Molecular Phylogenetics and Evolution 104 (2016) 14–20

Contents lists available at ScienceDirect

Molecular Phylogenetics and Evolution

journal homepage: www.elsevier.com/locate/ympev

Taxonomic status and phylogenetic relationship of tits based on mitogenomes and nuclear segments

Xuejuan Li a, Liliang Lin a, Aiming Cui a, Jie Bai a, Xiaoyang Wang a,b, Chao Xin a, Zhen Zhang a, Chao Yang c, ⇑ ⇑ Ruirui Gao a, Yuan Huang a, , Fumin Lei b, a Co-Innovation Center for Qinba Regions’ Sustainable Development, College of Life Sciences, Shaanxi Normal University, Xi’an 710062, China b Key Laboratory of the Zoological Systematics and Evolution, Institute of Zoology, The Chinese Academy of Sciences, Beijing 100101, China c Shaanxi Institute of Zoology, Xi’an 710032, China article info abstract

Article history: The phylogeny of tits has been studied using various molecular markers, but their phylogenetic relation- Received 6 May 2015 ships remain controversial. To further investigate their taxonomic status and phylogenetic relationships, Revised 20 June 2016 the entire mitochondrial genomes (mitogenomes) and five nuclear segments were sequenced from 10 Accepted 18 July 2016 species of tits and two outgroups (Sylviparus modestus and Remiz consobrinus), followed by the compar- Available online 19 July 2016 ison of mitogenomic characteristics and reconstruction of phylogenetic relationship based on the different datasets. The results revealed the following: the mitogenomes of 10 ingroup tits, each Keywords: 16,758–16,799 bp in length, displayed typical mitogenome organization and the gene order found in Tits most previously determined Passeriformes mitogenomes; close relationships existed between Mitogenome Nuclear segment major and P. monticolus, between P. montanus and P. palustris, and between P. ater and P. venustulus; Comparative genomics and Pseudopodoces humilis was a sister group to P. spilonotus, P. cyanus, or the clade containing P. major Phylogeny and P. monticolus. Ó 2016 Elsevier Inc. All rights reserved.

1. Introduction identified as a and has been moved to the family Paridae (James et al., 2003). In previous studies, Ps. humilis was identified The genus Parus (Passeriformes: Paridae) is commonly divided as a sister to various species or clades, such as P. major (Chen, into several subgenera based on overall morphological similarity. 2010), P. monticolus (Johansson et al., 2013), P. major and P. monti- Eck (1988) divided this genus into 12 subgroups, including a com- colus (James et al., 2003; Dai et al., 2010; Johansson et al., 2013), plex of 65 species, while Harrap and Quinn (1996) recognized 10 P. spilonotus (Chen, 2010; Dai et al., 2010), which generated great subgenera containing a total of 55 species, with the difference interest in the following question: What is the species most closely caused by the problems in defining subspecies and species. To phylogenetically related to Ps. humilis? In addition, the phyloge- facilitate evolutionary analyses among Parids, Gill et al. (2005) pro- netic relationships among these tits were not conclusively estab- posed the recognition of subgenera of Poecile, Baeolophus, Lopho- lished (Dai et al., 2010). The inconformity may be due to the phanes, Periparus, and Parus as distinct genera, in addition to poor phylogenetic performance of several mitochondrial genes or Melanochlora, Sylviparus and Pseudopodoces, which was adopted to the presence of poorly informative nuclear segments in the by Gosler and Clement (2007). previous phylogenetic studies. Therefore, a well-supported, well- Although numerous molecular markers have been applied to corroborated and presumably accurate phylogeny based on rela- study the phylogeny of tits (Sheldon et al., 1992; Gill et al., 1993; tively larger combined datasets is essential for understanding their Kvist et al., 1996; Slikas et al., 1996; Yang et al., 2006; Dai et al., evolutionary relationship and taxonomic status. 2010), the phylogenetic relationships remain controversial. For The mitogenome has been extensively used to infer avian instance, the Tibetan (Ps. humilis), which is endemic to genetic diversification with reasonable resolution. Combining the Tibetan Plateau (Yang et al., 2006) and was traditionally placed mitochondrial and nuclear markers is a powerful way to test phy- in the Corvidae (Monroe and Sibley, 1993), is now confidently logenetic and phylogeographic hypotheses (Toews and Brelsford, 2012). Most previous molecular studies of Paridae analyzed mito- chondrial genes (Dai et al., 2010), nuclear genes (Chen, 2010), or a ⇑ Corresponding authors. combination of mitochondrial and nuclear genes (Johansson et al., E-mail addresses: [email protected] (Y. Huang), [email protected] (F. Lei). http://dx.doi.org/10.1016/j.ympev.2016.07.022 1055-7903/Ó 2016 Elsevier Inc. All rights reserved. X. Li et al. / Molecular Phylogenetics and Evolution 104 (2016) 14–20 15

2013). Here, the entire mitogenomes and five nuclear segments of were 94 °C for 1 min, 50 °C for 1 min and 72 °C for 1 min, with a genes from 13 individuals from 10 tit species, as well as two out- final extension step of 72 °C for 7 min. groups (Sylviparus modestus and Remiz consobrinus), were PCR products were purified using a DNA Gel Purification Kit sequenced. In addition, the mitochondrial features of these species (Bioteke, Beijing, China) after separation by electrophoresis on a were thoroughly compared, and the phylogenetic relationships 1.0% agarose gel. After separation and purification, all PCR products were reconstructed. were directly sequenced by means of primer walking at Sangon Biotech (Shanghai) Co., Ltd. and Shenzhen Huada Gene Technology Co., Ltd. 2. Materials 2.3. Sequence assembly, annotation and analysis 2.1. Sample collection and DNA extraction Sequence data were assembled using the Staden package 1.7.0 Species here were selected with a focus on the taxa distributed (Staden et al., 2000). Most tRNA genes were identified using in China (Li et al., submitted for publication). All the samples were tRNAscan-SE 1.21 (Lowe and Eddy, 1997). The protein-coding preserved in 100% ethanol and stored at À20 °C. The voucher spec- genes, rRNAs, remaining tRNAs, and one control region were iden- imen was deposited in the Key Laboratory of Zoological Systematics tified by comparison with those of the sequenced using and Evolution, Institute of Zoology, Chinese Academy of Sciences. the Spin program of the Staden package (Staden et al., 2000). The total genomic DNA was extracted from the liver/muscle tissue Statistical analyses, such as nucleotide composition of different using the standard phenol/chloroform method (Sambrook and regions and codon usage of protein-coding genes, were carried Russell, 2001). out using MEGA 4.1 (Tamura et al., 2007). The pairwise genetic distance (P-distance) was also calculated in MEGA 4.1 based on 2.2. PCR amplification and sequencing the mitogenomic dataset with the third codon position of protein-coding genes not encoded in the RY-coding strategy. The sequences of primers used for PCR amplification and Based on the performance of the alignment software and the sequencing mitogenomes were obtained from Sorenson (2003) type of dataset (especially introns of nuclear segments) (Wang with minor changes, and the quality of primer pairs was evaluated et al., 2012), each mitochondrial gene or nuclear segment was indi- by Oligo 6.0 (Rychlik, 2004). Finally, 29 pair primers were selected vidually aligned using Muscle (Edgar, 2004). Each protein-coding with six pairs for LA-PCR (Li et al., submitted for publication; Ke gene was partitioned by different positions (first, second and third et al., 2010; Lin, 2011). sites), with the third sites using RY-coding strategy (changing puri- Six overlapping fragments of 2–5 kb were amplified using LA nes [A and G] to R and pyrimidines [C and T] to Y), which might be TaqTM DNA Polymerase (Takara) with an initial denaturation at valuable (Meiklejohn et al., 2014), while nuclear segments were 93 °C for 2 min, followed by 40 cycles of denaturation at 92 °C divided into exon and intron regions. for 10 s, annealing at 58–53 °C for 30 s, and elongation at 68 °C The mitogenomes of 13 individuals belonging to 10 tit species for 10 min in the initial 20 cycles and denaturation at 92 °C for were included in our analysis, with S. modestus (NC_026793) from 10 s, annealing at 53 °C for 30 s, and elongation at 68 °C for Paridae and R. consobrinus (NC_021641) from Remizidae used as 10 min with 20 s per cycle added to the elongation step in the suc- outgroups. Three main datasets were used to reconstruct the phy- ceeding 20 cycles, and a final elongation at 68 °C for 7 min after the logenetic relationships, dataset 1: 13 protein-coding genes of mito- last cycle. The LA-PCR amplifications were carried out in a volume genomes with the third sites employing RY-coding strategy, of 15 lL, which contained 1.5 lLof10ÂLA PCR Buffer (Mg2+ Free), dataset 2: the entire mitogenomes also with the third sites using

1.5–1.8 lL of 25 mM MgCl2, 2.4 lL of 2.5 mM dNTPs, 2.0 lL of each RY-coding strategy, and dataset 3: five nuclear segments. primer (10 lM), 0.5–2.0 lL of DNA templates, 0.18–0.20 lLof As partitioning datasets has been shown to be useful for analy- 5U/lL LA Taq polymerase (Takara, Dalian, China) and added to ses of mitogenomic data (Leavitt et al., 2013; Powell et al., 2013;

15 lL with ddH2O. Barker, 2014; Meiklejohn et al., 2014), the best scheme for parti- The 29 sets of primers were used for sub-PCR amplification of tioning datasets was selected in Partitionfinder v1.1.1 (Lanfear the 15 entire mitogenomes using LA-PCR products as templates. et al., 2012) using a greedy search algorithm and the Bayesian These short fragments were amplified using Takara Taq poly- information criterion. The heterogeneity of sequence divergence merase under the following conditions: 4 min at 95 °C, followed of datasets was analyzed using AliGROOVE (Kück et al., 2014). by 30–35 cycles of 45 s at 95 °C, 60 s at 58–53 °C, 60 s at 72 °C, The Iss metric as calculated by DAMBE (Xia et al., 2003) was used and a final elongation for 7 min at 72 °C. The sub-PCR amplifica- to test for substitution saturation. Bayesian Inference (BI) method tions were carried out in a volume of 25 lL, which contained was employed to analyze these datasets, using MrBayes 3.1.2 2+ 2.5 lL of PCR Buffer (Mg Free), 2.5 lL of 25 mM MgCl2, 2.5 lL (Ronquist and Huelsenbeck, 2003), with the best partitioned of 2.5 mM dNTPs, 2.0 lL of each primer (10 lM), 0.5–1.0 lLof scheme and optimal model analyzed in Partitionfinder v1.1.1. LA-PCR products, 0.25 lLof5U/lL LA Taq polymerase (Takara, Two runs were conducted, with each run containing one cold chain

Dalian, China) and added to 25 lL with ddH2O. and three heated chains. The analysis ran 10 million generations, The primer sequences of five nuclear segments, v-mos Moloney sampling every 100 generations. The effective sample size (ESS) murine sarcoma viral oncogene homolog (MOS), fibrinogen beta values assessed using Tracer v1.5 (Rambaut and Drummond, chain (FGB), aldolase B fructose-bisphosphate (ALDOB), pterin-4 2004), with ESS values greater than 200. The first 25% of trees were alpha-carbinolamine dehydratase (PCBD1), calbindin (CALB1), are discarded as burn-in, and the rest were used to generate the con- listed in Li et al. (submitted for publication) (Cox et al., 2007; sensus tree. Shen et al., 2014). To amplify these five nuclear segments, 50 lL Six independent datasets (the mitogenome with the third codon PCRs were performed using 25 lL of 2X Taq PCR StarMix, 20 lL position of protein-coding genes using RY-coding strategy, MOS, of ddH2O, 2 lL of each primer (10 lmol/L) and 1 lL of DNA tem- FGB, ALDOB, PCBD1, and CALB1) were used to construct gene trees. plate. The reaction comprised an initial denaturation step at The maximum likelihood (ML) method was employed to analyze 95 °C for 4 min, followed by 35 cycles of 94 °C for 1 min, 50– these datasets. ML analyses were conducted using the program 60 °C for 1 min and 72 °C for 1 min. The annealing step of each RAxML 7.0.3 (Stamatakis, 2006) with 1000 replications, employing cycle increased by 0.5 °C in the first 20 cycles, and the last 15 cycles the best partitioned scheme and optimal model as estimated by 16 X. Li et al. / Molecular Phylogenetics and Evolution 104 (2016) 14–20

MrModeltest2.2 (Nylander, 2008) or Partitionfinder v1.1.1 (Lanfear range of 0.090–0.101, which is closer to the average value et al., 2012). Then, these trees were used as input files to obtain a (0.094), confirming that Ps. humilis belongs to the tits. The results species tree in ASTRAL (Mirarab et al., 2014). among genera showed the highest p-distance value (0.102) between Pseudopodoces and Periparus and the lowest (0.071) 3. Results between Periparus and Pardaliparus (Li et al., submitted for publication). The p-distances are an important reference for divid- 3.1. Comparative genomics analyses ing a genus or subgenus, with p-distances of 0.038–0.102 between species in tits and of 0.071–0.102 between genera in tits, which 3.1.1. New mitogenomes may provide information for differentiating other species of tits. The mitogenomes of S. modestus, R. consobrinus and 13 individ- uals from 10 tit species were sequenced. The annotated mitogen- omes have been deposited in the GenBank database with the 3.1.3. A + T content and nucleotide skew accession numbers shown in Li et al. (submitted for publication). The A + T contents of mitogenomes in the 10 ingroup tit species The entire mitogenome sequences of the 10 ingroup tit species ranged from 50.8% (P. dichrous) to 52.4% (Ps. humilis), with an aver- were from 16,758 bp (Ps. humilis) to 16,799 bp (P. dichrous) in size age value of 51.8% (Li et al., submitted for publication). In the four and encoded the 37 genes typically found in most Passeriformes major partitions (protein-coding genes, rRNA, tRNA and control species: 13 protein-coding genes, 2 rRNA genes, and 22 tRNA genes region) from 10 ingroup tit species mitogenomes, the A + T content (Fig. 1). In addition, one non-coding control region was present in was tRNA > control region > rRNA > protein-coding genes, except the mitogenomes, located between trnE and trnF. The gene order of for Ps. humilis. The tRNA genes encoded in the H-strand showed the 10 ingroup tit species was consistent with that of most Passer- slightly greater A + T content than that in the L-strand. Based on iformes species. All genes were coded on the H-strand, except for the nucleotide compositions of rRNA, rrnL included more A + T one protein-coding gene (nad6) and eight tRNA genes [trnQ, trnA, content than did rrnS. In the comparison of A + T content in the trnN, trnC, trnY, trnS(ucn), trnP and trnE]. three codon positions of protein-coding genes, the second codon position contained higher values than the first and third codon 3.1.2. Genetic distances positions. In the results among tits, the p-distance value (0.102) calculated All 10 ingroup tit species mitogenomes showed A/C skew. The from the whole mitogenome dataset was the highest between skew degree of the four major partitions as mentioned above P. ater and Ps. humilis, while the lowest (0.038) was between showed that protein-coding genes and rRNA had A/C skew. There P. palustris and P. montanus (Li et al., submitted for publication). was obvious A/C skew exhibited in the third codon of protein- The p-distances between Ps. humilis and other tits were in the coding genes. The rrnS and rrnL genes showed A/C skew. The

Fig. 1. Gene map of the P. major mitogenome. Transfer RNA genes are designated by single-letter amino acid codes. L1, L2, S1, and S2 denote trnL(uur), trnL(cun), trnS(ucn) and trnS(agy), respectively. Clockwise arrow represented genes encoded by H-strand, while counterclockwise arrow represented genes encoded by L-strand. X. Li et al. / Molecular Phylogenetics and Evolution 104 (2016) 14–20 17

H-strand tRNA genes contained A/C skew, while L-strand genes suggesting that some phylogenetic relationships were well sup- showed T/G skew. ported by these gene trees.

3.1.4. Conserved sites in mitochondrial genes 4. Discussion The percentage of conserved sites in the whole 37 genes plus one control region revealed that trnS(ucn) was the most conserved 4.1. Characteristics of mitogenomes gene, while the control region varied most (Li et al., submitted for publication). Overall, RNA sequences (including tRNA and rRNA) There are two gene arrangement patterns between the cytb and tended to have more conserved sites than protein-coding genes rrnS genes in mitogenomes of Aves. One is the traditional model and the control region. For nucleotide sequences of all protein- with the trnT/trnP/nad6/trnE/control region/trnF arrangement, coding genes, NADH genes contained fewer conserved sites, while and the other is a rearranged model with the trnT/control region/ COX genes contained many more. The conserved site percentage in trnP/nad6/trnE/non-coding region/trnF arrangement. These two amino acid sequences of all protein-coding genes was significantly patterns exist simultaneously in Passeriformes. Some Passeri- higher than that of corresponding nucleotide sequences, with the formes species, such as Podoces hendersoni (Ke et al., 2010) and highest conserved site percentage occurring in cox1 and the lowest Pyrgilauda ruficollis (Ma et al., 2014), had the first pattern, while in atp8. The conserved site percentage was similar between the others, such as Alauda arvensis (Qian et al., 2013) and Progne chaly- two rRNAs. In comparing the tRNAs in 10 ingroup tit mitogenomes, bea (Cerasale et al., 2012) possessed the second pattern. The gene there were more conserved sites in trnS(ucn), trnL(uur), trnT and order of mitogenomes was consistent in the 15 individuals of trnL(cun) and fewer in trnE, trnP and trnF. Passeriformes, was similar to most Passeriformes species, and belonged to the first pattern. The common model used to explain 3.2. Phylogenetic relationships the mechanism of vertebrate mitochondrial gene rearrangements is the duplication and random deletion model (Jeffrey and The results of base compositional heterogeneity using Ali- Weslley, 1998; Sumida et al., 2001). Based on this model, Bensch GROOVE showed consistently positive scores in pairwise compar- and Härlid (2000) proposed that the rearrangement in the Aves isons (Li et al., submitted for publication). The heterogeneity was mitogenome was formed as follows: the trnP/nad6/trnE/control similar between the first and second datasets (Li et al., submitted region duplicated first, and then some genes were partially deleted. for publication), which may indicate that the third codon position Singh et al. (2007) studied the mitogenomes of three warblers might contain a lower rate of heterogeneity. The heterogeneity was (Sylvia atricapilla, Acrocephalus scirpaceus and S. crassirostris) and slightly stronger for datasets from protein-coding genes and mito- proposed a more detailed explanation based on the lowest cost genomes than for the dataset of nuclear segments. Pairwise principle from an evolutionary perspective, in which the arrange- sequence comparisons involving S. modestus, R. consobrinus and ment in Sylviidae was likely to undergo concerted evolution with other remaining species received similarity scores between those duplication of the control region to form control region 2 (which of other sequences in the mitochondrial datasets. The species con- was in agreement with Gibb et al., 2007), and then control region taining less sequence heterogeneity could be robustly placed in the 2 developed into various non-coding regions in different species phylogenetic tree. via base deletion and substitution. The Iss results showed Iss = 0.7450 and Iss.c = 0.8218 for the The A + T contents of mitogenomes in 10 ingroup tit species dataset 1; Iss = 0.5759 and Iss.c = 0.9353 for the dataset 2; and (50.8%–52.3%) were similar to those of other Passeriformes species, Iss = 0.1143 and Iss.c = 0.8148 for the dataset 3. For all of these such as P. ruficollis (55.0%) (Ma et al., 2014). Compared with other datasets, Iss is less than Iss.c, and the P value is equal to 0, which partitions, tRNA-J and the second codon position of protein-coding indicates that none of the datasets exhibited substitution genes have less A + T content, which may imply that these regions saturation. contain higher mutation rates. The A + T content may result from The final alignment resulted in 11,361 sites for dataset 1, 17,104 transcription-coupled repair and deamination processes sites for dataset 2, and 3095 sites for dataset 3. BI analyses (Francino and Ochman, 1997). The slightly higher A + T content employed the best partitioned scheme and optimal models of evo- in the second codon position of protein-coding genes than in the lution (Li et al., submitted for publication) and resulted in a robust other two positions of protein-coding genes indicates that they tree with relatively high posterior probabilities (Fig. 2). The require AT-rich nucleotides to encode amino acids. monophyly of genus Parus and Poecile was well supported in all Comparison of 10 ingroup tit species revealed A/C skew in the phylogenetic trees. The sister groups between P. major and mitogenomes. The mitogenomes revealed weak A skew and P. monticolus, between P. montanus and P. palustris, and between approximately moderate C skew. It is possible that the overall gen- P. ater and P. venustulus were consistent in phylogenetic trees ome A bias was driven by mutational pressure on the L-strand obtained from the three datasets. The phylogenetic positions of (Cameron and Whiting, 2007) and that the GC skew might be cor- P. spilonotus, P. cyanus and P. dichrous were different in the trees related with the asymmetric replication process of the mtDNA from the three datasets, and the position of Ps. humilis was incon- (Saccone et al., 1999). The more conserved genes (such as tRNAs, sistently associated with various species or clades as a sister group, rRNAs, and the amino acid sequences of protein-coding genes) i.e., P. spilonotus in dataset 1, P. cyanus in dataset 2, and the clade of and more varied genes (such as the nucleotide sequence of nad6 P. major/P. monticolus in dataset 3, all of which showed relatively and the control region) might provide a basis for selecting appro- high posterior probabilities (1.00). priate molecular markers in further phylogenetic studies of tits. The topologies were different among the six gene trees at the genus or even species level, with relatively lower bootstrap sup- port in some nodes (Li et al., submitted for publication). For exam- 4.2. Phylogenetic relationship ple, in the species tree, P. monticolus was placed in the basal position, whereas in the gene trees, R. consobrinus was placed in The mitogenome-based phylogenetic results in our study were the basal position. This incongruence may be due to the types of generally similar to the results of previous studies (Tietze and data used in each analysis. However, the consistent and well- Borthakur, 2012; Dai et al., 2010; Chen, 2010). The genera Parus, supported nodes estimated from most gene trees, such as the clade Machlolophus and Pseudopodoces showed closer relationships in containing P. ater and P. venustulus, were also in the species trees, both mitochondrial datasets and the nuclear segment dataset, as 18 X. Li et al. / Molecular Phylogenetics and Evolution 104 (2016) 14–20

Fig. 2. Phylogenetic reconstruction using Bayesian Inference method based on three datasets. A: dataset 1 including 13 protein-coding genes with the third sites employing RY-coding strategy; B: dataset 2 of mitogenomes also with the third sites using RY-coding strategy; C: dataset 3 containing five nuclear segments.

did Poecile, Periparus, Pardaliparus and Lophophanes. The sister genetic position of Ps. humilis needs to be re-evaluated with exten- group between Periparus and Pardaliparus was consistent in both sive species sampling during a future study. the mitochondrial and nuclear trees. The phylogenetic tree obtained from nuclear segments dataset All of our analyses revealed strong support for the hypothesis was incongruent with that of mitochondrial trees in some nodes (Li et al., submitted for publication) that P. major and P. monticolus, (Fig. 2). For example, P. cyanus had a closer relationship with the P. montanus and P. palustris, and P. ater and P. venustulus are paired clade of P. montanus/P. palustris/P. dichrous/P. ater/P. venustulus in sister taxa, which indicated their closer phylogenetic relationships nuclear trees, while it formed a sister group with Ps. humilis or (Dai et al., 2010; Chen, 2010). This was consistent with their mor- the clade P. major/P. monticolus/P. spilonotus/Ps. humilis in mito- phological features (Li et al., submitted for publication; Xu and chondrial trees. These incongruent branches of the phylogenetic Feng, 2004). For instance, there are numerous similarities between relationship might be explained by the following reasons. Indepen- P. major and P. monticolus, such as a rounded tail and no crest on dent mitogenome data are often inadequate for macroevolutionary the head; P. montanus and P. palustris are very similar in plumage phylogenetic analyses because they reflect only the matrilineal coloration, such as no blue or white on the body feathers and no genealogy (Shen et al., 2014). Thus, phylogenetic reconstruction solid yellow on the lower part of the body; P. ater and P. venustulus based only on mitogenomes might be insufficient. Moreover, the show many similarities, such as lacking two white stripes on the relatively slow mutation rate of nuclear genes also caused poor wings but showing splendid blue-black on the head (Zheng, 1964). resolution (Shen et al., 2014). The incongruent geographic patterns The P-distances between Ps. humilis and other tit species (from between mitochondrial and nuclear markers are mostly due to the 0.090 to 0.101) were similar to those of other tit species, which processes of adaptive introgression of mitogenomes, demographic confirmed that Ps. humilis belongs to the tits. Based on the whole disparities, sex-biased asymmetries, hybrid-zone movement, Wol- genome but limited species, Qu et al. (2013) also suggested that bachia infection and human introductions (Toews and Brelsford, Ps. humilis is more closely related to P. spilonotus than to P. major, 2012). For example, hybridization occurs between P. major and which was roughly consistent with the phylogenetic tree based P. minor in overlap regions of the middle Amur region, and strong on dataset 1. However, the phylogenetic position of Ps. humilis introgression was supported by combined mitochondrial and remains unstable, which was consistent with the hypothesis nuclear markers (Fedorov et al., 2009). Furthermore, mitochondrial (Fig. 2). The phylogenetic results based on mitochondrial genes and nuclear incongruence between P. minor and P. major (Kvist and showed that Ps. humilis was a sister group of P. cyanus,orof Rytkoenen, 2006) was proposed to be due to sex-biased asymme- P. spilonotus which was consistent with BI tree in the study of tries (Toews and Brelsford, 2012). Dai et al. (2010). However, the phylogenetic results based on The phylogenetic relationships based on different mitogenome nuclear segments showed that Ps. humilis formed a sister group regions (Fig. 2A and B) also showed incongruence but with high with a clade including P. major and P. monticolus, which was similar posterior probabilities. The incongruence among different mito- to the study of Johansson et al. (2013) and ML tree of Dai et al. chondrial regions has been observed in galliform (Cox et al., (2010). Therefore, the mitochondrial and nuclear discordant phylo- 2007; Meiklejohn et al., 2014). However, it is still difficult to ana- X. Li et al. / Molecular Phylogenetics and Evolution 104 (2016) 14–20 19 lyze mitochondrial data, despite testing multiple strategies to Kück, P., Meid, S.A., Groß, C., Wägele, J.W., Misof, B., 2014. AliGROOVE – reduce this incongruence (Meiklejohn et al., 2014). Therefore, the visualization of heterogeneous sequence divergence within multiple sequence alignments and detection of inflated branch support. BMC Bioinf. incongruence among different regions of mitogenomes should be 15, 294. received careful attention because mitochondrial genes still have Kvist, L., Ruokonen, M., Orell, M., Lumme, J., 1996. Evolutionary patterns and the potential to play an important role in phylogenetic phylogeny of tits and chickadees (genus Parus) based on the sequence of the mitochondrial cytochrome b gene. Ornis Fenn. 73, 145–156. reconstruction. Kvist, L., Rytkoenen, S., 2006. Characterization of a secondary contact zone of the great tit Parus major and the Japanese tit P. minor (Aves: Passeriformes) in far Acknowledgements eastern Siberia with DNA markers. Zootaxa 1325, 55–73. Lanfear, R., Calcott, B., Ho, S.Y., Guindon, S., 2012. Partitionfinder: combined selection of partitioning schemes and substitution models for phylogenetic This study was supported by the State Key Program of NSFC analyses. Mol. Biol. Evol. 29, 1695–1701. (31330073), the Strategic Priority Research Program of the Chinese Leavitt, J.R., Hiatt, K.D., Whiting, M.F., Song, H., 2013. Searching for the optimal data partitioning strategy in mitochondrial phylogenomics: a phylogeny of Academy of Sciences (Grant No. XDB13020300, XDA05080703), Acridoidea (Insecta: Orthoptera: Caelifera) as a case study. Mol. Phylogenet. the Ministry of Science and Technology of the People’s Republic Evol. 67 (2), 494–508. of China (MOST Grant No. 2011FY120200-3) and the Key Labora- Lin, L.L., 2011. Sequencing of Four Tits Mitochondrial Genomes and Mitochondrial Phylogenetic Analysis of Passeriformes. Shaanxi Normal University, Xi’an. tory of the Zoological Systematics and Evolution of the Chinese Lowe, T.M., Eddy, S.R., 1997. TRNAscan-SE: a program for improved detection of Academy of Sciences (No. O529YX5105). transfer RNA genes in genomic sequence. Nucleic Acids Res. 25, 955–964. Li, X.J., Lin, L.L., Cui, A.M., Bai, J., Wang, X.Y., Xin, C., Zhang, Z., Yang, C., Gao, R.R., Huang, Y., Lei, F.M., 2016. Data on taxonomic status and phylogenetic Appendix A. Supplementary material relationship of tits. Data in Brief (submitted for publication). Ma, Y.G., Huang, Y., Lei, F.M., 2014. Sequencing and phylogenetic analysis of the Supplementary data associated with this article can be found, in Pyrgilauda ruficollis (Aves, Passeridae) complete mitochondrial genome. Zool. Res. 35 (2), 81–91. the online version, at http://dx.doi.org/10.1016/j.ympev.2016.07. Meiklejohn, K.A., Danielson, M.J., Faircloth, B.C., Glenn, T.C., Braun, E.L., Kimball, R.T., 022. 2014. Incongruence among different mitochondrial regions: a case study using complete mitogenomes. Mol. Phylogenet. Evol. 78, 314–323. Mirarab, S., Reaz, R., Bayzid, M.S., Zimmermann, T., Swenson, M.S., Warnow, T., References 2014. ASTRAL: genome-scale coalescent-based species tree estimation. Bioinformatics 30 (17), i541–i548. Barker, F.K., 2014. Mitogenomic data resolve basal relationships among passeriform Monroe, B.L., Sibley, C.G., 1993. A World Checklist of Birds. Yale University Press, and passeridan birds. Mol. Phylogenet. Evol. 79, 313–324. New Haven and London. Bensch, S., Härlid, A., 2000. Mitochondrial genomic rearrangements in songbirds. Nylander, J.A.A., 2008. MRMODELTEST, Version 2.3 . Cameron, S.L., Whiting, M.F., 2007. Mitochondrial genomic comparisons of the Powell, A.F., Barker, F.K., Lanyon, S.M., 2013. Empirical evaluation of partitioning subterranean termites from the Genus Reticulitermes (Insecta: Isoptera: schemes for phylogenetic analyses of mitogenomic data: an avian case study. Rhinotermitidae). Genome 50, 188–202. Mol. Phylogenet. Evol. 66 (1), 69–79. Cerasale, D.J., Dor, R., Winkler, D.W., Lovette, I.J., 2012. Phylogeny of the Tachycineta Qian, C.J., Wang, Y.X., Guo, Z.C., Yang, J.K., Kan, X.Z., 2013. Complete mitochondrial genus of New World swallows: insights from complete mitochondrial genomes. genome of Skylark, Alauda arvensis (Aves: Passeriformes): the first Mol. Phylogenet. Evol. 63 (1), 64–71. representative of the family Alaudidae with two extensive heteroplasmic Chen, K., 2010. Molecular Phylogenetic Research of Paridae Based on Nuclear Genes. control regions. Mitochondr. DNA 24 (3), 246–248. Qufu Normal University, Ji’ning. Qu, Y.H., Zhao, H.W., Han, N.J., Zhou, G.Y., Song, G., Gao, B., Tian, S.L., Zhang, J.B., Cox, W.A., Kimball, R.T., Braun, E.L., 2007. Phylogenetic position of the New World Zhang, R.Y., Meng, X.H., Zhang, Y., Zhang, Y., Zhu, X.J., Wang, W.J., Lambert, D., Quail (Odontophoridae): eight nuclear loci and three mitochondrial regions Ericson, P.G.P., Yeung, C., Zhu, H.M., Jiang, Z., Li, R.Q., Lei, F.M., 2013. Ground tit contradict morphology and the Sibley-Ahlquist tapestry. Auk 124 (1), 71–84. genome reveals avian adaptation to living at high altitudes in the Tibetan Dai, C.Y., Chen, K., Zhang, R.Y., Yang, X.J., Yin, Z.H., Tian, H.J., Zhang, Z.M., Hu, Y., Lei, plateau. Nat. Commun. 4, 2071. F.M., 2010. Molecular phylogenetic analysis among species of Paridae, Rambaut, A., Drummond, A., 2004. Tracer. Published by the Authors. Remizidae and Aegithalos based on mtDNA sequences of COI and cyt b. Chin. Ronquist, F., Huelsenbeck, J.P., 2003. MrBayes 3: Bayesian phylogenetic inference Birds 1 (2), 112–123. under mixed models. Bioinformatics 19, 1572–1574. Eck, S., 1988. Gesichtspunkte zur Art-Systematik der Meisen (Paridae) (Aves). Zool. Rychlik, W.P., 2004. Oligo, Primer Analysis Software (Ver. 6.70). Molecular Biology Abh. (Dres.) 43, 101–134. Insights, Cascade, CO 80809, USA. Edgar, R.C., 2004. MUSCLE: multiple sequence alignment with high accuracy and Saccone, C., De, G.C., Gissi, C., Pesole, G., Reyes, A., 1999. Evolutionary genomics in high throughput. Nucleic Acids Res. 32, 1792–1797. Metazoa: the mitochondrial DNA as a model system. Gene 238, 195–209. Fedorov, V.V., Surin, V.L., Val’chuk, O.P., Kapitonova, L.V., Kerimov, A.B., Formozov, Sambrook, J., Russell, D.W., 2001. Molecular Cloning: A Laboratory Manual, third N.A., 2009. Maintaining of the morphological specificity and genetic ed. Cold Spring Harbor Laboratory Press, New York. introgression in populations of the great tit Parus major and the Japanese tit Sheldon, F.H., Slikas, B., Kinnarney, M., Gill, F.B., Zhao, E., Silverin, B., 1992. DNA- P. minor in the middle Amur region. Genetika 45 (7), 881–892. DNA hybridisation evidence of phylogenetic relationships among major Francino, M.P., Ochman, H., 1997. Strand asymmetries in DNA evolution. Trends lineages of Parus. Auk 109, 173–185. Genet. 13, 240–245. Shen, Y.Y., Dai, K., Cao, X., Murphy, R.W., Shen, X.J., Zhang, Y.P., 2014. The updated Gibb, G.C., Kardailsky, O., Kimball, R.T., Braun, E.L., Penny, D., 2007. Mitochondrial phylogenies of the Phasianidae based on combined data of nuclear and genomes and avian phylogeny: complex characters and resolvability without mitochondrial DNA. PLoS ONE 9 (4), e95786. explosive radiations. Mol. Biol. Evol. 24 (1), 269–280. Singh, T.R., Shneor, O., Huchon, D., 2007. mitochondrial gene order: insight Gill, F.B., Mostrom, A.M., Mack, A.L., 1993. Speciation in North American chickadees: from 3 warbler mitochondrial genomes. Mol. Biol. Evol. 25 (3), 475–477. I. Patterns of mtDNA genetic divergence. Evolution 47, 195–212. Slikas, B., Sheldon, H., Gill, F.G., 1996. Phylogeny of titmice (Paridae): I. Estimate of Gill, F.B., Slkas, B., Sheldon, F.H., 2005. Phylogeny of titmice (Paridae): II. Species relationships among subgenera based on DNA-DNA hybridization. J. Avian Biol. relationships based on sequences of the mitochondrial cytochrome-b gene. Auk 27, 70–82. 122, 121–143. Sorenson, M.D., 2003. Avian mtDNA Primers [OL] . Elliott, A., Christie, D.A. (Eds.), Handbook of the Birds of the World. Lynx Staden, R., Beal, K.F., Bonfield, J.K., 2000. The Staden package, 1998. Methods Mol. Edicions, Barcelona, pp. 662–744. Biol. 132, 115–130. Harrap, S., Quinn, D., 1996. Tits, Nuthatches and Treecreepers. Helm, London. Stamatakis, A., 2006. RAxML-VI-HPC: maximum likelihood-based phylogenetic James, H.F., Ericson, P.G.P., Slikas, B., Lei, F.M., Gill, F.B., Olson, S.L., 2003. analyses with thousands of taxa and mixed models. Bioinformatics 22, 2688– Pseudopodoces humilis, a misclassified terrestrial tit (Paridae) of the Tibetan 2690. Plateau: evolutionary consequences of shifting adaptive zones. Ibis 145, 185– Sumida, M., Kanamori, Y., Kaneda, H., Kato, Y., Nishioka, M., Hasegawa, M., 202. Yonekawa, H., 2001. Complete nucleotide sequence and gene rearrangement Jeffrey, L.B., Weslley, M.B., 1998. Big trees from little genomes: mitochondrial gene of the mitochondrial genome of the Japanese pond frog Rana nigromaculata. order as a phylogenetic tool. Curr. Opin. Genet. Dev. 8, 668–674. Genes Genet. Syst. 76, 311–325. Johansson, U.S., Ekman, J., Bowie, R.C., Halvarsson, P., Ohlson, J.I., Price, T.D., Ericson, Tamura, K., Dudley, J., Nei, M., Kumar, S., 2007. MEGA4: molecular evolutionary P.G., 2013. A complete multilocus species phylogeny of the tits and chickadees genetics analysis (MEGA) software version 4.0. Mol. Biol. Evol. 24 (8), 1596– (Aves: Paridae). Mol. Phylogenet. Evol. 69, 852–860. 1599. Ke, Y., Huang, Y., Lei, F.M., 2010. Sequencing and analysis of the complete Tietze, D.T., Borthakur, U., 2012. Historical biogeography of tits (Aves: Paridae, mitochondrial genome of Podoces hendersoni. Hereditas (Beijing) 32 (9), 951–960. Remizidae). Org. Divers. Evol. 12, 433–444. 20 X. Li et al. / Molecular Phylogenetics and Evolution 104 (2016) 14–20

Toews, D.P., Brelsford, A., 2012. The biogeography of mitochondrial and nuclear Xu, Z.W., Feng, N., 2004. A Field Guide to the Wild of Shaanxi. Shaanxi Tour discordance in animals. Mol. Ecol. 21, 3907–3930. Press, Xi’an. Wang, N., Braun, E.L., Kimball, R.T., 2012. Testing hypotheses about the sister group Yang, S.J., Yin, Z.H., Ma, X.M., Lei, F.M., 2006. Phylogeography of ground tit of the Passeriformes using an independent 30-locus data set. Mol. Biol. Evol. 29 (Pseudopodoces humilis) based on mtDNA: evidence of past fragmentation on (2), 737–750. the Tibetan Plateau. Mol. Phylogenet. Evol. 41, 257–265. Xia, X., Xie, Z., Salemi, M., Chen, L., Wang, Y., 2003. An index of substitution Zheng, Z.X., 1964. Chinese Birds System Retrieval. Science Press, Beijing. saturation and its application. Mol. Phylogenet. Evol. 26, 1–7.