Biodiversity Data Journal 11: e101942 OO) doi: 10.3897/BDJ.11.e101942 open access Research Article Mitochondrial genome analysis, phylogeny and divergence time evaluation of Strix aluco (Aves, Strigiformes, Strigidae) Yeying Wangt, Haofeng Zhan*, Yu Zhang*, Zhengmin Long}, Xiaofei Yang + Guizhou Normal University, Guiyang, China § Guizhou University, Guiyang, China Corresponding author: Yeying Wang (wangyeying0818@163.com), Xiaofei Yang (xfyang3@gzu.edu.cn) Academic editor: Caio J. Carlos Received: 14 Feb 2023 | Accepted: 11 Mar 2023 | Published: 20 Mar 2023 Citation: Wang Y, Zhan H, Zhang Y, Long Z, Yang X (2023) Mitochondrial genome analysis, phylogeny and divergence time evaluation of Strix aluco (Aves, Strigiformes, Strigidae). Biodiversity Data Journal 11: e101942. https://doi.org/10.3897/BDJ.11.e101942 Abstract Background Prior research has shown that the European peninsulas were the main sources of Siérix aluco colonisation of Northern Europe during the late glacial period. However, the phylogenetic relationship and the divergence time between S. aluco from Leigong Mountain Nature Reserve, Guizhou Province, China and the Strigiformes from overseas remains unclear. The mitochondrial genome structure of birds is a covalent double-chain loop structure that is highly conserved and, thus, suitable for phylogenetic analysis. This study examined the phylogenetic relationship and divergence time of Strix using the whole mitochondrial genome of S. aluco. New information In this study, the complete mitochondrial genome of Sérix aluco, with a total length of 18,632 bp, is reported for the first time. A total of 37 genes were found, including 22 tRNAs, two rRNAs, 13 protein-coding genes and two non-coding control regions. Certain species of Tytoninae were used as out-group and PhyloSuite software was applied to build the ML- © Wang Y et al. This is an open access article distributed under the terms of the Creative Commons Attribution License (CC BY 4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited. 2 Wang Y et al tree and Bl-tree of Strigiformes. Finally, the divergence time tree was constructed using BEAST 2.6.7 software and the age of Miosurnia diurna fossil-bearing sediments (6.0—9.5 Ma) was set as internal correction point. The common ancestor of Strix was confirmed to have diverged during the Pleistocene (2.58—0.01 Ma). The combined action of the dramatic uplift of the Qinling Mountains in the Middle Pleistocene and the climate oscillation of the Pleistocene caused Strix divergence between the northern and southern parts of mainland China. The isolation of glacial-interglacial rotation and glacier refuge was the main reason for the divergence of Strix uralensis and S. aluco from their common ancestor during this period. This study provides a reference for the evolutionary history of S. a/uco. Keywords Strix aluco, phylogeny, divergence time, Pleistocene, climate oscillation, mountains uplift Introduction Strix aluco belongs to Strigidae (Strigiformes) and is a medium-sized owl (Grytsyshina et al. 2016). S. aluco is a non-migratory and territorial nocturnal bird (Sunde et al. 2003, Dona et al. 2016) with a wide distribution throughout mountainous broadleaf forest and mixed forest in Eurasia; Israel is the southernmost country in the distribution area of S. aluco in the Northern Hemisphere (Obuch 2011, Comay et al. 2022). It can feed on mammals, fish, amphibians and even small birds such as sparrows (Obuch 2017) and voles are its most preferred food (Karell et al. 2009, Solonen 2022). The IUCN listed this species as of "Least Concern". The current population trend is stable and the estimated number of individuals ranges from 1,000,000 to 2,999,999 (IUCN 2016: https:/Awww.iucnredlist.org/). In China, S. aluco has been listed as a national class II protected animal. Mitochondria are characterised by maternal inheritance, high conservation, multiple copies in each cell, low sequence recombination rate and high evolutionary rate; therefore, mitochondria are widely used in phylogenetic studies (Yan et al. 2017, Sun et al. 2020). Using them enables researchers to accurately infer phylogenetic relationships in birds (Tuinen et al. 2000) and complete mitochondrial genomes generally achieve higher accuracy than partial mitochondrial genomes (Haring et al. 2001, Harrison et al. 2004). Through skeletal comparison, Strigidae has been divided into three subfamilies: Striginae (13 genera), Surniinae (eight genera) and Asioninae (two genera) (Ford 1967). Previous studies have defined the phylogenetic position of S. a/uco using a single gene or a combination of multiple mitochondrial genes (Heidrich and Wink 1994, Fuchs et al. 2008, Wood et al. 2017, Yu et al. 2021). Earlier research identified the monophyly of the Strigiformes phylogeny through the cytochrome B (Cyt B) gene (Wink and Heidrich 2000). In Strigiformes, the taxonomic relationship of subordinate branches of Strigidae has been hotly debated (Salter et al. 2020). Phylogenetic relationships through the Cyt B gene also showed that the order Strigiformes can be divided into two groups: (Tytonidae + Strigidae), Tytonidae consisting of Tytoninae (containing Tyto) and Phodilinae (containing Phodilus); and Strigidae can be divided into Striginae, Surniinae and Ninoxinae, amongst them, Mitochondrial genome analysis, phylogeny and divergence time evaluation ... 3 Striginae can be subdivided into a clade of ((Bubo + Strix) + (Pulsatrigini + Asio)) + (Psiloscops + Megascops) + Otus, with Surniinae consisting of two branches (with Surniini and Aegolius), Surniini containing (Glaucidium + Athene) and Ninoxinae being mainly composed of Ninox, possibly including Uroglaux and Sceloglaux (Wink et al. 2009, Wink and Sauer-GUu 2021). In addition, a growing number of scholars have described a framework for Strigiformes phylogeny. Zhang et al. (2016) completed the whole mitochondrial genome sequencing of Asio flammeus and determined the paraphyletic phylogenetic relationship amongst the three genera of Ofus, Ptilopsis and Asio; Kang et al. (2018) completed the whole mitochondrial genome sequencing of Sifrix uralensis and determined the inter-genus relationship of Ofus + (Asio + (Strix + Bubo)) by studying the mitochondrial genome of Strigidae; Salter et al. (2020) combined morphological characteristics and molecular biology, suggesting that a typical owl contains Striginae and Surniinae; they further suggested that Athene, Otus, Asio, Megascops, Bubo and Strix are paraphyletic, while Ninox and Glaucidium are polyphyletic; Koparde et al. (2018) showed that the Striginae and Surniinae form a paraphyletic group in the South Asian subcontinent population with Tytonidae as the out-group; their study showed that Strigidae and Tytonidae diverged at about 42.5-47.7 Ma (mega-annum, million years); Uva et al. (2018) Clarified the global distribution of Tytonidae and their time of divergence, their analysis showing that Tytonidae and S. aluco split from a common ancestor dating back to about 45 Ma. Prior research has shown that Strix and Tyto diverged roughly about 40-50 Ma (Prum et al. 2015). The phylogenetic relationship and timing of the divergence of Sfrix in China remain unclear. There are many reasons for the divergence of species, amongst which geological and climatic influences on species diversification cannot be ignored (Claramunt and Cracraft 2015). The Cretacean-Tertiary extinction event was a mass extinction event in Earth's history that occurred 65 Ma and wiped out most animals and plants at the time, including the dinosaurs. It also wiped out the direct ancestors of tree-dwelling waterbirds on Earth today with the few survivors evolving rapidly thereafter (Field et al. 2018). Bird ancestry began to increase exponentially at the end of the Eocene, from an original 100 species to the 10,000 species of today (Ksepka and Phillips 2015). Since the late Miocene, many birds in the Palaearctic have migrated on a large scale and their changing ranges have led to gene flows that have provided opportunities for the origin of various bird subfamilies (Drovetski 2003, Holm and Svenning 2014). Climatic oscillation during the Quaternary Period, especially throughout the Pleistocene (2.58-0.01 Ma), promoted the evolution of species on a global scale (Hewitt 2000, Hewitt 2004, Lamb et al. 2019). Pleistocene glacial gyre played a positive role in speciation (Zhao et al. 2013, Hung et al. 2014, Kozma et al. 2018, Mays et al. 2018). Brito (2005) studied 14 populations of S. a/uco in Western Europe and found that these could be divided into three branches originating from three glacial sanctuaries in the Iberian Peninsula and the Balkan Peninsula in Europe. This finding supports the "glacier refuge hypothesis" that describes the origin of S. aluco in estern Europe. However, the origin and divergence of S. aluco in mainland China remain a mystery. 4 Wang Y et al Divergence time analysis can provide a reference for the evolution process of species and provides a basis for further studies. To clarify the divergence time of species, it is necessary to obtain their gene sequence first; then, an appropriate evolutionary model needs to be selected and reliably calibrated, for example, by determining the age of fossils (Ho and Phillips 2009, Ho and Duchéne 2014). To clarify the phylogenetic position, divergence time and reasons for divergence of S. aluco from China, this study sequenced the complete mitochondrial genome of S. aluco and used it (combined with the mitochondrial genome of other birds in Strigiformes) to reconstruct the phylogenetic tree of Strigiformes. Fossil data are usually used to evaluate the divergence time of birds and the divergence time of Surniinae fossils was used as the correction point to analyse the divergence time of Strix. Possible reasons for its divergence are discussed in depth. Materials and methods Sample origin and DNA extraction Part of the muscle tissue was extracted from the leg of one individual of S. a/uco that died of an unknown cause in the Rescue Center of Leigong Mountain National Nature Reserve, Qiandongnan Prefecture, Guizhou Province, China (26° 49' 26.40" N, 104° 43' 33.60" E). The sample was stored in a refrigerated box with a built-in thermometer, the temperature was kept near freezing, until the sample was transported back to the laboratory for DNA extraction. To extract DNA, the standardised CTAB method was used (Lutz et al. 2011). Sequencing and assembly The whole genome shotgun strategy was used to construct the library (Roe 2004). Next generation sequencing technology was used for paired-end sequencing, based on the Illumina NovaSeq sequencing platform (Illumina NovaSeq, Illumina Inc., San Diego, California, USA). The concentration and purity of DNA extracted from the samples were assessed by Thermo Scientific NanoDrop 2000 (Thermo Scientific NanoDrop 2000, Thermo Fisher, Massachusetts, USA) and the integrity was assessed by agarose electrophoresis (Electrophoresis apparatus of Liuyi Company, Beijing, China) and Agilent 2100 Bioanalyzer (Agilent 2100 Bioanalyzer, Agilent Corporation, California, USA). The Covairs machine (Covairs machine of BRANSON Company in St. Louis, Missouri, USA) was used to break up and fragment DNA. The gene library was constructed according to the shotgun method described by Roe (2004). The Agilent 2100 Bioanalyzer was used to assess the size of the library and fluorescence quantitative detection was used to assess the total concentration of the library. The optimal amount of the library was selected and sequenced on the Illumina NovaSeq sequencing platform. A single-stranded library was used as a template for bridge PCR amplification and sequencing was performed during synthesis. After DNA extraction, purification, library construction and sequencing, a raw image file was first obtained by sequencing. The raw data that can be read in FASTQ format were Mitochondrial genome analysis, phylogeny and divergence time evaluation ... 5 generated after the multi-step transformation, i.e. the offline data. Data transformation is automatically completed by the sequencing platform. According to the statistics of raw data, 7,947,240 reads (each sequence read is called one read) were obtained, the total number of bases was 1,192,086,000 bp, the percentage of fuzzy bases (uncertain bases) was 0.0016% and the GC content was 44.58%. The base recognition accuracy exceeding 99.00% accounted for 95.61% and the base recognition accuracy exceeding 99.90% accounted for 90.44%. The quality of off-machine data was tested through quality control and the software used is FastQC (http:/Awww.bioinformatics.babraham.ac.uk/projects/ fastqc). Sequencing data contain low-quality reads with connectors, which will greatly interfere with subsequent analysis. To ensure the quality of subsequent information analysis, Fastp software (version 0.20.0) was used to remove sequencing connectors at the 3' end. Low- quality sequences (i.e. sequences with an average Q value of less than 20 and sequences with a sequence length shorter than 50 bp) were removed. The number of high-quality reads obtained was 7,611,984, accounting for 95.78% of the raw data and the number of bases of high-quality reads was 1,123,739,765 bp, accounting for 94.27% of the raw data (Chen et al. 2018). A5-miseq v20150522 (Coil et al. 2015) and SPAdesv 3.9.0 (Bankevich et al. 2012) were used for the de novo sequencing of high-quality next-generation sequencing data. To construct contig and scaffold sequences, the sequences were extracted according to the sequencing depth of de novo splicing sequences. The sequences with high sequencing depth were compared with the NT (Nucleotide) library on NCBI by Blastn (BLAST v2.2.31+) and the mitochondrial sequences of each splicing result were selected. To integrate splicing results, the mitochondrial splicing results obtained by the different software above were combined with reference sequences. Collinearity analysis was performed using mummer v.3.1 software (Kurtz et al. 2004) to determine the position relationship between contigs and fill gaps between contigs. The results were corrected by using pilon v.1.18 software (Walker et al. 2014) to obtain the final mitochondrial sequence. The complete mitochondrial genome sequence obtained by splicing was uploaded to the MITOS web server (http://mitos2.bioinf.uni-leipzig.de/index.py) for functional annotation (Bernt et al. 2013). RefSeq 81 Metazoa was selected as reference, the genetic code was set to a second set of vertebrate codons and other parameters were set according to the default parameters proposed by MITOS. Through the above methods, the base compositions of the whole mitochondrial genome, protein-coding genes and rRNA genes were obtained. CGview visualisation software was used to draw the mitochondrial complete genome circle map (Stothard and Wishart 2005). Mitochondrial genome data collection in Strigiformes Currently (until this study), in GenBank, there are 30 species with mitochondrial genomes greater than 10,000 bp, including 27 species of Strigidae and three species of Tytonidae. All taxonomic classifications of the species follow the current version of the IOC WORLD BIRD LIST (12.2) (http://dx.doi.org/10.14344/IOC.ML.12.2). The existing sequences of 6 Wang Y et al these thirty species were stored in a local folder using GenBank format. The registration number is shown in Table 1. Construction of phylogenetic trees The PhyloSuite software (downloaded from: hitps://github.com/dongzhang0725/ PhyloSuite/releases) (Zhang et al. 2020) was used to drag the 30 GenBank format files (downloaded from NCBI) and the GenBank format files of S. aluco sequence (obtained by the sequencing in this study) into the main interface. First, following the guided steps in the literature of Zhang et al. (2020), a series of standardised operations were conducted. Mitogenome sequence types were chosen; meanwhile, the annotation error tRNA file was exported and modified comments were uploaded to the ARWEN website (http://130.235.244.92/ARVWEN/). The site of the modified comments is copied and pasted for modification after the box. After that, the corrected 13 protein-coding genes (PCGs) and 24 RNAs were extracted successfully. The second set of two vertebrate mitochondrial codes is selected here and the extracted 13 PCGs and 24 RNAs are imported into MAFFT (PhyloSuite programme) for multiple sequence alignment. The 37 gene files exported by MAFFT were selected and imported into ‘concatenate sequence’ (PhyloSuite programme), using the '-auto' strategy and ‘normal’ alignment mode. The concatenated completion file was selected and PartitionFinder 2.0 (PhyloSuite programme) (Lanfear et al. 2017) was used to perform a greedy search, select the 'Nucleitide' mode and ‘branc-lengths' can be ‘linked’. Here, 'mrbayes' was chosen for models and the model supported by MrBayes was calculated. 'AlCc' is the model selection criterion recommended by PartitionFinder authors, the optimal partitioning strategy and model selection was calculated, this place using a separate GTR+G model for each data block automatically. Table 1. Mitochondrial genome sequences used in this study. Taxon GenBank accession Size (bp) Notes Reference Aegolius funereus MN122880 17166 Partial Direct Submission Asio flammeus KP889214 18966 Complete Zhang et al. (2016) Asio otus MG916810 17555 Complete Lee et al. (2018) Athene brama KF961185 16194 Partial Direct Submission Athene noctua MN122903 15776 Partial Direct Submission Bubo blakistoni LC099106 19379 Partial Direct Submission Bubo bubo MN206975 18956 Complete Direct Submission Bubo flavipes LC099100 19447 Partial Direct Submission Bubo scandiacus MG681084 18734 Complete Kang et al. (2018) Ciccaba nigrolineata MN356178 14875 Partial Feng et al. (2020) Glaucidium brasilianum MN356303 17717 Partial Feng et al. (2020) Mitochondrial genome analysis, phylogeny and divergence time evaluation ... 7 Taxon GenBank accession Size (bp) Notes Reference Glaucidium brodiei brodiei KP684122 17318 Complete Sun et al. (2016) Glaucidium cuculoides KY092431 17392 Complete Liu et al. (2019) Ninox novaeseelandiae AY309457 16223 Complete Harrison et al. (2004) Ninox scutulata KT943750 16208 Complete Direct Submission Ninox strenua KX529654 16206 Complete Sarker et al. (2016) Otus bakkamoena KT340631 17389 Complete Park et al. (2019a) Otus lettia MW364567 16951 Complete Yu et al. (2021) Otus scops KT340630 17413 Complete Park et al. (2019b) Otus semitorques LC541473 18834 Complete Direct Submission Otus sunia MF346692 17835 Complete Zhou et al. (2019) Sceloglaux albifacies KX098448 15565 Partial Wood et al. (2017) Strix aluco MN122823 16490 Partial Direct Submission Strix aluco OP850567 18832 Complete, This study Strix leptogrammica KC953095 16307 Complete Liu et al. (2014) Strix occidentalis MF431746 19889 Complete Hanna et al. (2017) Strix uralensis MG681081 18708 Complete Kang et al. (2018) Strix varia MF431745 18975 Complete Hanna et al. (2017) Out-group Phodilus badius KF961183 17086 Complete Mahmood et al. (2014) Tyto alba EU410491 16148 Partial Pratt et al. (2009) Tyto longimembris KP893332 18466 Partial Xu et al. (2016) The result file of PartitionFinder 2.0 was selected, the ML method was completed in I|Q-tree (Minh et al. 2020) mode (PhyloSuite programme) and the BI method was completed in MrBayes mode (PhyloSuite programme). Phodilus badius, Tyto alba and Tyto longimembris were set as out-groups. In |Q-tree mode, the Edge-linked partition style was employed for 10,000 replicates of ultrafast bootstrapping (Sterli et al. 2013, Hoang et al. 2018). In MrBayes mode, the result folder of PartitionFinder 2.0 was opened, out-groups were set, parameters were defined as Partition Models and algebra run as two parallel runs, four chains, for 2,000,000 generations (where it must be ensured that the average standard deviation of split frequencies values remains below 0.01), sampling freq is one sampling run for 1000 times and 25% of the initial samples were discarded as burn-in. Divergence time evaluation Miosurnia diurna fossils provide an approximate date of the origin of Surniinae and the age of the fossil-bearing sediments of the M. diurna is 6.0-9.5 Ma (Li et al. 2022), the origin times of Surniinae were set to 6.0 Ma and 9.5 Ma, which included Aegolius, Athene, and Glaucidium (Wink and Sauer-GUti 2021). The 'NEX' file obtained by concatenating 37 genes using the ‘concatenate sequence’ programme function in PhyloSuite was imported into 8 Wang Y et al BEAUti 2.6.7, (http:/www.beast2.org/), Hasegawa-Kishino-Yano model, with four gamma categories, Strict clock with 1.0 clock rate and with a Yule process (speciation) prior. “ Aegolius funereus, Athene brama, Athene noctua, Glaucidium brasilianum, Glaucidium brodiei brodiei and Glaucidium cuculoides” (sequence file name) was chosen and Prior was added. Then, the “monophyletic” option was checked, the Mean set to 6.0/9.5 and Sigma set to 0.1. A Markov Monte Carlo Chain Bayesian analysis with a chain length of 10,000,000 and with states recorded every 1000 iterations was run using BEAST 2.6.7. Log files were assessed using TRACER 1.7.2 (http://tree.bio.ed.ac.uk/software/tracer/) to ensure that posteriors were normally distributed and that all statistics had attained effective sample sizes of > 200. If ESS < 200, optimisation was employed by adding 5,000,000 iterations (chain length) each time. A burn-in of 10% was discarded, the maximum clade credibility tree was determined and mean heights were chosen using TreeAnnotator 2.6.7. Finally, FigTree 1.4.4 was used to assess the divergence time. Finally, Adobe Illustrator 1.0.0.2 was used for visual editing (all figures are the same). Results Genome annotation The total length of the mitochondrial genome sequence was 18,632 bp (GenBank entry number: OP850567). The results of genome annotation showed that the total number of genes was 39, including 13 protein-coding genes, 22 tRNA genes, two rRNA genes, two O H genes and 0 O, genes. Amongst them were eight tRNA genes (trn-Q, trn-A, trn-N, trn-C, trn-Y, trn-P, trn-E and trn-S2) and the PCG gene nad6 on the main chain (J chain). The remaining 14 tRNA genes were trn-F, trn-V, trn-L2, trn-l, trn-M, trn-W, trn-D, trn-K, trn-G, trn-R, trn-H, trn-S1, trn-L1 and trn-T. Further, the two rRNA genes rrn-S and rrn-L were found and 12 PCGs genes encoding nad1, nad2, nad3, nad4, nad4L, nad5, atp6, atp8, cox1, cOx2, cox3 and cytb on the secondary (N) chain were also found. There was no gene rearrangement (Fig. 1). The specific annotation results of each gene are shown in Suppl. material 1. Phylogenetic analysis In this study, both the ML-tree and Bl-tree showed the same tree topology with good support. The tree showed that Strigidae and Tytonidae are two distinct lineages under the owl shape. Athene noctua is a sister group of Athene brama; Glaucidium, Athene and Aegolius constitute the same group, Aegolius funereus is closely related to (G. cuculoides + G. brasilianum), but BI/ML (posterior probability/bootstrap) is 0.97/52, the phylogenetic relationship between them is G/aucidium brodiei brodiei + ((A. funereus + (G. cuculoides + G. brasilianum)) + (Athene noctua + A. brama). Ciccaba nigrolineata is nested in Strix. It shows that BI/ML is 1/100, S. al/uco in this study is a sister group of S. uralensis, Strix aluco MN122823 + (Strix aluco OP850567 + Strix uralensis) had formed with S. aluco; Strix is a sister to Bubo clade and forms an Asio + (Strix + Bubo) monophyletic group with Asio and a higher monophyletic group with Otus + [Asio + ((Strix + Ciccaba nigrolineata) + Bubo)]. Mitochondrial genome analysis, phylogeny and divergence time evaluation ... 2) Additionally, Sceloglaux albifacies is nested in Ninox, BI/ML is 1/99; this monophyly emerged simultaneously with (Sceloglaux albifacies + Ninox) and the monophyly exhibited as dyadic taxa (Fig. 2). Wi cps i tRNA MH rRNA Other HB GC content BB GC skew+ BB GC skew- trnF(gaa) — control region \ i) a ~ ! rl trnE (ttc), nad6 trnP(tgg), PotRNA 7imnL2(taa) control region, _nad1l jornltgat) —trnQ(ttg) NtrnM(cat) trnT(tgt)~ . ~askop = Strix aluco complete mitochondrial genomics a Length: 18,632 bp = 5kop - cob— —nad2 trnW(tca) ~LtrnA(tgc) —trnN(gtt) \ trnC(gca) trnY(gta) nad5—- ~“cox1 trnL1(tag)—_ trnS1(gct)~ trnH(gtg —trnS2(tga) \trnD(gtc) \cox2 nad4“ —trnk(ttt) nad4 : am latps trnR(tcg)_ “7. ~_ ——=, 2 atp6 nad3 trnG(tcc)©O%? Figure 1. EES Complete mitochondrial genome of S. aluco. The total length of the mitochondrial genome of S. aluco was 18,632bp. The genes located on the N strand or J strand are positioned inside or outside the circle. Contains two D-Loop regions. The GC Skew+ region contains more Guanine than Cytosine and the GC Skew- region contains more Cytosine than Guanine. Divergence time evaluation The divergence time tree, based on 37 genomes, shows that the time interval between Strigidae and Tytonidae from the common ancestor of Strigiformes was 8.05-12.75 Ma. However, in the out-group, 7yto alba, Tyto longimembris and Phodilus badius diverged from the common ancestor at about 4.23-6.69 Ma. The divergence of Strigidae began at about 6.32—10.01 Ma, the common ancestor of Ninoxinae and Striginae in Strigidae split 10 Wang Y et al into two species at 5.50-8.7 Ma and the earliest divergence of Surniinae occurred in Strigidae, Aegolius, Athene and Glaucidum occurred at about 6.0-9.5 Ma, Aegolius diverged from the common ancestor of Surniinae during 5.10-8.08 Ma, Athene and Glaucidium diverged completely into two species during 4.82—7.64 Ma. 0.97/52 Aegolius funereus MN122880 j 1/100 Glaucidium brasilianum MN356303 1/100 Glaucidium cuculoides KY 092431 1/100 1/100 Athene brama KF961185 Athene noctua MN122903 Glaucidium brodiei brodiei KP684122 1/100 Asio flammeus KP889214 Asio otus MG916810 1/100 Bubo blakistoni LC099106 1/100 1/100 Bubo flavipes LC099100 1/100 Bubo bubo MN206975 : Bubo scandiacus MG68 1084 1/100 1/100 1/100 Ciccaba nigrolineata MN356178 1/100 Strix varia varia MF431745 Strigidae j Strix occidentalis caurina MF431746 1/100 1/100 | 49 Strix aluco MN 122823 1/100 1/95 Strix aluco OP850567 *® < Strix uralensis MG681081 Strix leptogrammica KC953095 1/100 Otus bakkamoena KT340631 1/81 1/100 Otus semitorques pryeri LC541473 1/100 Otus lettia MW364567 1/100 Otus scops KT340630 Otus sunia MF346692 1/99 Ninox novaeseelandiae AY 309457 oA 1/100 Sceloglaux albifacies KX098448 1/100 Ninox strenua KX529654 Ninox scutulata KT943750 Phodilus badius KF961183 . 1/100 Tyto longimembris KP893332 Tytonidae (Out-group) Tyto alba EU410491 i Tree seale: 0.1 Figure 2. EESl BI/ML-tree, Bayesian phylogenetic tree of 37 genes (24 rRNAs, 13PCGs) from 31 species of Strigiformes. The node labels are BI/ML posterior probability and bootstrap support value, respectively and the scale indicates the probability of nucleotide change within each branch length. The GenBank of the sequences has been indicated next to the species name. Branches of different subfamilies are distinguished by different colours, with Tytoninae (with Phodilus badius, Tyto longimembris and Tyto alba) being the out-group. The Strix aluco mitochondrial genome obtained by this sequencing has been marked by *. The common ancestor of Strix and Bubo diverged completely during 3.49-5.53 Ma and, during 2.53—4.0 Ma, Strix began to gradually diverge into multiple species. In this study, the divergence time between S. aluco (OP850567) and S. aluco of Margaryan. A (IMN122823) was found to be about 1.47-2.33 Ma. The divergence time between S. aluco and S. uralensis in China was about 1.28—2.02 Ma (Fig. 3). Discussion The mitochondrial genome structure of birds is a covalent double-chain loop structure, with a total of 37 genes, including 22 tRNAs, two rRNAs, 13 PCGs and 1-2 non-coding control regions (D-loop). The nad6 and eight tRNA encoding genes (trnQ, trnA, trnN, trnC, trnY, trnS2, trnP and trnE) are located on the J chain (light chain). The remaining 14 tRNAs, two rRNAs, 12 protein-coding genes and 1—2 non-coding control regions are all located on the N chain (heavy chain) (Wolstenholme 1992, Boore 1999), which is consistent with the complete mitochondrial genome structure of all birds (Hanna et al. 2017). The complete mitochondrial genome sequence of S. a/uco obtained in this study was circular, with a total length of 18,632 bp and a GC content of 46.76%. Its composition was as follows: the ration of Adenine bases to the total base column (A%) was 29.57%; the ratio of the Guanine base Mitochondrial genome analysis, phylogeny and divergence time evaluation ... 11 to the total base column (G%) was 14.09%; the ratio of the Cytosine base to the total base column (C%) was 32.67%; the ratio of Thymidine to the total base column (T%) was 23.66%. The start codon of all 13 PCGs was ATG and the transcription stop codons were AGG, TAG and TAA. The content of A+T (53.23%) was higher than that of G+C (46.76%), which is consistent with the AT tendency of base bias in vertebrate mitochondrial genomes (Broughton et al. 2001, Ma et al. 2015). This result is consistent with the mitochondrial genome of other owls in Strigidae (Kang et al. 2018, Sun et al. 2020). TIME SCALE (Ma) [8.05, 12.75] ) Common Ancestor of Strigiformes Aegolius finereus MN122880 510.808 PROS « Athene brama KF961185 4.82, 7.64] —— « Athene noctua MN122903 aris « Glaucidium brasilianum MN356303 Glaucidium cuculoides KY 092431 « Glaucidium brodiei brodiei KP684122 BB Midevolution 192,304) Asio flammeus KP889214 —— Asio otus MG916810 Bubo blakistoni LC099106 Bubo flavipes LC099100 Bubo bubo MN206975 Bubo scandiacus MG68 1084 Ciccaba nigrolineata MN356178 Strix varia varia MF431745 Strix occidentalis caurina MF43 1746 4.63, 7.33] 2.53, 4.0] Strix aluco MN122823 {1.47,2.33 ee ears CYPR 138 2.03) Strix aluco OP850567 * [6.0, 9.50] (3.52, 5.57] 0.96, 1.52 14.03,6.39] | 65 94 4.65) 1.79, 2.84 © Approach to the present -(6.32, 10.01] 3.49, 5.53] TERE [I.61, 2.55] NOW (2.20, 3.48] Strix uralensis MG68 1081 Strix leptogrammica KC953095 (1.63,258] oe bakkamoena K’ eae? 4 Die ‘s a ts semitorques pryeri LC. pee 5:50, 8.70 ~ Lean Otus lettia MW364567 Sm ASP) (0.01, 0.02)" Otus scops KT340630 Ontus sunia MF346692 Ninox novaeseelandiae AY 309457 Sceloglaux albifacies (X098448 Ninox strenua KX529654 Ninox scutulata KT943750 Phodilus baditis KF961183 « Tyto albaEU410491 Tyto longimembris KP893332 2.0 (4.23, 6.69] Figure 3. EES Divergence time tree. Through the divergence time tree obtained by BEAST 2.6.7, based on the Bayesian method, the node horizontal bar indicates that the posterior probability of this age interval is 95% and the divergence time has been marked at the node. Phylogenetic analysis of S. aluco The BI and the ML tree have a consistent topology and each node has high posterior probability. The phylogenetic tree of Strigiformes obtained by the mitochondrial genome in this study is consistent with the phylogenetic tree obtained by Li et al. (2022) through morphology. Wink et al. (2009) compared Surniini (with Surnia, Gl/aucidium and Taenioglaux), Athenini (with Athene) and Aegolini (with Aegolius) under Surniinae, in their recent study, Surniini (with Surnia and G/aucidium) are monophyletic and cluster as a sister to Aegolius and they found that Jyto alba probably originated in Australia. They also believe that many owls that do not migrate will form new species in different places (Wink and Sauer-GU 2021). Both Salter et al. (2020) and Wink and Sauer-Gu (2021) agree on the phylogenetic relationship of (Glaucidium+Athene) + Aegolius and we agree that G. brodiei does not form an evolutionary clade with other Glaucidium. According to the genome 12 Wang Y et al analysis of Strigidae birds in Madagascar, Strix is most closely related to Bubo, followed by Otus (Fuchs et al. 2008). The phylogenetic relationship of Strigidae forms a phylogenetic relationship of Surniinae (with Surniini and Aegolius) + [Ninox + (Otus + (Asio + (Strix + Bubo)))] in this study. The conclusion of Oftus + (Asio + (Strix + Bubo)) is consistent with the conclusion of Kang et al. (2018). In the phylogenetic tree constructed by Yu et al. (2021), C. nigrolineata was also nested in Strix. S. albifacies has been extinct on the island of New Zealand and when Wood et al. (2017) extracted its mitochondrial genome from museum specimens, they suggested changing its name to Ninox albifacies because it has the same morphological structure and phylogenetic position as Ninox, in our phylogenetic tree, with relatively high posterior probability and bootstrap supporting this point. S. aluco O P850567 forms a sister group with S. uralensis, which was uploaded to GenBank by Kang et al. (2018). However, S. aluco uploaded with Margaryan. A forms a monophyly of Stnx aluco MN122823 + (Strix aluco OP850567 + Strix uralensis). Compared with the mitochondrial genome of S. a/uco obtained by Margaryan. A, S. aluco in China is more closely related to S. uralensis MG681081, which came from northeast China. Extant Eurasian birds communicated through woodland corridors during the Pleistocene interglacial. Combined with divergence-time tree analysis, such communication may have existed in the common ancestor of S. aluco IMN122823 and S. aluco OP850567 (Voelker 2010). The likely reason is that, at the beginning of the Pleistocene, the common ancestor of S. aluco MN122823 and S. aluco OP850567 had already been geographically isolated. The isolation of the Pleistocene refugium led to the divergence of the whole genome of the common ancestor of the forest owl both in China and internationally. Foreign studies have shown that the Quaternary Period is characterised by a series of glacial-interglacial cycles (Woodruff 2010), with the ancestors of modern species seeking refuge in suitable environments. The existing species on the Qinghai-Tibet Plateau may be the result of rapid population expansion in relatively warm refugia during the Pleistocene glaciation and interglacial period, which formed the current distribution pattern and genetic diversity (Gao et al. 2015). Leigong Mountain in Guizhou Province just played the role of a refugium for S. aluco during the Pleistocene glaciation period. Mitochondrial phylogeographic studies (Brito 2005) showed that the origin of S. aluco in Western Europe supports the "glacial refuge hypothesis"; further, the species survived in three allopatric refugees in Italy, the Iberian Peninsula and the Balkans, becoming the main source of S. a/uco in Europe during the late glacial period. DNA barcoding technology also showed that the geographical barrier of the Strait of Gibraltar played an extremely important role in the phylogenetic history of S. aluco (Dona et al. 2016). Divergence time evaluation of S. aluco The Pleistocene began 2.58 million years ago (2.58 Ma). The Pleistocene (especially climate change) had a profound effect on the phylogeographic structure of existing populations (Lamb et al. 2019). On the Qinghai-Tibet Plateau, the impact of mountain uplift on the formation of modern species (< 2.0 Ma) is limited and researchers have suggested that climate fluctuations played a key role in the formation of species during the Middle Pleistocene (Renner 2016, Wang et al. 2018). During the Quaternary Period and Pleistocene (1.60—2.70 Ma), there were severe climate shocks (Lisiecki and Raymo 2005), Mitochondrial genome analysis, phylogeny and divergence time evaluation ... 13 which positively promoted the formation of species (Rull 2008, Rull 2011, Rull 2015). Climatic fluctuations during this period, especially during the ice age, affected the distribution of forests in the Northern Hemisphere and the evolution of forest living species (Song et al. 2021). This series of climate fluctuations in the Pleistocene promoted species variation, which has led to species differentiation (Leonard et al. 2015). Glaciation has played an important role in influencing the population size, species and community genetic structure of today's species (Hewitt 2000, Hewitt 2004, Svendsen et al. 2004). The glacial- interglacial gyrations of the same period also affected the distribution of species (Zhao et al. 2013, Hung et al. 2014, Kozma et al. 2018, Mays et al. 2018). Glacial-interglacial cycles led to periodic shifts in glacial refuges for Pleistocene birds (Nadachowska-Bizyska et al. 2015) and the isolation of glacier refugia led to the divergence of the whole genome of species, thus forming different species (Provost et al. 2022). This is likely also the reason why the common ancestor of S. aluco MN122823 and S. aluco OP850567 (this study) diverged at 1.47—2.33 Ma. Genetic divergence of the same lineage because of the isolation of refugees leads to divergence of lineages. Various kinds of species generally begin to migrate to the best habitat during warm climate periods (Claramunt and Cracraft 2015); in particular, species adapted to low altitudes in the early stage of climate change will move to high altitudes at this time, resulting in the reproductive isolation of species in the two separated places (Wiens 2004). During the Pleistocene-Holocene (1.10—0.60 Ma), the Qinghai-Tibet Plateau experienced three stages of rapid uplift, with the formation of mountains, the climate changing from moist and warm to dry and cold and the retreat of forests to the edge of the plateau (Wang et a/. 2008). Thus, the forest landscape became what it is today. The Quaternary Period climate shock led to the initial formation of the existing forest and mountain distribution pattern in the Northern Hemisphere. Birds began to distribute widely after leaving the glacier refuge at the end of the glaciation and initially formed the existing distribution pattern (Pujolar et al. 2022). The S. uralensis may have moved north at this time and thus diverged from S. aluco. In addition, the rapid uplift of the Qinling Mountains from the end of the Early Pleistocene to the Middle Pleistocene may have formed the Qinling Mountains as a barrier to north-south bird communication (Li et al. 2019). The rapid uplift of the Qinling Mountains prevented communication between S. aluco and the common ancestor of S. uralensis, which was originally distributed on the north and south sides. Conclusions By sequencing the complete mitochondrial genome of S. a/uco and mapping its phylogenetic tree and divergence time tree, the phylogenetic relationship of Strigiformes (Tytoninae + Phodilinae) + (Striginae + Ninoxinae + Surniinae) has been summarised. Tytonidae, including Tytoninae (with Tyto) and Phodilinae (with Phodilus), are defined as the out-group; Strigidae comprises Striginae (with Asio, Bubo, Strix, Ciccaba and Otus) + Ninoxinae + Surniinae (with Athene, Aegolius and Glaucidium). The divergence time tree shows that the divergence time between S. a/luco of China and S. aluco of other countries was about 1.47-2.33 Ma, suggesting that the common ancestor of S. aluco was separated by geographical isolation at the beginning of the Pleistocene. The divergence between S. 14 Wang Y etal aluco and S. uralensis in China was about 1.28—2.02 Ma. During this time, the rapid uplift of the Qinling Mountains led to the divergence of the ancestors of Strix on the north and south sides of the Chinese mainland. At the same time, because of climatic oscillations during the Pleistocene, the existing S. aluco population on the Qinghai-Tibet Plateau may have rapidly expanded in relatively warm shelters, such as Leigong Mountain to form the current distribution pattern. Data ability The complete mitochondrial genome of Strix aluco has been uploaded to NCBI, GenBank accession number: OP850567. Acknowledgements The authors would like to thank the Rescue Center of Leigong Mountain National Nature Reserve, Qiandongnan Prefecture, Guizhou Province for providing the tissue slice of Strix aluco. Funding program Guizhou Provincial Science and Technology Foundation, Grant/Award Number: Qiankehe LH [2020] 1Y080; Project supported by the Joint Fund of the National Natural Science Foundation of China and the Karst Science Research Center of Guizhou Province, Grant/Award Number: U1812401; Science and Technology Foundation of Guizhou Forestry Bureau (Qianlinkehe [2020] 09), Guizhou Universtiy Dr. Scientific Research Fund (Guidarenjihe (2018) 07). Author contributions Yeying Wang and Haofeng Zhan contributed equally to this study. Conceptualisation: Yeying Wang, Haofeng Zhang, Yu Zhan, Zhengmin Long; Methodology: Yeying Wang, Haofeng Zhang; Software: Haofeng Zhang; Formal analysis: Haofeng Zhang; Investigation: Yeying Wang, Xiaofei Yang, Yu Zhan, Zhengmin Long; Resources: Yeying Wang, Yu Zhan, Zhengmin Long, Rescue Center of Leigong Mountain National Nature Reserve; Data curation: Haofeng Zhang; Writing - original draft: Yeying Wang, Haofeng Zhang; Writing - review & editing: Yu Zhan, Zhengmin Long, Xiaofei Yang; Visualisation: Haofeng Zhang, Yeying Wang; Supervision: Xiaofei Yang; Project administration: Xiaofei Yang; Funding acquisition: Xiaofei Yang, Yeying Wang. Mitochondrial genome analysis, phylogeny and divergence time evaluation ... 15 Conflicts of interest The authors declare no competing or financial interests. References ° Bankevich A, Nurk S, Antipov D, Gurevich A, Dvorkin M, Kulikov A, Lesin V, Nikolenko S, Pham S, Prjibelski A, Pyshkin A, Sirotkin A, Vyahhi N, Tesler G, Alekseyev M, Pevzner P (2012) SPAdes: A new genome assembly algorithm and its applications to single-cell sequencing. Journal of Computational Biology 19 (5): 455-477. https:// doi.org/10.1089/cmb.2012.0021 ° Bernt M, Donath A, Juhling F, Externbrink F, Florentz C, Fritzsch G, Putz J, Middendorf M, Stadler P (2013) MITOS: Improved de novo metazoan mitochondrial genome annotation. Molecular Phylogenetics and Evolution 69 (2): 313-319. https://doi.org/ 10.1016/j.ympev.2012.08.023 ° Boore JL (1999) Animal mitochondrial genomes. Nucleic Acids Research 27 (8): 1767-1780. https://doi.org/10.1093/nar/27.8.1767 ° Brito P (2005) The influence of Pleistocene glacial refugia on tawny owl genetic diversity and phylogeography in western Europe. Molecular Ecology 14 (10): 3077-3094. https://doi.org/10.1111/j.1365-294x.2005.02663.x ° Broughton R, Milam J, Roe B (2001) The complete sequence of the zebrafish (Danio rerio) mitochondrial genome and evolutionary patterns in vertebrate mitochondrial DNA. Genome Research 11 (11): 1958-1967. https://doi.org/10.1101/gr.156801 ° Chen S, Zhou Y, Chen Y, Gu J (2018) fastp: an ultra-fast all-in-one FASTQ preprocessor. bioRxiv 34: 884-890. https://doi.org/10.1101/274100 ° Claramunt S, Cracraft J (2015) A new time tree reveals Earth history’s imprint on the evolution of modern birds. Science Advances 1 (11): e€1501005. https://doi.org/10.1126/ sciadv.1501005 ° Coil D, Jospin G, Darling A (2015) A5-miseq: an updated pipeline to assemble microbial genomes from Illumina MiSeq data. Bioinformatics 31 (4): 587-589. https://doi.org/ 10.1093/bioinformatics/btu661 ° Comay O, Ezov E, Yom-Tov Y, Dayan T (2022) In its southern edge of distribution, the tawny owl (Strix aluco) is more sensitive to extreme temperatures than to rural development. Animals 12 (5): 641. https://doi.org/10.3390/ani12050641 ° Dona J, Ruiz-Ruano F, Jovani R (2016) DNA barcoding of Iberian Peninsula and North Africa tawny owls Strix aluco suggests the Strait of Gibraltar as an important barrier for phylogeography. Mitochondrial DNA Part A 27 (6): 4475-4478. https://doi.org/ 10.3109/19401736.2015.1089573 ° Drovetski S (2003) Plio-Pleistocene climatic oscilations, Holarctic biogeography and speciation in an avian subfamily. Journal of Biogeography 30 (8): 1173-1181. https:// doi.org/10.1046/j.1365-2699.2003.00920.x ° Feng §S, Stiller J, Deng Y, Armstrong J, Fang Q, Reeve AH, Xie D, Chen G, Guo C, Faircloth B, Petersen B, Wang Z, Zhou Q, Diekhans M, Chen W, Andreu-Sanchez S, Margaryan A, Howard JT, Parent C, Pacheco G, Sinding M, Puetz L, Cavill E, Ribeiro A, Eckhart L, Fjeldsa J, Hosner P, Brumfield R, Christidis L, Bertelsen M, Sicheritz-Ponten T, Tietze DT, Robertson B, Song G, Borgia G, Claramunt S, Lovette |, Cowen S, 16 Wang Y et al Njoroge P, Dumbacher JP, Ryder O, Fuchs J, Bunce M, Burt D, Cracraft J, Meng G, Hackett S, Ryan P, J@nsson KA, Jamieson I, da Fonseca R, Braun E, Houde P, Mirarab S, SuhA, Hansson B, Ponnikas S, Sigeman H, Stervander M, Frandsen P, van der Zwan H, van der Sluis R, Visser C, Balakrishnan C, Clark A, Fitzpatrick J, Bowman R, Chen N, Cloutier A, Sackton T, Edwards S, Foote D, Shakya S, Sheldon F, Vignal A, Soares AR, Shapiro B, Gonzalez-Solis J, Ferrer-Obiol J, Rozas J, Riutort M, Tigano A, Friesen V, Dalén L, Urrutia A, Székely T, Liu Y, Campana M, Corvelo A, Fleischer R, Rutherford K, Gemmell N, Dussex N, Mouritsen H, Thiele N, Delmore K, Liedvogel M, Franke A, Hoeppner M, Krone O, Fudickar A, Mila B, Ketterson E, Fidler AE, Friis G, Parody-Merino A, Battley P, Cox M, Lima NCB, Prosdocimi F, Parchman TL, Schlinger B, Loiselle B, Blake J, Lim HC, Day L, Fuxjager M, Baldwin M, Braun M, Wirthlin M, Dikow R, Ryder TB, Camenisch G, Keller L, DaCosta J, Hauber M, Louder MM, Witt C, McGuire J, Mudge J, Megna L, Carling M, Wang B, Taylor S, Del-Rio G, Aleixo A, Vasconcelos ATR, Mello C, Weir J, Haussler D, Li Q, Yang H, Wang J, Lei F, Rahbek C, Gilbert MTP, Graves G, Jarvis E, Paten B, Zhang G (2020) Author Correction: Dense sampling of bird diversity increases power of comparative genomics. Nature 592 (7856): 252-257. https://doi.org/10.1038/s41586-021-03473-8 Field D, Bercovici A, Berv J, Dunn R, Fastovsky D, Lyson T, Vajda V, Gauthier J (2018) Early evolution of modern birds structured by global forest collapse at the end- Cretaceous Mass Extinction. Current Biology 28 (11): 1825-1831. https://doi.org/ 10.1016/j.cub.2018.04.062 Ford NL (1967) A systematic study of the owls based on comparative osteology. Ph.D. dissertation, University of Michigan, Ann Arbor, MI, USA.. Fuchs J, Pons J, Goodman SM, Bretagnolle V, Melo M, Bowie RC, Currie D, Safford R, Virani MZ, Thomsett S, Hija A, Cruaud C, Pasquet E (2008) Tracing the colonization history of the Indian Ocean scops-owls (Strigiformes: Otus) with further insight into the spatio-temporal origin of the Malagasy avifauna. BMC Evolutionary Biology 8 (1): 1-1. https://doi.org/10.1186/1471-2148-8-197 Gao Y, Zhang Y, Gao X, Zhu Z (2015) Pleistocene glaciations, demographic expansion and subsequent isolation promoted morphological heterogeneity: A phylogeographic study of the alpine Rosa sericea complex (Rosaceae). Scientific Reports 5 (1): 1-15. https://doi.org/10.1038/srep11698 Grytsyshina EE, Kuznetsov AN, Panyutina AA (2016) Kinematic constituents of the extreme head turn of Strix aluco estimated by means of CT-scanning. Doklady Biological Sciences 466 (1): 24-27. https://doi.org/10.1134/s0012496616010087 Hanna Z, Henderson J, Sellas A, Fuchs J, Bowie RK, Dumbacher J (2017) Complete mitochondrial genome sequences of the northern spotted owl (Strix occidentalis caurina ) and the barred owl (Strix varia; Aves: Strigiformes: Strigidae) confirm the presence of a duplicated control region. PeerJ 5 https://doi.org/10.7717/peerj.3901 Haring E, Kruckenhauser L, Gamauf A, Riesing M, Pinsker W (2001) The complete sequence of the mitochondrial genome of Buteo buteo (Aves, Accipitridae) indicates an early split in the phylogeny of raptors. Molecular Biology and Evolution 18 (10): 1892-1904. https://doi.org/10.1093/oxfordjournals.molbev.a003730 Harrison GLA, McLenachan PA, Phillips MJ, Slack K, Cooper A, Penny D (2004) Four new avian mitochondrial genomes help get to basic evolutionary questions in the Late Cretaceous. Molecular Biology and Evolution 21 (6): 974-983. https://doi.org/10.1093/ molbev/msh065 Mitochondrial genome analysis, phylogeny and divergence time evaluation ... 17 Heidrich P, Wink M (1994) Tawny owl (Strix aluco) and Hume's tawny owl (Strix butler!) are distinct species: Evidence from nucleotide sequences of the cytochrome b gene. Zeitschrift fur Naturforschung C 49: 230-234. httos://doi.org/10.1515/znc-1994-3-411 Hewitt G (2000) The genetic legacy of the Quaternary ice ages. Nature 405 (6789): 907-913. https://doi.org/10.1038/35016000 Hewitt GM (2004) Genetic consequences of climatic oscillations in the Quaternary. Philosophical Transactions of the Royal Society of London. Series B: Biological Sciences 359 (1442): 183-195. https://doi.org/10.1098/rstb.2003.1388 Hoang DT, Chernomor O, von Haeseler A, Minh BQ, Vinh LS (2018) UFBoot2: Improving the ultrafast bootstrap approximation. Molecular Biology and Evolution 35 (2): 518-522. https://doi.org/10.1093/molbev/msx281 Holm SR, Svenning J (2014) 180,000 years of climate change in Europe: Avifaunal responses and vegetation implications. PLoS One 9 (4): e94021. httos://doi.org/ 10.1371/journal.pone.0094021 Ho SW, Phillips M (2009) Accounting for calibration uncertainty in phylogenetic estimation of evolutionary divergence times. Systematic Biology 58 (3): 367-380. https:// doi.org/10.1093/sysbio/syp035 Ho SW, Duchéne S (2014) Molecular-clock methods for estimating evolutionary rates and timescales. Molecular Ecology 23 (24): 5947-5965. https://doi.org/10.1111/mec. 12953 Hung C, Shaner P, Zink R, Liu W, Chu T, Huang W, Li S (2014) Drastic population fluctuations explain the rapid extinction of the passenger pigeon. Proceedings of the National Academy of Sciences 111 (29): 10636-10641. hitps://doi.org/10.1073/pnas. 1401526111 Kang H, Li B, Ma X, Xu Y (2018) Evolutionary progression of mitochondrial gene rearrangements and phylogenetic relationships in Strigidae (Strigiformes). Gene 674: 8-14. https://doi.org/10.1016/j.gene.2018.06.066 Karell P, Ahola K, Karstinen T, Zolei A, Brommer J (2009) Population dynamics in a cyclic environment: consequences of cyclic food abundance on tawny owl reproduction and survival. Journal of Animal Ecology 78 (5): 1050-1062. https://doi.org/10.1111/j. 1365-2656.2009.01563.x Koparde P, Mehta P, Reddy S, Ramakrishnan U, Mukherjee S, Robin VV (2018) The critically endangered forest owlet Heteroglaux blewitti is nested within the currently recognized Athene clade: A century-old debate addressed. PLoS One 13 (2): €0192359. https://doi.org/10.1371/journal.pone.0192359 Kozma R, Lillie M, Benito B, Svenning J, Héglund J (2018) Past and potential future population dynamics of three grouse species using ecological and whole genome coalescent modeling. Ecology and Evolution 8 (13): 6671-6681. https://doi.org/10.1002/ ece3.4163 Ksepka D, Phillips M (2015) Avian diversification patterns across the K-Pg boundary: influence of calibrations, datasets, and model misspecification. Annals of the Missouri Botanical Garden 100 (4): 300-328. hitps://doi.org/10.3417/2014032 Kurtz S, Phillippy A, Delcher AL, Smoot M, Shumway M, Antonescu C, Salzberg SL (2004) Versatile and open software for comparing large genomes. Genome Biology 5 (2): R12. https://doi.org/10.1186/gb-2004-5-2-r12 Lamb AM, Gongalves da Silva A, Joseph L, Sunnucks P, Paviova A (2019) Pleistocene- dated biogeographic barriers drove divergence within the Australo-Papuan region in a 18 Wang Y et al sex-specific manner: an example in a widespread Australian songbird. Heredity 123 (5): 608-621. https://doi.org/10.1038/s41437-019-0206-2 Lanfear R, Frandsen P, Wright A, Senfeld T, Calcott B (2017) PartitionFinder 2: New methods for selecting partitioned models of evolution for molecular and morphological phylogenetic analyses. Molecular Biology and Evolution 34: 772-77. https://doi.org/ 10.1093/molbev/msw260 Lee M, Lee S, Jeon HS, Lee S, Park J, An J (2018) Complete mitochondrial genome of the Northern Long-eared Owl (Asio otus Linnaeus, 1758) determined using next- generation sequencing. Mitochondrial DNA Part B 3 (2): 494-495. httos://doi.org/ 10.1080/23802359.2018.1451260 Leonard J, den Tex R, Hawkins MR, Mufhoz-Fuentes V, Thorington R, Maldonado J (2015) Phylogeography of vertebrates on the Sunda Shelf: a multi-species comparison. Journal of Biogeography 42 (5): 871-879. htips://doi.org/10.1111/jbi.12465 Li J, Song G, Liu N, Chang Y, Bao X (2019) Deep south-north genetic divergence in Godlewski’s bunting (Emberiza godlewskil) related to uplift of the Qinghai-Tibet Plateau and habitat preferences. BMC Evolutionary Biology 19 (1). https://doi.org/10.1186/ $12862-019-1487-z Lisiecki L, Raymo M (2005) A Pliocene-Pleistocene stack of 57 globally distributed benthic 5'8O records. Paleoceanography 20 (1): 1-17. https://doi.org/ 10.1029/2004pa001071 Liu G, Zhou L, Gu C (2014) The complete mitochondrial genome of Brown wood owl Strix leptogrammica (Strigiformes: Strigidae). Mitochondrial DNA 25 (5): 370-371. https://doi.org/10.3109/19401736.2013.803540 Liu G, Zhou L, Zhao G (2019) Complete mitochondrial genomes of five raptors and implications for the phylogenetic relationships between owls and nightjars. PeerJ Preprints No. e27478v1. https://doi.org/10.7287/peerj.preprints.27478v1 Li Z, Stidham T, Zheng X, Wang Y, Zhao T, Deng T, Zhou Z (2022) Early evolution of diurnal habits in owls (Aves, Strigiformes) documented by a new and exquisitely preserved Miocene owl fossil from China. Proceedings of the National Academy of Sciences 119 (15): e211921711. https://doi.org/10.1073/pnas.2119217119 Lutz KA, Wang W, Zdepski A, Michael TP (2011) Isolation and analysis of high quality nuclear DNA with reduced organellar DNA for plant genome sequencing and resequencing. BMC Biotechnology 11 (1): 1-9. https://doi.org/10.1186/1472-6750-11-54 Mahmood MT, McLenachan PA, Gibb GC, Penny D (2014) Phylogenetic position of avian nocturnal and diurnal raptors. Genome Biology and Evolution 6 (2): 326-332. https://doi.org/10.1093/qbe/evu016 Mays H, Hung C, Shaner P, Denvir J, Justice M, Yang S, Roth T, Oehler D, Fan J, Rekulapally S, Primerano D (2018) Genomic analysis of demographic history and ecological niche modeling in the endangered Sumatran rhinoceros Dicerorhinus sumatrensis. Current Biology 28 (1): 70-76. https://doi.org/10.1016/j.cub.2017.11.021 Ma Z, Yang X, Bercsenyi M, Wu J, Yu Y, Wei K, Fan Q, Yang R (2015) Comparative mitogenomics of the genus Odontobutis (Perciformes: Gobioidei: Odontobutidae) revealed conserved gene rearrangement and high sequence variations. International Journal of Molecular Sciences 16 (10): 25031-25049. htips://doi.org/10.3390/ iims161025031 Minh BQ, Schmidt HA, Chernomor O, Schrempf D, Woodhams MD, von Haeseler A, Lanfear R (2020) |Q-TREE 2: New models and efficient methods for phylogenetic Mitochondrial genome analysis, phylogeny and divergence time evaluation ... 19 inference in the genomic era. Molecular Biology and Evolution 37 (5): 1530-1534. https://doi.org/10.1093/molbev/msaa015 Nadachowska-Brzyska K, Li C, Smeds L, Zhang G, Ellegren H (2015) Temporal dynamics of avian populations during Pleistocene revealed by whole-genome sequences. Current Biology 25 (10): 1375-1380. https://doi.org/10.1016/j.cub. 2015.03.047 Obuch J (2011) Spatial and temporal diversity of the diet of the tawny owl (Strix aluco). Slovak Raptor Journal 5 (1): 1-120. https://doi.org/10.2478/v10262-012-0057-8 Park CE, Kim M, Ibal JCP, Pham HQ, Park HC, Shin J (2019a) The complete mitochondrial genome sequence of Ofus bakkamoena (Aves, Strigiformes, Strigidae). Mitochondrial DNA Part B 4 (1): 775-776. https://doi.org/ 10.1080/23802359.2019.1565979 Park CE, Kim M, Ibal JCP, Pham HQ, Park HC, Shin J (2019b) The complete mitochondrial genome sequence of Ofus scops (Aves, Strigiformes, Strigidae). Mitochondrial DNA Part B 4 (1): 764-765. https://doi.org/ 10.1080/23802359.2019.1565973 Pratt RC, Gibb GC, Morgan-Richards M, Phillips MJ, Hendy MD, Penny D (2009) Toward resolving deep neoaves phylogeny: data, signal enhancement, and priors. Molecular Biology and Evolution 26 (2): 313-326. https://doi.org/10.1093/molbev/ msn248 Provost K, Shue SY, Forcellati M, Smith BT (2022) The genomic landscapes of desert birds form over multiple time scales. Molecular Biology and Evolution 39 (10). https:// doi.org/10.1093/molbev/msac200 Prum R, Berv J, Dornburg A, Field D, Townsend J, Lemmon EM, Lemmon A (2015) A comprehensive phylogeny of birds (Aves) using targeted next-generation DNA sequencing. Nature 526 (7574): 569-573. https://doi.org/10.1038/nature15697 Pujolar JM, Blom MK, Reeve AH, Kennedy J, Marki PZ, Korneliussen T, Freeman B, Sam K, Linck E, Haryoko T, lova B, Koane B, Maiah G, Paul L, Irestedt M, Jansson KA (2022) The formation of avian montane diversity across barriers and along elevational gradients. Nature Communications 13 (1): 268. htips://doi.org/10.1038/ $41467-021-27858-5 Renner S (2016) Available data point to a 4-km-high Tibetan Plateau by 40 Ma, but 100 molecular-clock papers have linked supposed recent uplift to young node ages. Journal of Biogeography 43 (8): 1479-1487. httos://doi.org/10.1111/jbi.12755 Roe B (2004) Shotgun Library Construction for DNA Sequencing. Bacterial Artificial Chromosomes171-188. https://doi.org/10.1385/1-59259-752-1:171 Rull V (2008) Speciation timing and neotropical biodiversity: the Tertiary-Quaternary debate in the light of molecular phylogenetic evidence. Molecular Ecology 17 (11): 2722-2729. https://doi.org/10.1111/j.1365-294x.2008.03789.x Rull V (2011) Neotropical biodiversity: timing and potential drivers. Trends in Ecology & Evolution 26 (10): 508-513. https://doi.org/10.1016/j.tree.2011.05.011 Rull V (2015) Pleistocene speciation is not refuge speciation. Journal of Biogeography 42 (3): 602-604. https://doi.org/10.1111/jbi.12440 Salter JF, Oliveros CH, Hosner PA, Manthey JD, Robbins MB, Moyle RG, Brumfield RT, Faircloth BC (2020) Extensive paraphyly in the typical owl family (Strigidae). The Auk 137 (1). https://doi.org/10.1093/auk/ukz070 20 Wang Y et al Sarker S, Das S, Forwood J, Helbig K, Raidal S (2016) The complete mitochondrial genome sequence of an Endangered powerful owl (Ninox strenua). Mitochondrial DNA Part B 1 (1): 722-723. https://doi.org/10.1080/23802359.2016.1229588 Solonen T (2022) Body condition in the tawny owl Strix aluco near the northern limit of its range: effects of individual characteristics and environmental conditions. Animals 12 (20): 2843. https://doi.org/10.3390/ani12202843 Song K, Gao B, Halvarsson P, Fang Y, Klaus S, Jiang Y, Swenson J, Sun Y, Héglund J (2021) Demographic history and divergence of sibling grouse species inferred from whole genome sequencing reveal past effects of climate change. BMC Ecology and Evolution 21 (1): 1-10. https://doi.org/10.1186/s12862-021-01921-7 Sterli J, Pol D, Laurin M (2013) Incorporating phylogenetic uncertainty on phylogeny- based palaeontological dating and the timing of turtle diversification. Cladistics 29 (3): 233-246. https://doi.org/10.1111/j.1096-0031.2012.00425.x Stothard P, Wishart D (2005) Circular genome visualization and exploration using CGView. Bioinformatics 21 (4): 537-539. https://doi.org/10.1093/bioinformatics/bti054 Sun C, Liu H, Min X, Lu C (2020) Mitogenome of the little owl Athene noctua and phylogenetic analysis of Strigidae. International Journal of Biological Macromolecules 151: 924-931. https://doi.org/10.1016/j.ijbiomac.2020.02.238 Sunde P, Bglstad M, Desfor K (2003) Diurnal exposure as a risk sensitive behaviour in tawny owls Sinx aluco? Journal of Avian Biology 34 (4): 409-418. https://doi.org/ 10.1111/j.0908-8857.2003.03105.x Sun X, Zhou W, Sun Z, Qian L, Zhang Y, Pan T, Zhang B (2016) The complete mitochondrial genome of Glaucidium brodiei (Strigiformes: Strigidae). Mitochondrial DNA Part A 27 (4): 2508-2509. https://doi.org/10.3109/19401736.2015.1036252 Svendsen JI, Alexanderson H, Astakhov VI, Demidov |, Dowdeswell JA, Funder S, Gataullin V, Henriksen M, Hjort C, Houmark-Nielsen M, Hubberten HW, Ingélfsson O, Jakobsson M, Kjzer KH, Larsen E, Lokrantz H, Lunkka JP, Lysa A, Mangerud J, Matiouchkov A, Murray A, Moller P, Niessen F, Nikolskaya O, Polyak L, Saarnisto M, Siegert C, Siegert MJ, Spielhagen RF, Stein R (2004) Late Quaternary ice sheet history of northern Eurasia. Quaternary Science Reviews 23: 1229-1271. https://doi.org/ 10.1016/j.quascirev.2003.12.008 Tuinen Mv, Sibley C, Hedges SB (2000) The early history of modern birds inferred from DNA sequences of nuclear and mitochondrial ribosomal genes. Molecular Biology and Evolution 17 (3): 451-457. https://doi.org/10.1093/oxfordjournals.molbev.a026324 Uva V, Packert M, Cibois A, Fumagalli L, Roulin A (2018) Comprehensive molecular phylogeny of barn owls and relatives (Family: Tytonidae), and their six major Pleistocene radiations. Molecular Phylogenetics and Evolution 125: 127-137. https:// doi.org/10.1016/j.ympev.2018.03.013 Voelker G (2010) Repeated vicariance of Eurasian songbird lineages since the Late Miocene. Journal of Biogeography 37 (7): 1251-1261. https://doi.org/10.1111/j. 1365-2699.2010.02313.x Walker B, Abeel T, Shea T, Priest M, Abouelliel A, Sakthikumar S, Cuomo C, Zeng Q, Wortman J, Young S, Earl A (2014) Pilon: An integrated tool for comprehensive microbial variant detection and genome assembly improvement. PLoS One 9 (11): e112963. https://doi.org/10.1371/journal.pone.0112963 Wang C, Zhao X, Liu Z, Lippert P, Graham S, Coe R, Yi H, Zhu L, Liu S, Li Y (2008) Constraints on the early uplift history of the Tibetan Plateau. Proceedings of the Mitochondrial genome analysis, phylogeny and divergence time evaluation ... 21 National Academy of Sciences 105 (13): 4987-4992. https://doi.org/10.1073/pnas. 0703595105 Wang P, Yao H, Gilbert K, Lu Q, Hao Y, Zhang Z, Wang N (2018) Glaciation-based isolation contributed to speciation in a Palearctic alpine biodiversity hotspot: Evidence from endemic species. Molecular Phylogenetics and Evolution 129: 315-324. hitps:// doi.org/10.1016/j.ympev.2018.09.006 Wiens J (2004) Speciation and ecology revisited: phylogenetic niche conservatism and the origin of species. Evolution 58 (1): 193-197. https://doi.org/10.1111/j. 0014-3820.2004.tb01586.x Wink M, Heidrich P (2000) Molecular systematics of owls Strigiformes based on DNA- sequences of the mitochondrial cytochrome b gene. Raptors at Risk819-828. Wink M, El-Sayed A, Sauer-Girth H, Gonzalez J (2009) Molecular phylogeny of owls (Strigiformes) inferred from DNA sequences of the mitochondrial cytochrome b and the nuclear RAG-7 gene. Ardea 97 (4): 581-591. https://doi.org/10.5253/078.097.0425 Wink M, Sauer-Gi H (2021) Molecular Taxonomy and Systematics of Owls (Strigiformes)-An Update. AIRO 29: 475-488. Wolstenholme D (1992) Animal mitochondrial DNA: structure and evolution. International Review of Cytology 141: 173-216. https://doi.org/10.1016/ s0074-7696(08)62066-5 Wood J, Mitchell K, Scofield RP, De Pietri V, Rawlence N, Cooper A (2017) Phylogenetic relationships and terrestrial adaptations of the extinct laughing owl, Sceloglaux albifacies (Aves: Strigidae). Zoological Journal of the Linnean Society 179: 907-918. https://doi.org/10.1111/Z0j.12483 Woodruff D (2010) Biogeography and conservation in Southeast Asia: how 2.7 million years of repeated environmental fluctuations affect today’s patterns and the future of the remaining refugial-phase biodiversity. Biodiversity and Conservation 19 (4): 919-941. https://doi.org/10.1007/s10531-010-9783-3 Xu P, Li Y, Miao L, Xie G, Huang Y (2016) Complete mitochondrial genome of the Tyto longimembris (Strigiformes: Tytonidae). Mitochondrial DNA Part A 27 (4): 2481-2482. https://doi.org/10.3109/19401736.2015.1033708 Yan C, Mou B, Meng Y, Tu F, Fan Z, Price M, Yue B, Zhang X (2017) A novel mitochondrial genome of Arborophila and new insight into Arborophila evolutionary history. PLoS One 12 (7): e0181649. https://doi.org/10.1371/journal.pone.0181649 Yu J, Liu J, LiC, Wu W, Feng F, Wang Q, Ying X, Qi D, Qi G (2021) Characterization of the complete mitochondrial genome of Otus /ettia: exploring the mitochondrial evolution and phylogeny of owls (Strigiformes). Mitochondrial DNA Part B 6 (12): 3443-3451. https://doi.org/10.1080/23802359.2021.1995517 Zhang D, Gao F, Jakovli¢ |, Zou H, Zhang J, Li W, Wang G (2020) PhyloSuite: An integrated and scalable desktop platform for streamlined molecular sequence data management and evolutionary phylogenetics studies. Molecular Ecology Resources 20 (1): 348-355. https://doi.org/10.1111/1755-0998.13096 Zhang Y, Song T, Pan T, Sun X, Sun Z, Qian L, Zhang B (2016) Complete sequence and gene organization of the mitochondrial genome of Asio flammeus (Strigiformes, strigidae). Mitochondrial DNA Part A 27 (4): 2665-2667. https://doi.org/ 10.3109/19401736.2015.1043538 Zhao S, Zheng P, Dong S, Zhan X, Wu Q, Guo X, Hu Y, He W, Zhang S, Fan W, Zhu L, Li D, Zhang X, Chen Q, Zhang H, Zhang Z, Jin X, Zhang J, Yang H, Wang J, Wang J, 22 Wang Y et al Wei F (2013) Whole-genome sequencing of giant pandas provides insights into demographic history and local adaptation. Nature Genetics 45 (1): 67-71. httos://doi.org/ 10.1038/ng.2494 ° Zhou C, Chen Y, Hao Y, Meng Y, Yue B, Zeng T (2019) Characterization of the complete mitochondrial genome and phylogenetic analysis of Otus sunia (Strigiformes: Strigidae). Mitochondrial DNA Part B 4 (1): 804-805. https://doi.org/ 10.1080/23802359.2019.1574643 Supplementary material Suppl. material 1: Analysis of mitochondrial genome feature EE Authors: Yeying Wang, Haofeng Zhan Data type: Table Brief description: The genome annotation results showed that the total number of genes was 39, including 13 protein-coding genes, 22 tRNA genes, two rRNA genes, two O, genes and 0 O, genes. Amongst them, eight tRNA genes (trn-Q, trn-A, trn-N, trn-C, trn-Y, trn-P, trn-E and trn-S2), one PCGs gene: nad6, are on the main chain (J chain); and the remaining 14 tRNA genes are trn- F, trn-V, trn-L2, trn-l, trn-M, trn-W, trn-D, trn-K, trn-G, trn-R, trn-H, trn-S1, trn-L1 and trn-T; Two rRNA genes: rrn-S, rrn-L;with 12 PCGs genes encoding: nad1, nad2, nad3, nad4, nad4L, nad5, atp6, atp8, cox1, cox2, cox3 and cytb on the secondary (N) chain. Download file (8.23 kb)