo bioassays in ecotoxicological studies. Chemosphere 240, 124819. https://doi.org/10.1016/j.chemosphere.2019.124819 (2020).Article 
    ADS 
    CAS 

    Google Scholar 
    Uthicke, S. & Karez, R. Sediment patch selectivity in tropical sea cucumbers (Holothuroidea: Aspidochirotida) analysed with multiple choice experiments. J. Exp. Mar. Biol. Ecol. 236, 69–87. https://doi.org/10.1016/S0022-0981(98)00190-7 (1999).Article 

    Google Scholar 
    MacTavish, T., Stenton-Dozey, J., Vopel, K. & Savage, C. Deposit-feeding sea cucumbers enhance mineralization and nutrient cycling in organically-enriched coastal sediments. PLoS ONE 7, 1–11. https://doi.org/10.1371/journal.pone.0050031 (2012).Article 
    CAS 

    Google Scholar 
    Rakaj, A. et al. Towards sea cucumbers as a new model in embryo-larval bioassays: Holothuria tubulosa as test species for the assessment of marine pollution. Sci. Total Environ. 787, 147593. https://doi.org/10.1016/j.scitotenv.2021.147593 (2021).Article 
    ADS 
    CAS 

    Google Scholar 
    Purcell, S., Conand, C., Uthicke, S. & Byrne, M. Ecological roles of exploited sea cucumbers. Oceanogr. Mar. Biol. 54, 367–386. https://doi.org/10.1201/9781315368597-8 (2016).Article 

    Google Scholar 
    Zamora, L. N., Yuan, X., Carton, A. G., Slater, M. J. & Marine, L. Role of deposit-feeding sea cucumbers in integrated multitrophic aquaculture: Progress, problems, potential and future challenges. Rev. Aquac. 10, 57–74. https://doi.org/10.1111/raq.12147 (2016).Article 

    Google Scholar 
    Slater, M. J. & Carton, A. G. Survivorship and growth of the sea cucumber Australostichopus (Stichopus) mollis (Hutton 1872) in polyculture trials with green-lipped mussel farms. Aquaculture 272, 389–398. https://doi.org/10.1016/j.aquaculture.2007.07.230 (2007).Article 

    Google Scholar 
    Slater, M. J. & Carton, A. G. Effect of sea cucumber (Australostichopus mollis) grazing on coastal sediments impacted by mussel farm deposition. Mar. Pollut. Bull. 58, 1123–1129. https://doi.org/10.1016/j.marpolbul.2009.04.008 (2009).Article 
    CAS 

    Google Scholar 
    Slater, M. J. & Carton, A. G. Sea cucumber habitat differentiation and site retention as determined by intraspecific stable isotope variation. Aquac. Res. 41, 695–702. https://doi.org/10.1111/j.1365-2109.2010.02607.x (2010).Article 
    CAS 

    Google Scholar 
    Stenton-Dozey, J. Finding hidden treasure in aquaculture waste. Water Atmos. 15, 9–11 (2007).
    Google Scholar 
    Slater, M. J., Jeffs, A. G. & Carton, A. G. The use of the waste from green-lipped mussels as a food source for juvenile sea cucumber, Australostichopus mollis. Aquaculture 292, 219–224. https://doi.org/10.1016/j.aquaculture.2009.04.027 (2009).Article 

    Google Scholar 
    Stenton-Dozey, J. & Heath, P. A first for New Zealand: Culturing our endemic sea cucumber for overseas markets. Water Atmos. 17, 20–21 (2009).
    Google Scholar 
    Zamora, L. N. & Jeffs, A. G. Feeding, selection, digestion and absorption of the organic matter from mussel waste by juveniles of the deposit-feeding sea cucumber, Australostichopus mollis. Aquaculture 317, 223–228. https://doi.org/10.1016/j.aquaculture.2011.04.011 (2011).Article 

    Google Scholar 
    Zamora, L. N. & Jeffs, A. G. The ability of the deposit-feeding sea cucumber Australostichopus mollis to use natural variation in the biodeposits beneath mussel farms. Aquaculture 326, 116–122. https://doi.org/10.1016/J.AQUACULTURE.2011.11.015 (2012).Article 

    Google Scholar 
    Zamora, L. N. & Jeffs, A. G. A Review of the research on the Australasian Sea Cucumber, Australostichopus mollis (Echinodermata: Holothuroidea) (Hutton 1872), with emphasis on aquaculture. J. Shellfish Res. 32, 613–627. https://doi.org/10.2983/035.032.0301 (2013).Article 

    Google Scholar 
    Zamora, L. N. & Jeffs, A. G. Macronutrient selection, absorption and energy budget of juveniles of the Australasian sea cucumber, Australostichopus mollis, feeding on mussel biodeposits at different temperatures. Aquac. Nutr. 21, 162–172. https://doi.org/10.1111/ANU.12144 (2015).Article 
    CAS 

    Google Scholar 
    Chatzivasileiou, D. et al. An IMTA in Greece: Co-culture of fish, bivalves, and holothurians. J. Mar. Sci. Eng. 10, 776. https://doi.org/10.3390/jmse10060776 (2022).Article 

    Google Scholar 
    Rakaj, A. et al. Spawning and rearing of Holothuria tubulosa: A new candidate for aquaculture in the Mediterranean region. Aquac. Res. 49, 557–568. https://doi.org/10.1111/are.13487 (2018).Article 
    CAS 

    Google Scholar 
    Rakaj, A., Fianchini, A., Boncagni, P., Scardi, M. & Cataudella, S. Artificial reproduction of Holothuria polii: A new candidate for aquaculture. Aquaculture 498, 444–453. https://doi.org/10.1016/j.aquaculture.2018.08.060 (2019).Article 

    Google Scholar 
    González-Wangüemert, M., Aydin, M. & Conand, C. Assessment of sea cucumber populations from the Aegean Sea (Turkey): First insights to sustainable management of new fisheries. Ocean Coast. Manag. 92, 87–94. https://doi.org/10.1016/J.OCECOAMAN.2014.02.014 (2014).Article 

    Google Scholar 
    González-Wangüemert, M., Valente, S. & Aydin, M. Effects of fishery protection on biometry and genetic structure of two target sea cucumber species from the Mediterranean Sea. Hydrobiologia 743, 65–74. https://doi.org/10.1007/s10750-014-2006-2 (2015).Article 

    Google Scholar 
    González-Wangüemert, M., Domínguez-Godino, J. A. & Cánovas, F. The fast development of sea cucumber fisheries in the Mediterranean and NE Atlantic waters: From a new marine resource to its over-exploitation. Ocean Coast. Manag. 151, 165–177. https://doi.org/10.1016/j.ocecoaman.2017.10.002 (2018).Article 

    Google Scholar 
    González-Wangüemert, M. & Godino, J. Sea cucumbers as new marine resource in Europe. Front. Mar. Sci. 3, 112 (2016).
    Google Scholar 
    Domínguez-Godino, J. A., Slater, M. J., Hannon, C. & González-Wangüermert, M. A new species for sea cucumber ranching and aquaculture: Breeding and rearing of Holothuria arguinensis. Aquaculture 438, 122–128. https://doi.org/10.1016/J.AQUACULTURE.2015.01.004 (2015).Article 

    Google Scholar 
    Günay, D., Emiroğlu, D., Tolon, T., Özden, O. & Saygi, H. Growth and survival rate of Juvenile Sea Cucumbers (Holothuria tubulosa, Gmelin, 1788) at Various Temperatures. Turk. J. Fish. Aquat. Sci. 15, 533–541. https://doi.org/10.4194/1303-2712-v15_2_41 (2015).Article 

    Google Scholar 
    Tolon, T. Effect of salinity on growth and survival of the juvenile sea cucumbers Holothuria tubulosa (Gmelin, 1788) and Holothuria poli (Delle Chiaje, 1923). Fresenius Environ. Bull. 26, 3930–3935 (2017).CAS 

    Google Scholar 
    Tolon, T., Emiroğlu, D., Günay, D. & Hancı, B. Effect of stocking density on growth performance of juvenile sea cucumber Holothuria tubulosa (Gmelin, 1788). Aquac. Res. 48, 4124–4131. https://doi.org/10.1111/are.13232 (2017).Article 

    Google Scholar 
    Tolon, M. T., Emiroglu, D., Gunay, D. & Ozgul, A. Sea cucumber (Holothuria tubulosa Gmelin, 1790) culture under marine fish net cages for potential use in integrated multi-trophic aquaculture (IMTA). Indian J. Geol. Mar. Sci. 46, 749–756 (2017).
    Google Scholar 
    Neofitou, N. et al. Contribution of sea cucumber Holothuria tubulosa on organic load reduction from fish farming operation. Aquaculture 501, 97–103. https://doi.org/10.1016/j.aquaculture.2018.10.071 (2019).Article 

    Google Scholar 
    Sadoul, B. et al. Aquaculture Is Holothuria tubulosa the golden goose of ecological aquaculture in the Mediterranean Sea? Aquaculture 554, 738149. https://doi.org/10.1016/j.aquaculture.2022.738149 (2022).Article 
    CAS 

    Google Scholar 
    Cutajar, K. et al. Culturing the sea cucumber Holothuria poli in open-water integrated multi-trophic aquaculture at a coastal Mediterranean fish farm. Aquaculture 550, 737881. https://doi.org/10.1016/j.aquaculture.2021.737881 (2022).Article 
    CAS 

    Google Scholar 
    Grosso, L. et al. Integrated Multi-Trophic Aquaculture (IMTA) system combining the sea urchin Paracentrotus lividus, as primary species, and the sea cucumber Holothuria tubulosa as extractive species. Aquaculture 534, 736268. https://doi.org/10.1016/J.AQUACULTURE.2020.736268 (2021).Article 
    CAS 

    Google Scholar 
    González-Wangüemert, M., Valente, S., Henriques, F., Domínguez-Godino, J. A. & Serrão, E. A. Setting preliminary biometric baselines for new target sea cucumbers species of the NE Atlantic and Mediterranean fisheries. Fish. Res. 179, 57–66. https://doi.org/10.1016/J.FISHRES.2016.02.008 (2016).Article 

    Google Scholar 
    Aydin, M. Biometry, density and the biomass of the commercial sea cucumber population of the Aegean Sea. Turk. J. Fish. Aquat. Sci 19, 463–474. https://doi.org/10.4194/1303-2712-v19_6_02 (2018).Article 

    Google Scholar 
    Whitlock, M. C. & Schluter, D. Analisi Statistica dei Dati Biologici, Zanichelli (2010)Hammer, O. & Harper, D. A. T. PAST PAleontological STatistics Version 3 Reference Manual (University of Oslo, 2013).Zhou, Y. et al. Feeding and growth on bivalve biodeposits by the deposit feeder Stichopus japonicus Selenka (Echinodermata: Holothuroidea) co-cultured in lantern nets. Aquaculture 256, 510–520. https://doi.org/10.1016/j.aquaculture.2006.02.005 (2006).Article 

    Google Scholar 
    Pensa, D. et al. Population status, distribution and trophic implications of Pinna nobilis along the South-eastern Italian coast. Npj Biodivers. https://doi.org/10.21203/rs.3.rs-1425249/v1 (2022).Article 

    Google Scholar 
    Francour, P. Predation on holothurians: A literature review. Invert. Bio. 116, 52–60. https://doi.org/10.2307/3226924 (1997).Article 

    Google Scholar 
    Mecheta, A. & Mezali, K. A biometric study to determine the economic and nutritional value of sea cucumbers (Holothuroidea: Echinodermata) collected from Algeria’s shallow water areas. Beche-de-mer Inf. Bull. 39, 65–70 (2019).
    Google Scholar 
    Sun, J., Hamel, J. F., Gianasi, B. L., Graham, M. & Mercier, A. Growth, health and biochemical composition of the sea cucumber Cucumaria frondosa after multi-year holding in effluent waters of land-based salmon culture. Aquac. Environ. Interact. 12, 139–151. https://doi.org/10.3354/aei00356 (2020).Article 

    Google Scholar 
    Boncagni, P., Rakaj, A., Fianchini, A. & Vizzini, S. Preferential assimilation of seagrass detritus by two coexisting Mediterranean sea cucumbers: Holothuria polii and Holothuria tubulosa. Estuar. Coast. Shelf Sci. 231, 106464. https://doi.org/10.1016/j.ecss.2019.106464 (2019).Article 
    CAS 

    Google Scholar 
    Rakaj, A., and Fianchini, A. Mediterranean sea cucumbers—Biology, ecology, and exploitation, Chapter. In The World of Sea Cucumbers Challenges, Advances, and Innovations (Mercier, A., Hamel, J.-F, Suhrbier, A. & Pearce, C.) (2023)Massin, C. & Jangoux, M. Observations écologiques sur Holothuria tubulosa, Holothuria poli et Holothuria forskali (Echinodermata, Holothuroidea) et comportement alimentaire de H. tubulosa. Référ. Cah. Biol. Mar. 17, 45–59 (1976).
    Google Scholar 
    Coulon, P. & Jangoux, M. Feeding rate and sediment reworking by the holothuroid Holothuria tubulosa (Echinodermata) in a Mediterranean seagrass bed off Ischia Island, Italy. Mar. Ecol. Progr. Ser. 92, 201–204 (1993).Article 
    ADS 

    Google Scholar 
    Belbachir, N., Mezali, K. & Soualili, D. L. Selective feeding behaviour in some aspidochirotid holothurians (Echinodermata: Holothuroidea) at Stidia, Mostaganem Province, Algeria (2014).Grosso, L. et al. Trophic requirements of the sea urchin Paracentrotus lividus varies at different life stages: comprehension of species ecology and implications for effective feeding formulations. Front. Mar. Sci. 9, 865450. https://doi.org/10.3389/fmars.2022.865450 (2022).Article 

    Google Scholar 
    Sun, Z. L., Gao, Q. F., Dong, S. L., Shin, P. K. & Wang, F. Estimates of carbon turnover rates in the sea cucumber Apostichopus japonicus (Selenka) using stable isotope analysis: The role of metabolism and growth. Mar. Ecol. Prog. Ser. 457, 101–112. https://doi.org/10.3354/meps09760 (2012).Article 
    ADS 

    Google Scholar 
    Yuan, X. T. et al. Effects of aestivation on the energy budget of sea cucumber Apostichopus japonicus (Selenka) (Echinaodermata: Holothuroidea). Acta. Ecol. Sin. 27, 3155−3161. https://doi.org/10.1016/S1872-2032(07)60070-5 (2007).Article 

    Google Scholar 
    Liu, Y., Dong, S. L., Tian, X. L., Wang, F. & Gao, Q. F. Effects ofdietary sea mud and yellow soil on growth and energybudget of the sea cucumber Apostichopus japonicas (Selenka). Aquaculture 286, 266–270. https://doi.org/10.1016/j.aquaculture.2008.09.029 (2009).Article 

    Google Scholar 
    Brown, N. P. & Eddy, S. D. Echinoderm Aquaculture (Wiley, 2015).Book 

    Google Scholar 
    Qiu, T., Zhang, L., Zhang, T., Bai, Y. & Yang, H. Effect of culture methods on individual variation in the growth of sea cucumber Apostichopus japonicus within a cohort and family. Chin. J. Oceanol. Limnol. 32, 737–742. https://doi.org/10.1007/S00343-014-3131-5 (2014).Article 
    ADS 

    Google Scholar 
    Zappes, I. A. et al. New data on Weddell seal (Leptonychotes weddellii) colonies: A genetic analysis of a top predator from the Ross Sea, Antarctica. PLoS ONE 12, 0182922. https://doi.org/10.1371/journal.pone.0182922 (2017).Article 
    CAS 

    Google Scholar 
    Paltzat, D. L., Pearce, C. M., Barnes, P. A. & McKinley, R. S. Growth and production of California sea cucumbers (Parastichopus californicus Stimpson) co-cultured with suspended Pacific oysters (Crassostrea gigas Thunberg). Aquaculture 275, 124–137. https://doi.org/10.1016/j.aquaculture.2007.12.014 (2008).Article 

    Google Scholar 
    Dong, S. et al. Intra-specific effects of sea cucumber (Apostichopus japonicus) with reference to stocking density and body size. Aquac. Res. 41, 1170–1178. https://doi.org/10.1111/J.1365-2109.2009.02404.X (2010).Article 

    Google Scholar 
    Pei, S., Dong, S., Wang, F., Gao, Q. & Tian, X. Effects of stocking density and body physical contact on growth of sea cucumber, Apostichopus japonicus. Aquac. Res. 45, 629–636. https://doi.org/10.1111/ARE.12004 (2014).Article 

    Google Scholar 
    Xia, B., Ren, Y., Wang, J., Sun, Y. & Zhang, Z. Effects of feeding frequency and density on growth, energy budget and physiological performance of sea cucumber Apostichopus japonicus (Selenka). Aquaculture 466, 26–32. https://doi.org/10.1016/J.AQUACULTURE.2016.09.039 (2017).Article 

    Google Scholar 
    Domínguez-Godino, J. A. & González-Wangüemert, M. Does space matter? Optimizing stocking density of Holothuria arguinensis and Holothuria mammata. Aquac. Res. 49, 3107–3115. https://doi.org/10.1111/are.13773 (2018).Article 

    Google Scholar 
    Rugnini, L., Rossi, C., Antonaroli, S., Rakaj, A. & Bruno, L. The influence of light and nutrient starvation on morphology, biomass and lipid content in seven strains of green microalgae as a source of biodiesel. Microorganisms 8, 1254. https://doi.org/10.3390/microorganisms8081254 (2020).Article 
    CAS 

    Google Scholar  More