Skip to main content

Filling the gap - COI barcode resolution in eastern Palearctic birds

Abstract

Background

The Palearctic region supports relatively few avian species, yet recent molecular studies have revealed that cryptic lineages likely still persist unrecognized. A broad survey of cytochrome c oxidase I (COI) sequences, or DNA barcodes, can aid on this front by providing molecular diagnostics for species assignment. Barcodes have already been extensively surveyed in the Nearctic, which provides an interesting comparison to this region; faunal interchange between these regions has been very dynamic. We explored COI sequence divergence within and between species of Palearctic birds, including samples from Russia, Kazakhstan, and Mongolia. As of yet, there is no consensus on the best method to analyze barcode data. We used this opportunity to compare and contrast three different methods routinely employed in barcoding studies: clustering-based, distance-based, and character-based methods.

Results

We produced COI sequences from 1,674 specimens representing 398 Palearctic species. These were merged with published COI sequences from North American congeners, creating a final dataset of 2,523 sequences for 599 species. Ninety-six percent of the species analyzed could be accurately identified using one or a combination of the methods employed. Most species could be rapidly assigned using the cluster-based or distance-based approach alone. For a few select groups of species, the character-based method offered an additional level of resolution. Of the five groups of indistinguishable species, most were pairs, save for a larger group comprising the herring gull complex. Up to 44 species exhibited deep intraspecific divergences, many of which corresponded to previously described phylogeographic patterns and endemism hotspots.

Conclusion

COI sequence divergence within eastern Palearctic birds is largely consistent with that observed in birds from other temperate regions. Sequence variation is primarily congruent with taxonomic boundaries; deviations from this trend reveal overlooked biological patterns, and in some cases, overlooked species. More research is needed to further refine the taxonomic status of some Palearctic birds, but large genetic surveys such as this may facilitate this effort. DNA barcodes are a practical means for rapid species assignment, although efficient analytical methods will likely require a two-tiered approach to differentiate closely related pairs of species.

Background

DNA barcoding employs sequences from a short standardized gene region to identify species [1]. The mitochondrial gene cytochrome c oxidase I (COI) has been firmly established as the core barcode region for animals [2] and its performance has been evaluated in birds from several regions, including North America [3], Brazil [4, 5], Argentina [6], and Korea [7]. While most bird species are readily identifiable through morphological traits [8], their well-developed taxonomy makes them a valuable group to test the efficacy of barcoding. Additionally, avian taxonomy is not immune to change, and in recent decades DNA evidence has clarified many species boundaries. Broad surveys, such as DNA barcoding, can expedite this process by quickly spotlighting species that merit further taxonomic investigation [9–11]. This capacity is illustrated by several recently described species that were earlier revealed as divergent lineages during barcode surveys [12–14].

Although the avian diversity of the Palearctic is relatively depauperate [15] and its taxonomy was stable for decades, modern molecular techniques have spurred the recognition of overlooked species [16]. These new species were often hidden within morphologically cryptic assemblages, which impeded their discovery [e.g. [17, 18]]. In other cases, biological species hypotheses could not be tested because divergent populations had allopatric distributions [19–21]. Molecular analyses continue to illuminate the phylogeographic structure of birds in this region [20, 22–28]. A recent barcoding survey of Scandinavian birds by Johnsen et al. [29] revealed high species resolution plus a few divergent lineages, including some between European and North American populations of trans-Atlantic species. The Atlantic Ocean serves as a relatively impermeable barrier to dispersal for non-pelagic birds [15, 30], but the situation is very different in the eastern Palearctic, where intercontinental exchange across the Bering Strait is more frequent [19, 24, 31]. Johnsen et al. [29] also highlighted sequence divergences within a few species that failed to correspond to known subspecies or logical geographical patterns - a pattern not observed in a comprehensive survey of Nearctic birds [3]. To determine if this pattern is recurrent, to highlight further cases of cryptic divergences, and to explore general patterns in sequence divergence, we advance COI barcode coverage in this study to include the breeding birds of the eastern Palearctic region, including Russia, Ukraine, Kazakhstan, and Mongolia.

Despite the growth of DNA barcode libraries, no consensus has yet emerged on the best method to analyze DNA barcode data [32]. Some of the original tools proposed to delimit species using COI sequences, such as neighbour-joining profiles [33] and distance thresholds [34], have been criticized by several authors for not realistically addressing the complexity of species boundaries [35–38]. More recent tools have gained complexity, incorporating coalescent theory and more elaborate statistical methods, though at the cost of computational time and power [38–40]. The situation is further complicated by the dual purposes proposed for barcoding: species identification and species discovery [41]. The majority of new generation tools require pre-defined species designations and consequently cannot be used to identify divergent genetic lineages within known groups. Although the use of DNA barcodes to "discover" species is contentious, it is generally accepted that barcode data can be used to flag potentially distinct taxa for further hypothesis testing [42]. Because the taxonomy of Holarctic birds is relatively mature [35], we take this opportunity to compare and contrast some of the more commonly used analytical methods.

Methods

Sampling

We examined 1,674 individuals representing 398 Palearctic species, with 83% of these taxa represented by multiple individuals. Species coverage was not uniformly distributed across orders and families due to specimen availability; nearly two-thirds of resident passerines were represented, versus less than 38% of non-passerine birds. We used frozen tissue (typically pectoral muscle) from museum specimens; all but six tissues were linked to vouchered specimens. All tissue specimens originated from either the ornithology collection at the Burke Museum of Natural History and Culture (87.5%) or from the Zoological Museum of Moscow University (12.5%), and were collected in the field during the past 20 years. To capture geographical variation, individuals collected from widely dispersed sites were preferentially sampled for each species whenever possible (see Figure 1 for distribution of collecting sites). Additional sequences from North American congeners were also contributed (see below). As a taxonomic reference, we followed Clements [43], including corrections and updates up to 8 October 2007 with the exception of treating Corvus cornix as conspecific with C. corone [sensu[44]].

Figure 1
figure 1

Distribution of collecting sites. Map of the eastern Palearctic region detailing the collecting sites for all specimens used in this study. Red circles indicate sampling sites. Sampling intensity is indicated by the brightness of each circle.

Laboratory methods

DNA extraction, PCR, and sequencing reactions follow the procedures described in Kerr et al. [6]. Only sequences greater than 500 bp and containing fewer than 10 ambiguous base calls were included in analyses. The sequence from one Anas crecca specimen was omitted from analysis due to suspicion that it was actually an A. crecca × A. carolinensis hybrid based on morphology and molecular results. Collection data, sequences, and trace files are available from the project 'Birds of the eastern Palearctic' at http://www.barcodinglife.org. All sequences have also been deposited in GenBank (Accession nos GQ481247 - GQ482920). A complete list of the museum catalog numbers, BOLD process identification numbers, and GenBank accession numbers for each specimen analyzed is included in Additional file 1.

We supplemented the data gathered in this study with sequences from North American congeners (accessible from the "Birds of North America - Phase II" project folder at http://www.barcodinglife.org) to examine divergences within transcontinental species and between sister species pairs. This added 849 sequences from 227 species, of which 66 species were shared with the Palearctic dataset. A list of BOLD process identification numbers and GenBank accession numbers for these sequences are listed in the Additional file 2. In total, 2,523 sequences from 559 species were included in the analyses.

Data analysis

To assess the discriminatory power of COI barcodes, we compared three different methods commonly deployed in DNA barcoding studies: neighbour-joining (NJ) clusters, distance-based thresholds, and character-based assignment. We avoided more computationally intensive methods in favour of programs that could be executed in real time. For the clustering method, we used MEGA version 3.1 [45] to construct an NJ tree using the Kimura 2 parameter distance model (K2P). More sophisticated tree-building methods exist, but since we are concerned about terminal branches, not deeper branching patterns, this method is sufficient. Support for monophyletic clusters was determined using 500 bootstrap replicates. Species were accepted as being monophyletic providing they comprised the smallest diagnosable cluster with greater than 95% bootstrap support [46]. Though bootstrap support cannot be determined for species represented by a single sequence, they were included in the analysis to observe if they created paraphyly in neighbouring taxa. Species that could be divided into two or more well-supported clusters were flagged as potentially cryptic taxa.

For the threshold-based approach, we blindly grouped sequences into provisional species clusters using a molecular operational taxonomic unit (MOTU) assignment program originally developed for nematodes [47]. The program, 'MOTU_define.pl' v2.07 (R. Floyd and M. Blaxter, unpublished; available from http://www.nematodes.org/bioinformatics/MOTU/index.shtml), clusters sequences together based on BLAST similarity using a user-defined base difference cut-off. Rather than use an arbitrary cut-off value, we determined the optimum threshold, or OT [36], by pooling our new data with the published North American bird dataset [3] and generating a cumulative error plot using all species with multiple representatives (see Figure 2). We adopted a liberal threshold of 11 base differences based on this result, which approximately equates to 1.6% divergence. Program parameters only included sequences greater than 500 bp with a minimum alignment overlap of 400 bp; however, this did not exclude any sequences from analysis.

Figure 2
figure 2

Cumulative error plots. Cumulative error plots of type I (false positive) and type II (false negative) errors for different divergence thresholds. Plot is based on 979 Holarctic bird species. The optimum threshold occurs at 1.6% divergence.

For the character-based identification method, we used the character assignment system CAOS, which automates the identification of conserved character states (in this case, different nucleotides) from a tree of pre-defined species [48]. The system comprises two programs: P-Gnome and P-Elf [48]. P-Gnome is used to identify the diagnostic sequence characters that separate species and uses them to generate a rule set for species identification; P-Elf classifies new sequences to species using the rule set. We used the programs PAUP v4.0b10 [49] and MESQUITE v2.6 [50] respectively to produce the input NJ trees and nexus files for P-Gnome in accordance with the CAOS manual. We executed P-Gnome using several subsets of our data. First, we tried all of the Palearctic species included in this study to determine if diagnostic characters could be identified to separate a wide range of species. The input tree for P-Gnome requires that all species nodes be collapsed to single polytomies, which is an arduous task for large numbers of species. We only used a single representative from each species to circumvent this issue with the drawback that intraspecific variation is ignored during rule generation. To test the character-based method on a finer scale, we ran the program independently on the three largest genera sampled: Emberiza (n = 23), Phylloscopus (n = 13), and Turdus (n = 13). For species with multiple representatives, the shortest sequence was omitted from rule generation and used later to test species assignment.

For the first two tests (NJ and MOTU), all species exhibiting type I error, wherein a single species produced two or more discernable clusters of sequences, were compiled. Additional lines of evidence (e.g. alternative genes, morphological differences, song differences, etc.) were sought from previous studies to support or refute the likelihood of species differences in such cases. However, no formal recommendations are made here. We also performed the two-cluster test using Lintre [51] to determine if sequences from these species had evolved in a clock-like manner. For type II errors, wherein multiple species grouped together to form one well-supported cluster, sequences from each cluster were run through P-Gnome to ascertain if diagnostic characters could be identified that distinguish these close species.

Results

Neighbour-joining clusters

Of the 559 species analyzed, 72 had only a single representative and thus no bootstrap support could be calculated. However, all of these formed independent branches on the NJ tree that did not compromise the identification of other species. The remaining species were categorized into four patterns (Figure 3). Ninety percent formed well-supported (> 95% bootstrap) monophyletic groups (Figure 3a), and an additional 4% were monophyletic but with less than 95% bootstrap support (Figure 3b). Ten species, 2% of the total, were paraphyletic (Larus canus, Thalasseus sandviciensis, Motacilla citreola, M. flava, Saxicola maurus, Sitta europaea, Certhia familiaris, Lanius collurio, L. excubitor, and Pica pica)(Figure 3c). The remaining taxa (4%) formed monophyletic clusters that contained two or more species (Figure 3d; Table 1). These were mostly limited to pairs of sister taxa, with the notable exception of one cluster containing 10 species in the Herring gull complex (Larus californicus, L. fuscus, L. glaucescens, L. glaucoides, L. heuglini, L. hyperboreus, L. occidentalis, L. smithsonianus, L. thayeri, and L. vegae).

Table 1 Species with limited COI divergence
Figure 3
figure 3

Divergence patterns of closely related species. Examples of divergence patterns illustrated in the NJ tree a) Species are monophlyletic with > 95% bootstrap support, b) Species are monophyletic, but is support is weak, c) Species are not monophyletic (i.e. paraphyly occurs), d) Multiple species form a single monophyletic group.

Forty-two species showed evidence of having divergent lineages (Table 2). Twenty-two species formed two or more well-supported (> 95% bootstrap) monophyletic clusters. Another four species formed two distinct clusters, but with one cluster possessing only 90-94% bootstrap support. These cases included 7 of the 10 paraphyletic species. In an additional 16 species, a single specimen was divergent from the rest, but further sampling is necessary to adequately evaluate these cases. Table 2 lists all species with divergent lineages. The total number of species recognized via this method is difficult to gauge due to inclusion of single representatives for some species and divergent lineages.

Table 2 Species bearing divergent COI lineages

Distance-based assignment

The MOTU analysis identified 570 clusters, or taxonomic units, versus the 559 recognized by traditional taxonomy. The similarity of these numbers disguises discrepancies in species assignment. Poor resolution occurred in 22 groups representing 61 species (Table 1). These lumped taxa, as with the NJ clustering method, were mostly limited to pairs of species, save for two triplets (Somateria spp. and Turdus spp.) and thirteen large white-headed gulls (Larus canus, L. delawarensis, L. marinus, and the aforementioned members of the Herring gull complex). Divergent groups were recognized in 42 species (Table 2); 95% of these overlapped with those recognized via NJ. Most were divided into two clusters, though three or more clusters were detected in five species. In two of the paraphyletic species (Motacilla flava, Lanius collurio), one lineage was lumped with a closely related species while the other lineage was divergent.

Character-based assignment

P-Gnome failed to produce a diagnostic rule set that that could distinguish all 398 species sequenced in this study. Results using subsets of the data were more successful. Complete diagnostic rule sets were generated and successfully tested for both Phylloscopus and Turdus. The rule set for Emberiza could not distinguish between sequences of E. leucocephalos and E. citrinella due to their near congruence. In addition, P-Elf failed to correctly identify single sequences from the species E. chrysophrys and E. elegans. The former sequence was short (594 bp) and might have lacked important diagnostic characters. However, the latter sequence was of typical length (694 bp) and only exceptional in that it contained 5 polymorphic sites from the sequence used to generate the rule set. Both of these species were incorrectly identified as E. aureola, though this identification would vary if the input tree were altered.

Of 22 groups of lumped species, all but five could be resolved using diagnostic characters (see Table 1). For example, the species pair Coturnix coturnix and C. japonica possessed 10 diagnostic nucleotide sites, two short of recognition by the MOTU threshold but still easily distinguishable. More complex rule sets were required when more species were involved (e.g. Aythya ducks). The remaining groups featured virtually no variation between species. These include 10 members of the herring gull complex (Larus spp.) and the species pairs Gallinago gallinago/G. delicata, Cuculus canorus/C. optatus, Carduelis flammea/C. hornemanni, and Emberiza citrinella/E. leucocephalos.

Discussion

Species boundaries in Palearctic Birds

Divergence levels between closely related species were highly variable, ranging from approximately 0-16%; however, some of these values may be inflated for under-sampled genera and families. Recent studies have detached rate variation in the mitochondrial genome from factors such as population size, body size, and other life-history traits [52–54]. While some authors contend that rate variation in birds is highly irregular [53], a recent thorough review demonstrated relatively minor variation and upheld the occurrence of clock-like evolution [55]. Consequently, we attribute the limited divergence between some sister species to recent speciation events. Studies documenting recent and rapid diversifications often address subspecific variants rather than full species [56, 57]. Still, low sequence divergence does not necessarily indicate that species should be synonymised [58]. Low sequence divergence is particularly common in superspecies complexes, including those divided between continents, but the species within them remain valid units for both ecological studies and conservation.

Four species pairs and the large white-headed gulls included in this study featured virtually no variation for COI and could not be distinguished using any of the approaches employed in this study. Low divergence in mitochondrial markers had been previously demonstrated in each of these cases. Lumping has been considered for some, including Carduelis flammea/hornemanni[59] and the recently split Gallinago gallinago/delicata[35], but more evidence is required. The cause of shared mitochondrial haplotypes between Cuculus canorus and C. optatus has not been resolved (hybrids have never been documented [60]), but their taxonomic distinction has been asserted based on song differences [61]. Emberiza citrinella and E. leucocephalos are exceptionally interesting in that they are the most phenotypically distinct of these pairs and a survey of nuclear markers revealed genetic divergence [62]. They are known to hybridize extensively and introgression is a likely explanation [62]. Species boundaries in the large white-headed gulls may have also been confused by contemporary hybridization, though shallow history and slowed rates of evolution have also been implicated [63, 64].

Nearly one tenth of the species (7.5%) analyzed in this study contained divergent mitochondrial lineages, with divergences averaging 3.6%. While divergence at a single mitochondrial gene alone is insufficient evidence to define new species boundaries, it is cause for new hypothesis testing. Several recently split species that are morphologically similar to their nearest relative, such as the swallow Riparia diluta and the warbler Locustella amnicola, represent taxa that barcodes would flag for closer scrutiny. Distributions of most of the divergent lineages in this study conform to one of four previously documented phylogeographic trends (summarized in Table 2): a unique lineage in the Caucasus region [65]; a unique lineage in the Sakhalin region [66]; divergent lineages divided into eastern and western populations [25]; divergent lineages on either side of the Bering Strait [19]. Species with multiple lineages can display more than one of these patterns. A few lineages appear to be parapatric, which could indicate areas of overlap or hybrid zones [67]. Past climate change and its effect on historical habitat distribution is likely responsible for shaping patterns of genetic divergence in modern populations, but whether or not these populations were divided by the same historical events is difficult to determine without dating divergence times. While the COI sequences mostly appear to be evolving in a clocklike fashion, dating is risky given the absence of adequate calibration points and the reliance on various assumptions [24, 55].

Most species exhibited surprisingly limited variation between Old World and New World populations. Of the approximately 140 species with Holarctic distributions, 43% are represented in this study. Only 11 of these 61 species (18%) possessed intraspecific divergences great enough to signal likely species-level differences by either the NJ or MOTU method. The Bering Sea has served a variable but clear role as a barrier to gene flow for birds, particularly non-marine species. Several trans-Beringian species have already been split in recent years, due partly to molecular evidence (e.g. Brachyramphus marmoratus/B. perdix[21], Picoides tridactylus/P. dorsalis[19], Pica pica/P. hudsoni[68]). Still, caution must be exercised when identifying species boundaries between allopatric populations. For example, one of the Palearctic Lanius excubitor specimens from this study appears to belong to the North American clade, suggesting that some modern exchange might occur between the continents. Though it is more common for Palearctic species to invade the Nearctic, the reverse pattern has also been observed [69]. Correct interpretation of this result requires further study with additional specimens.

This survey has identified a number of species that demand further taxonomic scrutiny (see Table 2). It is likely that some of the divergent lineages identified here represent distinct species. Of course, genetic distances do not always correspond to species limits [19, 69]. Alternative explanations for the divergent lineages observed include historical phylogeographic isolation, female-restricted dispersal, or male-biased gene flow [35]. The common phylogeographic patterns observed in many of the divergent lineages support the idea of historical isolation. Areas of secondary contact must be further studied to evaluate the gene flow between lineages [70]. In a few exceptional cases genetic lineages appear largely sympatric, including within Alauda arvensis, Delichon dasypus, and Phoenicurus phoenicurus. Nuclear copies of mitochondrial sequences (numts) are an unlikely explanation given the absence of stop codons and heterozygous peaks. Phoenicurus phoenicurus was also noted by Johnsen et al. [29], who attributed the aberrant phylogeographic pattern to admixture of historically separated lineages. This situation is paradoxical compared to suspected introgressed genomes used to explain limited divergence in sister species. Selective sweeps are frequently invoked to explain the limited variation observed in mitochondrial markers [6, 71], which raises the question of how two mtDNA lineages manage to persist in one species but not another. Ongoing research of species limits and evolutionary histories is clearly still necessary in the Palearctic.

Methods comparison

The MOTU assignment program used in this study was originally developed for meiofauna with few morphological characters [47]. Applying it to a group with better-established taxonomy allows more conclusive tests of its performance. Our results indicated a type II error rate of 10.9%, but this is inflated by the diversity of named white-headed gull species (Larus spp.); with these species eliminated, error is reduced to 8.8%. At this point, we don't consider type I errors a fault of this method since these cases are biologically interesting, do not necessarily impair identification, and may represent over-looked species [34, 35]. The major drawback to the program in its current form is the difficulty in associating any level of statistical support with species assignments, which may differ slightly depending on the input order of sequences. Although the program does allow a random re-sampling scheme, the output is not summarized, making statistical inference on the stability of taxonomic units virtually impossible. The major impediment now for biologists applying this method to microscopic invertebrates still lies in determining an operational threshold.

The use of a distance-based threshold technique has been a major point of contention in the DNA barcoding endeavour [37, 72, 73]. While COI variation represents a product of evolution, an arbitrary cut-off value does not reflect what is known about the evolutionary processes responsible for this variation. The threshold approach depends on the existence of a gap between levels of intraspecific variation and interspecific divergence, which opponents argue does not exist. Early success in identifying a "barcoding gap" in North American birds was attributed to insufficient sampling of closely related species [35, 37]. We found the original "10× rule" proposed by Hebert et al. [34] to be too conservative to recognize recently diverged species and opted for a more liberal threshold of 1.6%. While this value was more effective at species identification, some sister species exhibited little or no variation, which eliminates the possibility of identifying a gap. However, invalidating the use of distance-based methods based on the failure of thresholds might be going too far. Identifying the nearest matches to a query sequence is still useful, even if a conclusive assignment is not provided [74].

The development of an NJ profile for identification depends on the coalescence of species and not an arbitrary level of divergence [36]; in theory, species that failed recognition via the threshold approach may still be recognized. However, we found that the same species were typically problematic for both approaches (see Table 1). This is not surprising: high bootstrap support is unlikely when a slight aberration in the data would alter the results [75], which is the case when sequences are highly similar. Critics have argued that the bootstrap test for monophyly is simply too conservative and incorrectly rejects monophyly in too many cases [76]. This is apparent from the 4% of species that appear monophyletic but with limited support. Alternative forms of statistical support based on coalescent theory suggest that increased sampling decreases the risk of monophyly by chance, which would support the reality of these patterns despite low bootstrap values [77]. A modified NJ algorithm with non-parametric bootstrapping has been proposed to offer fast barcode-based identifications, but success still depends on the completeness of the reference database and weakly divergent species remain problematic [78].

The character-based method was effective, but did not feature the same scalability as the previous two methods. We found that the CAOS system was severely constrained by limits on the number of species that could be included for rule generation. More thorough benchmarking is necessary to determine the upper limits of the program, but at this point in time they are unclear. We also found that comprehensive sampling for each taxon is vital for accurate rules that account for intraspecific polymorphisms. When operating with smaller sets of taxa, the programs were successful in both identifying diagnostic characters and in subsequently identifying new sequences to species. However, we did find P-Elf to be highly susceptible to erroneous identifications for unrepresented species, counter to previous claims [79]. When using smaller datasets, sequences introduced from novel taxa were typically given a species level identification, even when those taxa derived from a different order (data not shown).

Both distance-based and clustering-based methods appear to share the same computational strengths, handling even large datasets quickly. However, both methods are also impaired by the same issues: limited divergence between sister taxa. The results of the character-based method appear to complement the former two methods. While it is precise and able to detect minor differences in closely related taxa [80], it is unable to handle large numbers of sequences. It is also susceptible to errors when the appropriate taxa have not been comprehensively sampled. When it comes to species identification, we propose that the best method might actually be a multi-tiered approach, where an initial method is used to narrow the identification to a select group of taxa and an alternate method is used to differentiate similar taxa. Similarly, Munch et al. [78] recommend incorporating methods that model population level variation to distinguish between closely allied species. For cases of limited divergence, sampling a longer stretch of COI or even alternative genes would increase support for identifications.

Conclusion

The utility of DNA barcodes in avian research is two-fold. Preliminary investigations, such as this, offer fresh insight to aid the ongoing effort to refine avian taxonomy. And secondly, a comprehensive library of COI sequences provides an invaluable tool for species assignment when differences in morphology are difficult to measure or otherwise assess. This includes species with cryptic morphological differences (e.g. Phylloscopus warblers, Calandrella larks, and Empidonax flycatchers) but also scenarios where identification is desired but only fragmentary remains are available (e.g. air strikes, nest contents, diet analysis, etc.). This study reaffirms these possibilities, demonstrating that COI sequence variation is largely congruent with species boundaries. Departures from this congruence are typically indicative of overlooked biological processes; historically separated lineages in the case of within species divergence, and recent or historical gene flow in the case of shared haplotypes between species. Molecular analysis is novel for some of these taxonomic groups or geographic areas, and the resultant observations highlight areas in need of further taxonomic study.

The efficacy of DNA barcodes for use in species assignment is dependent on two factors: the construction of thorough COI libraries and efficient tools to assign sequences to species. This study substantiates the need for dense taxonomic sampling. It further demonstrates that standardized gene libraries are easily amalgamated to examine geographically broad areas or taxonomically diverse groups. Current analytical methods for barcode data appear insufficient for handling recently evolved species. Though less of a problem for known cases of shallow divergence, where pairs of species may often be further scrutinized using a multi-tiered approach, these cases may be more problematic for those who wish to use barcodes as a tool to accelerate species discovery in poorly studied groups.

References

  1. Hebert PDN, Cywinska A, Ball SL, DeWaard JR: Biological identifications through DNA barcodes. Proceedings of the Royal Society of London Series B-Biological Sciences. 2003, 270: 313-321. 10.1098/rspb.2002.2218.

    Article  CAS  Google Scholar 

  2. Frezal L, Leblois R: Four years of DNA barcoding: Current advances and prospects. Infection Genetics and Evolution. 2008, 8: 727-736. 10.1016/j.meegid.2008.05.005.

    Article  CAS  Google Scholar 

  3. Kerr KCR, Stoeckle MY, Dove CJ, Weigt LA, Francis CM, Hebert PDN: Comprehensive DNA barcoding coverage of North American birds. Molecular Ecology Notes. 2007, 7: 535-543. 10.1111/j.1471-8286.2007.01670.x.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  4. Vilaça ST, Lacerda DR, Sari EHR, Santos FR: DNA-based identification applied to Thamnophilidae (Passeriformes) species: the first barcodes of Neotropical birds. Revista Brasileira de Ornitologia. 2006, 14: 7-13.

    Google Scholar 

  5. Chaves AV, Clozato CL, Lacerda DR, Sari EHR, Santos FR: Molecular taxonomy of brazilian tyrant-flycatchers (Passeriformes: Tyrannidae). Molecular Ecology Resources. 2008, 8: 1169-1177. 10.1111/j.1755-0998.2008.02218.x.

    Article  CAS  PubMed  Google Scholar 

  6. Kerr KCR, Lijtmaer DA, Barreira AS, Hebert PDN, Tubaro PL: Probing evolutionary patterns in Neotropical birds through DNA barcodes. PLoS One. 2009, 4: 6-10.1371/journal.pone.0004379.

    Article  Google Scholar 

  7. Yoo HS, Eah JY, Kim JS, Kim YJ, Min MS, Paek WK, Lee H, Kim CB: DNA barcoding Korean birds. Molecules and Cells. 2006, 22: 323-327.

    CAS  PubMed  Google Scholar 

  8. Watson DM: Diagnosable versus distinct: Evaluating species limits in birds. Bioscience. 2005, 55: 60-68. 10.1641/0006-3568(2005)055[0060:DVDESL]2.0.CO;2.

    Article  Google Scholar 

  9. Elias-Gutierrez M, Valdez-Moreno M: A now cryptic species of Leberis Smirnov, 1989 (Crustacea, Cladocera, Chydoridae) from the Mexican semi-desert region, highlighted by DNA barcoding. Hidrobiologica. 2008, 18: 63-74.

    Google Scholar 

  10. Gibbs J: Integrative taxonomy identifies new (and old) species in the Lasioglossum (Dialictus) tegulare (Robertson) species group (Hymenoptera, Halictidae). Zootaxa. 2009, 1-38.

    Google Scholar 

  11. Yassin A: Molecular and Morphometrical Revision of the Zaprionus tuberculatus Species Subgroup (Diptera: Drosophilidae), with Descriptions of Two Cryptic Species. Annals of the Entomological Society of America. 2008, 101: 978-988. 10.1603/0013-8746-101.6.978.

    Article  Google Scholar 

  12. Toews DPL, Irwin DE: Cryptic speciation in a Holarctic passerine revealed by genetic and bioacoustic analyses. Molecular Ecology. 2008, 17: 2691-2705. 10.1111/j.1365-294X.2008.03769.x.

    Article  CAS  PubMed  Google Scholar 

  13. Areta JI, Pearman M: Natural history, morphology, evolution, and taxonomic status of the earthcreeper Upucerthia saturatior (Furnariidae) from the Patagonian forests of South America. Condor. 2009, 111: 135-149. 10.1525/cond.2009.080009.

    Article  Google Scholar 

  14. Barker FK, Vandergon AJ, Lanyon SM: Assessment of species limits among yellow-breasted meadowlarks (Sturnella spp.) using mitochondrial and sex-linked markers. Auk. 2008, 125: 869-879. 10.1525/auk.2008.07148.

    Article  Google Scholar 

  15. Newton I: The Speciation and Biogeography of Birds. 2000, New York: Academic Press

    Google Scholar 

  16. Knox AG, Collinson M, Helbig AJ, Parkin DT, Sangster G: Taxonomic recommendations for British birds. Ibis. 2002, 144: 707-710. 10.1046/j.1474-919X.2002.00110.x.

    Article  Google Scholar 

  17. Li W, Zhang Y-y: Subspecific taxonomy of Ficedula parva based on sequences of mitochondrial cytochrome b gene. Zoological Research. 2004, 25: 127-131.

    CAS  Google Scholar 

  18. Illera JC, Richardson DS, Helm B, Atienza JC, Emerson BC: Phylogenetic relationships, biogeography and speciation in the avian genus Saxicola. Molecular Phylogenetics and Evolution. 2008, 48: 1145-1154. 10.1016/j.ympev.2008.05.016.

    Article  PubMed  Google Scholar 

  19. Zink RM, Rohwer S, Andreev AV, Dittmann DL: Trans-Beringia comparisons of mitochrondrial DNA differentiation in birds. Condor. 1995, 97: 639-649. 10.2307/1369173.

    Article  Google Scholar 

  20. Zink RM, Rohwer S, Drovetski S, Blackwell-Rago RC, Farrell SL: Holarctic phylogeography and species limits of Three-toed Woodpeckers. Condor. 2002, 104: 167-170. 10.1650/0010-5422(2002)104[0167:HPASLO]2.0.CO;2.

    Article  Google Scholar 

  21. Friesen VL, Piatt JF, Baker AJ: Evidence from cytochrome b sequences and allozymes for a 'new' species of alcid: The long-billed Murrelet (Brachyramphus perdix). Condor. 1996, 98: 681-690. 10.2307/1369851.

    Article  Google Scholar 

  22. Drovetski SV, Zink RM, Rohwer S, Fadeev IV, Nesterov EV, Karagodin I, Koblik EA, Red'kin YA: Complex biogeographic history of a Holarctic passerine. Proceedings of the Royal Society of London Series B-Biological Sciences. 2004, 271: 545-551. 10.1098/rspb.2003.2638.

    Article  Google Scholar 

  23. Pavlova A, Zink RM, Drovetski SV, Red'kin Y, Rohwer S: Phylogeographic patterns in Motacilla flava and Motacilla citreola: Species limits and population history. Auk. 2003, 120: 744-758. 10.1642/0004-8038(2003)120[0744:PPIMFA]2.0.CO;2.

    Article  Google Scholar 

  24. Pavlova A, Zink RM, Drovetski SV, Rohwer S: Pleistocene evolution of closely related sand martins Riparia riparia and R-diluta. Molecular Phylogenetics and Evolution. 2008, 48: 61-73. 10.1016/j.ympev.2008.03.030.

    Article  CAS  PubMed  Google Scholar 

  25. Zink RM, Pavlova A, Drovetski S, Rohwer S: Mitochondrial phylogeographies of five widespread Eurasian bird species. Journal of Ornithology. 2008, 149: 399-413. 10.1007/s10336-008-0276-z.

    Article  Google Scholar 

  26. Zink RM, Drovetski SV, Rohwer S: Selective neutrality of mitochondrial ND2 sequences, phylogeography and species limits in Sitta europaea. Molecular Phylogenetics and Evolution. 2006, 40: 679-686. 10.1016/j.ympev.2005.11.002.

    Article  CAS  PubMed  Google Scholar 

  27. Drovetski SV, Zink RM, Fadeev IV, Nesterov EV, Koblik EA, Red'kin YA, Rohwer S: Mitochondrial phylogeny of Locustella and related genera. Journal of Avian Biology. 2004, 35: 105-110. 10.1111/j.0908-8857.2004.03217.x.

    Article  Google Scholar 

  28. Pavlova A, Rohwer S, Drovetski SV, Zink RM: Different post-pleistocene histories of Eurasian parids. Journal of Heredity. 2006, 97: 389-402. 10.1093/jhered/esl011.

    Article  CAS  PubMed  Google Scholar 

  29. Johnsen A, Rindal E, Ericson PGP, Zuccon D, Kerr KCR, Stoeckle MY, Lifjeld JT: DNA barcoding of Scandinavian birds reveals divergent lineages in trans-Atlantic species. Journal of Ornithology. 2010,

    Google Scholar 

  30. Voelker G, Rohwer S, Outlaw DC, Bowie RCK: Repeated trans-Atlantic dispersal catalysed a global songbird radiation. Global Ecology and Biogeography. 2009, 18: 41-49. 10.1111/j.1466-8238.2008.00423.x.

    Article  Google Scholar 

  31. Reeves AB, Drovetski SV, Fadeev IV: Mitochondrial DNA data imply a stepping-stone colonization of Beringia by arctic warbler Phylloscopus borealis. Journal of Avian Biology. 2008, 39: 567-575. 10.1111/j.0908-8857.2008.04421.x.

    Article  Google Scholar 

  32. Ferri E, Barbuto M, Bain O, Galimberti A, Uni S, Guerrero R, Ferte H, Bandi C, Martin C, Casiraghi M: Integrated taxonomy: traditional approach and DNA barcoding for the identification of filarioid worms and related parasites (Nematoda). Frontiers in Zoology. 2009, 6: 1-10.1186/1742-9994-6-1.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  33. Barrett RDH, Hebert PDN: Identifying spiders through DNA barcodes. Canadian Journal of Zoology. 2005, 83: 481-491. 10.1139/z05-024.

    Article  CAS  Google Scholar 

  34. Hebert PDN, Stoeckle MY, Zemlak TS, Francis CM: Identification of birds through DNA barcodes. Plos Biology. 2004, 2: 1657-1663. 10.1371/journal.pbio.0020312.

    Article  CAS  Google Scholar 

  35. Baker AJ, Tavares ES, Elbourne RF: Countering criticisms of single mitochondrial DNA gene barcoding in birds. Molecular Ecology Resources. 2009, 9: 257-267. 10.1111/j.1755-0998.2009.02650.x.

    Article  CAS  PubMed  Google Scholar 

  36. Wiemers M, Fiedler K: Does the DNA barcoding gap exist? - a case study in blue butterflies (Lepidoptera: Lycaenidae). Frontiers in Zoology. 2007, 4: 8-10.1186/1742-9994-4-8.

    Article  PubMed Central  PubMed  Google Scholar 

  37. Moritz C, Cicero C: DNA barcoding: Promise and pitfalls. Plos Biology. 2004, 2: 1529-1531. 10.1371/journal.pbio.0020354.

    Article  CAS  Google Scholar 

  38. Zhang AB, Sikes DS, Muster C, Li SQ: Inferring species membership using DNA sequences with back-propagation neural networks. Systematic Biology. 2008, 57: 202-215. 10.1080/10635150802032982.

    Article  CAS  PubMed  Google Scholar 

  39. Abdo Z, Golding GB: A step toward barcoding life: A model-based, decision-theoretic method to assign genes to preexisting species groups. Systematic Biology. 2007, 56: 44-56. 10.1080/10635150601167005.

    Article  CAS  PubMed  Google Scholar 

  40. Matz MV, Nielsen R: A likelihood ratio test for species membership based on DNA sequence data. Philosophical Transactions of the Royal Society B-Biological Sciences. 2005, 360: 1969-1974. 10.1098/rstb.2005.1728.

    Article  PubMed Central  CAS  Google Scholar 

  41. DeSalle R: Species discovery versus species identification in DNA barcoding efforts: response to Rubinoff. Conservation Biology. 2006, 20: 1545-1547. 10.1111/j.1523-1739.2006.00543.x.

    Article  PubMed  Google Scholar 

  42. Rach J, DeSalle R, Sarkar IN, Schierwater B, Hadrys H: Character-based DNA barcoding allows discrimination of genera, species and populations in Odonata. Proceedings of the Royal Society B-Biological Sciences. 2008, 275: 237-247. 10.1098/rspb.2007.1290.

    Article  PubMed Central  CAS  Google Scholar 

  43. Clements JF: The Clements checklist of the birds of the world. 2007, New York: Cornell University Press, 6

    Google Scholar 

  44. Haring E, Gamauf A, Kryukov A: Phylogeographic patterns in widespread corvid birds. Molecular Phylogenetics and Evolution. 2007, 45: 840-862.

    Article  CAS  PubMed  Google Scholar 

  45. Kumar S, Tamura K, Nei M: MEGA3: Integrated software for molecular evolutionary genetics analysis and sequence alignment. Briefings in Bioinformatics. 2004, 5: 150-163. 10.1093/bib/5.2.150.

    Article  CAS  PubMed  Google Scholar 

  46. Felsenstein J: Confidence limits on phylogenies - an approach using the bootstrap. Evolution. 1985, 39: 783-791. 10.2307/2408678.

    Article  Google Scholar 

  47. Floyd R, Abebe E, Papert A, Blaxter M: Molecular barcodes for soil nematode identification. Molecular Ecology. 2002, 11: 839-850. 10.1046/j.1365-294X.2002.01485.x.

    Article  CAS  PubMed  Google Scholar 

  48. Sarkar IN, Planet PJ, Desalle R: CAOS software for use in character-based DNA barcoding. Molecular Ecology Resources. 2008, 8: 1256-1259. 10.1111/j.1755-0998.2008.02235.x.

    Article  CAS  PubMed  Google Scholar 

  49. Swofford DL: PAUP*: Phylogenetic analysis using parsimony (*and other methods). Version 4 edition. 2002, Sunderland, Massachusetts: Sinauer Associates

    Google Scholar 

  50. Maddison WP, Maddison DR: Mesquite: a modular system for evolutionary analysis. 2009, [http://mesquiteproject.org]2.6

    Google Scholar 

  51. Takezaki N, Rzhetsky A, Nei M: Phylogenetic test of the molecular clock and linearized trees. Molecular Biology and Evolution. 1995, 12: 823-833.

    CAS  PubMed  Google Scholar 

  52. Bazin E, Glemin S, Galtier N: Population size does not influence mitochondrial genetic diversity in animals. Science. 2006, 312: 570-572. 10.1126/science.1122033.

    Article  CAS  PubMed  Google Scholar 

  53. Nabholz B, Glemin S, Galtier N: The erratic mitochondrial clock: variations of mutation rate, not population size, affect mtDNA diversity across birds and mammals. Bmc Evolutionary Biology. 2009, 9: 54-10.1186/1471-2148-9-54.

    Article  PubMed Central  PubMed  Google Scholar 

  54. Nabholz B, Mauffrey JF, Bazin E, Galtier N, Glemin S: Determination of mitochondrial genetic diversity in mammals. Genetics. 2008, 178: 351-361. 10.1534/genetics.107.073346.

    Article  PubMed Central  PubMed  Google Scholar 

  55. Weir JT, Schluter D: Calibrating the avian molecular clock. Molecular Ecology. 2008, 17: 2321-2328. 10.1111/j.1365-294X.2008.03742.x.

    Article  CAS  PubMed  Google Scholar 

  56. Mila B, McCormack JE, Castaneda G, Wayne RK, Smith TB: Recent postglacial range expansion drives the rapid diversification of a songbird lineage in the genus Junco. Proceedings of the Royal Society B-Biological Sciences. 2007, 274: 2653-2660. 10.1098/rspb.2007.0852.

    Article  PubMed Central  CAS  Google Scholar 

  57. Mila B, Smith TB, Wayne RK: Speciation and rapid phenotypic differentiation in the yellow-rumped warbler Dendroica coronata complex. Molecular Ecology. 2007, 16: 159-173. 10.1111/j.1365-294X.2006.03119.x.

    Article  CAS  PubMed  Google Scholar 

  58. Joseph L, Omland KE: Phylogeography: its development and impact in Australo-Papuan ornithology with special reference to paraphyly in Australian birds. Emu. 2009, 109: 1-23. 10.1071/MU08024.

    Article  Google Scholar 

  59. Marthinsen G, Wennerberg L, Lifjeld JT: Low support for separate species within the redpoll complex (Carduelis flammea-hornemanni-cabaret) from analyses of mtDNA and microsatellite markers. Molecular Phylogenetics and Evolution. 2008, 47: 1005-1017. 10.1016/j.ympev.2008.03.027.

    Article  CAS  PubMed  Google Scholar 

  60. Sorenson MD, Payne RB: A molecular genetic analysis of cuckoo phylogeny. Bird Families of the World: Cuckoos. Edited by: Payne RB. 2005, New York: Oxford University Press, 68-94.

    Google Scholar 

  61. Payne RB: Bird Families of the World: Cuckoos. 2005, New York: Oxford University Press

    Google Scholar 

  62. Irwin DE, Rubstov AS, Panov EV: Mitochondrial introgression and replacement between yellowhammers (Emberiza citrinella) and pine buntings (E. leucocephalos; Aves, Passeriformes). Biological Journal of the Linnean Society. 2009, 98: 422-438. 10.1111/j.1095-8312.2009.01282.x.

    Article  Google Scholar 

  63. Crochet PA, Desmarais E: Slow rate of evolution in the mitochondrial control region of gulls (Aves: Laridae). Molecular Biology and Evolution. 2000, 17: 1797-1806.

    Article  CAS  PubMed  Google Scholar 

  64. Liebers D, de Knijff P, Helbig AJ: The herring gull complex is not a ring species. Proceedings of the Royal Society of London Series B-Biological Sciences. 2004, 271: 893-901. 10.1098/rspb.2004.2679.

    Article  Google Scholar 

  65. Hewitt G: The genetic legacy of the Quaternary ice ages. Nature. 2000, 405: 907-913. 10.1038/35016000.

    Article  CAS  PubMed  Google Scholar 

  66. Zink RM, Drovetski SV, Rohwer S: Phylogeographic patterns in the great spotted woodpecker Dendrocopos major across Eurasia. Journal of Avian Biology. 2002, 33: 175-178. 10.1034/j.1600-048X.2002.330208.x.

    Article  Google Scholar 

  67. Aliabadian M, Roselaar CS, Nijman V, Sluys R, Vences M: Identifying contact zone hotspots of passerine birds in the Palaearctic region. Biology Letters. 2005, 1: 21-23. 10.1098/rsbl.2004.0258.

    Article  PubMed Central  PubMed  Google Scholar 

  68. Banks RC, Cicero C, Dunn JL, Kratter AW, Ouellet H, Rasmussen PC, Remsen JV, Rising JA, Stotz DF: Forty-second supplement to the American Ornithologists' Union check-list of North American birds. Auk. 2000, 117: 847-858. 10.1642/0004-8038(2000)117[0847:FSSTTA]2.0.CO;2.

    Article  Google Scholar 

  69. Zink RM, Pavlova A, Rohwer S, Drovetski SV: Barn swallows before barns: population histories and intercontinental colonization. Proceedings of the Royal Society B-Biological Sciences. 2006, 273: 1245-1251. 10.1098/rspb.2005.3414.

    Article  PubMed Central  Google Scholar 

  70. Moritz C, Hoskin CJ, MacKenzie JB, Phillips BL, Tonione M, Silva N, VanDerWal J, Williams SE, Graham CH: Identification and dynamics of a cryptic suture zone in tropical rainforest. Proceedings of the Royal Society B-Biological Sciences. 2009, 276: 1235-1244. 10.1098/rspb.2008.1622.

    Article  PubMed Central  CAS  Google Scholar 

  71. Irwin DE, Rubstov AS, Panov EV: Mitochondrial introgression and replacement between yellowhammers (Emberiza citrinella) and pine buntings (E. leucocephalos; Aves, Passeriformes). Biological Journal of the Linnean Society. 2009, 92: 422-438. 10.1111/j.1095-8312.2009.01282.x.

    Article  Google Scholar 

  72. Hickerson MJ, Meyer CP, Moritz C: DNA barcoding will often fail to discover new animal species over broad parameter space. Systematic Biology. 2006, 55: 729-739. 10.1080/10635150600969898.

    Article  PubMed  Google Scholar 

  73. Meyer CP, Paulay G: DNA barcoding: Error rates based on comprehensive sampling. Plos Biology. 2005, 3: 2229-2238.

    CAS  Google Scholar 

  74. Ratnasingham S, Hebert PDN: BOLD: The Barcode of Life Data System. Molecular Ecology Notes. 2007, 7: 355-364. 10.1111/j.1471-8286.2007.01678.x. [http://www.barcodinglife.org]

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  75. Holmes S: Bootstrapping phylogenetic trees: theory and methods. Statistical Science. 2003, 18: 241-255. 10.1214/ss/1063994979.

    Article  Google Scholar 

  76. Rodrigo AG: Calibrating the bootstrap test of monophyly. International Journal for Parasitology. 1993, 23: 507-514. 10.1016/0020-7519(93)90040-6.

    Article  CAS  PubMed  Google Scholar 

  77. Rosenberg NA: Statistical tests for taxonomic distinctiveness from observations of monophyly. Evolution. 2007, 61: 317-323. 10.1111/j.1558-5646.2007.00023.x.

    Article  PubMed  Google Scholar 

  78. Munch K, Boomsma W, Willerslev E, Nielsen R: Fast phylogenetic DNA barcoding. Philosophical Transactions of the Royal Society B-Biological Sciences. 2008, 363: 3997-4002. 10.1098/rstb.2008.0169.

    Article  PubMed Central  CAS  Google Scholar 

  79. Kelly RP, Sarkar IN, Eernisse DJ, Desalle R: DNA barcoding using chitons (genus Mopalia). Molecular Ecology Notes. 2007, 7: 177-183. 10.1111/j.1471-8286.2006.01641.x.

    Article  CAS  Google Scholar 

  80. Wong EHK, Shivji MS, Hanner RH: Identifying sharks with DNA barcodes: assessing the utility of a nucleotide diagnostic approach. Molecular Ecology Resources. 2009, 9: 243-256. 10.1111/j.1755-0998.2009.02653.x.

    Article  CAS  PubMed  Google Scholar 

  81. Efe MA, Tavares ES, Baker AJ, Bonatto SL: Multigene phylogeny and DNA barcoding indicate that the Sandwich tern complex (Thalasseus sandvicensis, Laridae, Sternini) comprises two species. Molecular Phylogenetics and Evolution. 2009, 52: 263-267. 10.1016/j.ympev.2009.03.030.

    Article  CAS  PubMed  Google Scholar 

  82. Koopman NE, McDonald DB, Hayward GD, Eldegard K, Sonerud GA, Sermach SG: Genetic similarity among Eurasian subspecies of boreal owls Aegolius funereus. Journal of Avian Biology. 2005, 36: 179-183. 10.1111/j.0908-8857.2005.03509.x.

    Article  Google Scholar 

  83. Zink RM, Pavlova A, Drovetski S, Wink M, Rohwer S: Taxonomic status and evolutionary history of the Saxicola torquata complex. Molecular Phylogenetics and Evolution. 2009, 52: 769-773. 10.1016/j.ympev.2009.05.016.

    Article  CAS  PubMed  Google Scholar 

  84. Irwin DE, Bensch S, Price TD: Speciation in a ring. Nature. 2001, 409: 333-337. 10.1038/35053059.

    Article  CAS  PubMed  Google Scholar 

  85. Packert M, Martens J, Severinghaus LL: The Taiwan Firecrest (Regulus goodfellowi) belongs to the Goldcrest assemblage (Regulus regulus s. l.): evidence from mitochondrial DNA and the territorial song of the Regulidae. Journal of Ornithology. 2009, 150: 205-220. 10.1007/s10336-008-0335-5.

    Article  Google Scholar 

  86. Kvist L, Martens J, Higuchi H, Nazarenko AA, Valchuk OP, Orell M: Evolution and genetic structure of the great tit (Parus major) complex. Proceedings of the Royal Society of London Series B-Biological Sciences. 2003, 270: 1447-1454. 10.1098/rspb.2002.2321.

    Article  Google Scholar 

  87. Akimova A, Haring E, Kryukov S, Kryukov A: First insights into a DNA sequence based phylogeny of the Eurasian Jay Garrulus glandarius. Russian Journal of Ornithology. 2007, 16: 567-575.

    Google Scholar 

Download references

Acknowledgements

We thank Rob Faucett, Chris Wood, Sievert Rohwer, Annamaria Clark, Stephanie Sundier, Kevin Epperly, and Elizabeth Broughton of the Burke Museum, University of Washington, Eugeniy Yakhontov, and Aleksandra Panyutina of Moscow State University for aiding the assembly of specimens for analysis, in addition to the original field collectors and collaborators, who make large-scale studies such as this possible. We thank Nataly Ivanova and staff at the Biodiversity Institute of Ontario for assistance with DNA sequencing. We thank Robin Floyd for assistance with the MOTU program and Naoko Takezaki for assistance with Lintre. We thank Jan Lifjeld for useful input on species taxonomy and John Wilson, Bob Hanner, Mark Stoeckle, and two anonymous reviewers for thoughtful commentary on earlier drafts of this manuscript. This research was funded by grants from NSERC and Genome Canada via the Ontario Genomics Institute to Paul Hebert.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kevin CR Kerr.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors' contributions

KCRK coordinated the study, carried out the molecular and statistical analyses, and drafted the manuscript. SMB and MVK provided specimens for the study, participated in its coordination, and helped with the manuscript. YAR and EAK contributed to the interpretation of the results and helped with the manuscript. PDNH conceived of the study, participated in its design and coordination, and helped with the manuscript. All authors read and approved the final manuscript.

Electronic supplementary material

12983_2009_117_MOESM1_ESM.DOC

Additional file 1: List of sampled specimens. Complete list of museum accession numbers, BOLD process identification numbers, and GenBank accession numbers for each specimen analyzed in this study. (DOC 2 MB)

12983_2009_117_MOESM2_ESM.DOC

Additional file 2: List of sequences acquired from BOLD. Complete list of BOLD process identification numbers and GenBank accession numbers for all sequences used in this study from the "Birds of North America - Phase II" project in BOLD. (DOC 990 KB)

Authors’ original submitted files for images

Rights and permissions

This article is published under license to BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Kerr, K.C., Birks, S.M., Kalyakin, M.V. et al. Filling the gap - COI barcode resolution in eastern Palearctic birds. Front Zool 6, 29 (2009). https://doi.org/10.1186/1742-9994-6-29

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1742-9994-6-29

Keywords