Skip to main content
  • Research article
  • Open access
  • Published:

Site specific rates of mitochondrial genomes and the phylogeny of eutheria

Abstract

Background

Traditionally, most studies employing data from whole mitochondrial genomes to diagnose relationships among the major lineages of mammals have attempted to exclude regions that potentially complicate phylogenetic analysis. Components generally excluded are 3rd codon positions of protein-encoding genes, the control region, rRNAs, tRNAs, and the ND6 gene (encoded on the opposite strand). We present an approach that includes all the data, with the exception of the control region. This approach is based on a site-specific rate model that accommodates excessive homoplasy and that utilizes secondary structure as a reference for proper alignment of rRNAs and tRNAs.

Results

Mitochondrial genomic data for 78 eutherian mammals, 8 metatherians, and 3 monotremes were analyzed with a Bayesian analysis and our site specific rate model. The resultant phylogeny revealed strong support for most nodes and was highly congruent with more recent phylogenies based on nuclear DNA sequences. In addition, many of the conflicting relationships observed by earlier mitochondrial-based analyses were resolved without need for the exclusion of large subsets of the data.

Conclusion

Rather than exclusion of data to minimize presumed noise associated with non-protein encoding genes in the mitochondrial genome, our results indicate that selection of an appropriate model that accommodates rate heterogeneity across data partitions and proper treatment of RNA genes can result in a mitochondrial genome-based phylogeny of eutherian mammals that is reasonably congruent with recent phylogenies derived from nuclear genes.

Background

The class Mammalia provides a classic example of an adaptive radiation, characterized by a proliferation of lineages displaying a diverse array of ecomorphological specializations for feeding and locomotion [1]. Many additional biological attributes (e.g., behavior, physiology), coupled with this diversity in form and function, have allowed mammals to exploit a broad range of habitats worldwide. There are approximately 135 families of living mammals apportioned into 26 orders and two major subclasses, Prototheria and Theria, with the former subclass containing the order Monotremata (duck-billed platypus and spiny-anteaters) and the latter containing the infraclasses Metatheria (marsupials) and Eutheria (placentals), which are subdivided into 7 and 18 orders, respectively [2, 3]. Lineage-specific rate heterogeneity in terms of morphological diversification [4] and molecular divergence [5–7] is a trademark of the various orders and families of mammals, especially within the Eutheria, and this has complicated efforts to resolve phylogenetic relationships among the higher categories of mammals.

Until relatively recently, most contributions to the "mammal tree of life," as it relates to phylogeny and classification, were made by functional morphologists and paleontologists [2, 8–10]. More recent molecular efforts have resulted in confirmation of some previous hypotheses, the refutation of others, and the proposal of novel arrangements [10–13].

The most severe disagreements between morphology and molecules originated from studies based on mitochondrial genome sequences. For example, monophyly of Rodentia (the most speciose order of mammals) is based on a combination of dentition, skull morphology, soft anatomy, the postcranial skeleton, and the jaw mechanism [14], and early classifications never questioned the naturalness of this clade. Nevertheless, several early studies of nuclear genes [15–17] and mitochondrial genomes [18–20] argued that guinea pigs and presumably their relatives (hystricognath rodents from South America and Africa) were "not rodents," but represented a separate and more basal eutherian lineage, apart from muroid rodents (rats and mice). These same data challenged the monophyly of Glires, a group recognized on the basis of morphology [10, 21] and containing the orders Lagomorpha (rabbits) and Rodentia, by suggesting a sister-group relationship between lagomorphs and primates [22]. The morphological placement of the order Xenarthra (armadillos, sloths, and anteaters) at the base of the eutherian radiation was also challenged, with mitochondrial data suggesting either the Erinaecidae [hedgehogs; [23]] or rodents at the base. In contrast to the morphology, xenarthrans were considered sister to a clade containing the orders Carnivora, Perrisodactyla (horses, rhinos, and elephants), Artiodactyla (pigs, antelope, deer, camels, etc.), and Cetacea (whales and dolphins) [24]. Two of the more startling results from the analysis of mitochondrial genomes included: 1) the placement the order Monotremata as sister to Metatheria, thus making the subclass Theria paraphyletic [25], and 2) a sister-group relationship between the anthropoid primates and Dermoptera (flying lemurs), thus rendering the order Primates paraphyletic [26]. Neither of these hypotheses is supported from either other molecular data or morphology [9, 10, 27–29].

More extensive studies employing greater taxon sampling as well as larger amounts of nucleotide sequence data from mitochondrial RNA (primarily rRNA) and/or nuclear genes [30–38] have resulted in higher levels of congruence with earlier morphological studies, including increased support for a more basal position of Xenarthra, the monophyly of Rodentia, Glires, and Primates, a monophyletic Theria, the Paenungulata (containing elephants, hyraxes, and sirenians), Tetytheria (elephants and sirenians), and Euarchonta (Scandentia, Dermoptera, Primates).

In contrast to recent studies employing primarily nuclear DNA sequences, a more recent study of whole mitochondrial genomes [26] failed to retrieve many of the well-supported clades identified by nuclear gene studies. Springer et al.'s [36] comparison of mitochondrial and nuclear gene sequences implied that mitochondrial data are less effective at resolving relationships at deeper nodes of the mammalian tree, and in many cases mitochondrial sequences failed to recover "benchmark clades," that are well-supported by both morphology and nuclear genes. In this particular comparison, nuclear genes apparently outperformed mitochondrial genomes because they evolve at a rate appropriate for resolving more divergent relationships among major lineages of mammals.

Unless mitochondrial genomes are evolving at rates where saturation becomes a problem at deeper nodes, one would expect the inclusion of analytical procedures that accommodate asymmetries observed for mtDNA [29, 39–42], coupled with appropriate placement of the root of the eutherian tree [30, 40, 43] and increased taxon sampling [44–47], to result in mitochondrial phylogenies that are more congruent with the consensus reached by nuclear genes. For the most part, a consideration of these factors has improved more recent results, primarily because model-based analyses of more mitochondrial genomes were employed [41]. Nevertheless, as with earlier studies employing whole mitochondrial genomes, Reyes et al. [41] excluded several regions of the genome prior to analysis with a model that accommodated multiple rates of substitution. For instance, 3rd codon positions, first positions involving leucine, and the control region are generally excluded to reduce homoplasy resulting saturation effects. The ND6 gene, encoded on the L-strand, is omitted because of presumed differences in constraints (e.g., base composition) relative to genes encoded on the H-strand. Finally, ribosomal genes (rRNAs) and transfer RNAs (tRNAs) are frequently left out, presumably because they are difficult to align.

It is our contention that exclusion of data is unnecessary if appropriate model-based analyses are employed. If fast evolving sites like 3rd codon positions can be appropriately modeled, then there is little reason for excluding them from a likelihood-based analysis. Similarly, if rRNAs and tRNAs can be reasonably well aligned with secondary structure, we see little justification for excluding these characters. In this paper we provide an analysis of whole mitochondrial genomes from 89 mammalian taxa and investigate relationships among major lineages of eutherians. Except for the control region, which is difficult to align across highly divergent taxa, all sequences were used in an analysis employing a pseudoreplicate-generated, site-specific rate model, first proposed by Kjer et al. [48]. Our major goal is to evaluate the effectiveness of this model to negate a prior exclusion of potentially useful data, and we base our conclusions on comparison of results to more extensive studies based on a large panel of nuclear gene sequences and extensive taxon sampling.

Results

The annotated Nexus file consists of 14,740 nucleotides, includes 3,783 amino acid characters as well as additional taxa (not used in this analysis), and is available on Kjer's website [49]. The Nexus file on the website includes character set definitions ("charsets") that allow the user to identify and analyze single gene partitions, codon positions, and rate classes separately, and taxon set definitions ("taxsets") that allow the user to evaluate relationships among specific taxa. The most likely tree from the Bayesian analysis is shown in Fig. 1. This phylogeny reveals strong support for several major groups of eutherians including: 1) a monophyletic Afrotheria, a basal clade containing Proboscidea (elephants), Sirenia (manatees and dugongs), Hyracoidea (hyraxes), Macroscelidea (elephant shrews), Tubulidentata (aardvarks), Afrosoricidea (insectivore families Chrysochloridae or golden moles and Tenrecidae or tenrecs); 2) a monophyletic Xenarthra sister to Afrotheria; 3) Euarchontoglires represented by two major clades, one containing the Primates (including Anthropoidea, Tarsiformes, and Lemuriformes), with Dermoptera (flying lemurs) nested inside, and the other containing a monophyletic Glires (rabbits and rodents); 4) euarchontan order Scandentia (tree shrews) sister to the two major groups of Euarchontoglires; 5) Laurasiatheria containing a paraphyletic Eulipotyphyla (representing the insectivore families Erinaceidae and Soricidae, and Talpidae), Chiroptera (bats), Pholidota (pangolins), Cetartiodactyla (Artiodactyla and Cetacea or whales and dolphins), Perrisodactyla (horses, rhinos, tapirs), and Carnivora; 6) a sister-group relationship between Euarchontoglires and Laurasiatheria. In addition to these major clades, monophyly of Paenungulata (containing the orders Proboscidea, Sirenia, and Hyracoidea), Tethytheria (Sirenia and Proboscidea), and Cetartiodactyla (Artiodactyla and Cetacea) with cetaceans sister to hippo is strongly supported.

Figure 1
figure 1

Most likely phylogram derived from the Bayesian Analysis (-ln 533753.675). Numerals indicate estimated posterior probability. These values are either placed on top of the node they represent (or with arrows pointing to the top of the internode) or directly to the left of the node. Nodes without numerals are supported at 100%. Higher taxa are indicated either on top of their representative internode, directly to the left of the node or to the right of the clade, and are delimited with brackets.

Table 1 shows the number of characters in each class, the rescaled consistency indices (RC), the mean model parameters and rate classes associated with the six partitions. The RC values show that the rate classes are very different in terms of how well the data map onto the tree. The fastest rate class is C-T rich (80%), just as C-T transitions are the fastest substitution class while slower rate classes are much less biased in terms of nucleotide composition (Table 1). Among site rate variation is most pronounced at the slowest and the fastest rate classes. Figure 2 shows a characterization of the partitions in terms of codon position and RNAs. RNA sequences tended to be conservative, and in terms of rates were similar to 2nd codon positions of protein-encoding genes. As expected, 3rd codon positions were associated with the faster rate classes, although a portion of 3rd positions evolved slowly (approximately 200 in rate classes 3–6). There were more parsimony-informative RNA characters (786), as well as first and second codon position characters (1928) in the "fast" rate class 2, than in rate class 6 (the slowest; 197 parsimony informative rRNA sites, and 258 parsimony informative 1st and 2nd codon sites). There were about the same number of variable RNA characters in rate class 6 (532) as there were second codon sites (543). We note that many 1st and 2nd codon sites are fast-evolving (2,206 in the fastest two rate classes), and 186 parsimony-informative (of 1800) RNA characters that have been discarded from other analyses are members of the slowest rate class, which is comparable to 131 (of 2541) parsimony-informative second codon positions in rate class 6.

Table 1 Mean model parameters and six character partitions and rate classes
Figure 2
figure 2

Rate Classes and Partition of Variable Sites – Top: A visualization with pie graphs of the proportion of sites in each rate-class partition that are RNAs (white), first codon positions (light grey), second codon positions (dark grey), and third codon positions (black). Rate classes are listed across the top, from fastest (class 1) to slowest (class 6). Bottom: A bar-graph visualization of the numbers of each of these classes among partitions, using the same color coding, as indicated in the key. Constant sites, found only in rate class six, are indicated with hatched bars. Raw numbers of each of the values in the bar graph are given below the bars. Fifteen sites from the origin belong in rate class 6, one in rate class 4, and two in rate class 3 (not shown).

Discussion

This analysis shows that third codon positions, redundant first codon (leucine) positions, the ND6, and the RNA genes can be included in a combined model-based analysis without drastically contradicting the general consensus from previous molecular studies. In fact, all benchmark clades for eutherian mammals that could be compared to the list provided by Springer et al. [36] were retrieved in our analysis and received high support. These benchmark clades include (all posterior probabilities 100): 1) Carnivora (Feliformia + Caniformia); 2) Cetacea (toothed whales and dolphins + baleen whales); 3) Cetartiodactyla (Artiodactyla + Cetacea); 4) Chiroptera (bats); 5) Diprotodontia (wombats, wallaroos, and brush-tailed possums); 6) Paenungulata (hyrax + elephants and Sirenia); 7) Perrisodactyla (rhino and tapir + horses); 8) Rumantia (bovines, sheep, deer); and 9) Xenarthra (armadillo + tamandua). The mitochondrial genome-based phylogeny shown in Fig. 1 is congruent with previous nuclear gene studies [32–34, 50] in several respects. Although placement of the root varies among studies [51], the nuclear gene studies and our study place the groups Afrotheria and Xenarthra at the base of the eutherian phylogeny followed by a sister-group relationship between the monophyletic groups Euarchontoglires and Laurasiatheria (collectively called the Boreoeutheria). Several other monophyletic groups appear to be well-supported and congruent between our mtDNA and previous nuclear DNA studies including Paenungulata (Hyracoidea, Sirenia, and Proboscidae), Cetartiodactyla (Artiodactyla and Cetacea), Chiroptera, and Glires (Lagomorpha and Rodentia).

Although several groups are identified by both our whole mitochondrial genome analysis and nuclear genes, not all of these molecularly-defined groups are necessarily congruent with morphological data. For instance, some morphological studies support a monophyletic Archonta containing the euarchontans as well as Chiroptera [9, 10], and although a relationship between the orders Artiodactyla and Cetacea has support from morphology, a sister-group relationship between Cetacea and the family Hippopotamidae (hippos) denoted by both nuclear genes and mitochondrial genomes [52] is supported by some [53] but not all morphological analyses [54, 55]. Some earlier morphological comparisons [9], but none of the molecular data, support Volitantia, a group containing Chiroptera and Dermoptera. More recent molecular studies, including the one presented here, have indicated paraphyly for the chiropteran suborder Microchiroptera with the family Rhinolophidae grouping closer to the Megachiroptera, a clade containing non-echolocating taxa [56–58], and this is not corroborated by morphological data.

Our phylogenetic results are similar to those presented by Reyes et al. [41], which was based on a GTR+I+G Bayesian analysis that excluded RNAs, ND6, and redundant codon positions. Gibson et al. [39] also showed that there were lineage and gene specific biases of C and T compositions, and performed an analysis with a model that reduced the character complexity of these nucleotides to Y, creating a three-state model. While Gibson et al. [39] included RNAs, they also excluded third codon positions and the ND6, resulting in a dataset of 7,402 sites. While we agree with the corrections proposed by both Gibson et al. [39] and Reyes et al. [41] in reducing the influence of homoplastic and biased characters, our approach differed in including a site specific rate model that rendered noisy sites less influential at deeper nodes, while retaining them as characters toward the tips of the tree. Our matrix is nearly twice the size of the largest previous analyses. In performing the pseudoreplicate reweighting, the noisiest sites are presumably identified and accommodated in a model. Many different partitions, including those that were excluded by others, can be explored by downloading the Nexus file and including specific "charsets" such as the ND6. For example, a parsimony analysis of the ND6 gene results in the recovery of therians, metatherians, eutherians, anthropoid primates (in the same order as the combined analysis), whales, and carnivores, among other groups (not shown). Clearly, the ND6 contains some non-random signal, including 26% of its 535 nucleotides in rate class 6 (the slowest).

The trees in our analysis of the combined data differ from others in the placement of Xenarthra; ours with Afrotheria (Fig. 1), supporting a northern-southern hemisphere split, and Gibson et al. [39] and Reyes et al. [41] with Euarchontoglires. Note, both this analysis and the analysis of Gibson et al. [39] compensate for the large number of homoplastic C-T transitions but in different ways. Kriegs et al. [59], using retrotransposed elements (which they suppose to be "homoplasy free"), supported Xenarthra as the sister taxon of the rest of Eutheria. While we agree that the two retrotransposed elements supporting this relationship are exceedingly strong characters, we prefer to consider the independent loss of these in the sloth and the armadillo as "possible but unlikely." The rest of Kriegs et al.'s [59] conclusions are supported by our analysis. The placement of Manis (pangolin) also differs between this hypothesis and Gibson et al. [39] and Reyes et al [41]. Although we show 100% posterior probability for our hypothesis, we also note the exceedingly short branch length of the internode placing Manis as the sister taxon to (Cetartiodactyla(Perissodactyla(Carnivora))). Lewis et al. [60] describe conditions under which Bayesian posterior probabilities may be inflated, and we have not corrected for potentially inflated support for our placement of both Manis and Xenartha. The placement of Xenarthra with Afrotheria and the position of Manis in our phylogenetic hypothesis are congruent with Hudelot et al. [31], who used a 7-state doublet model to accommodate paired RNA sites. Similarities between this study and Hudelot et al [31] could be attributed to the inclusion of RNAs in both studies, while differences are more likely due to differences between models.

Finally, the mitochondrial genome data, even after inclusion of all sequences and a model that incorporates multiple rate classes, reveal several anomalies that are not congruent with recent nuclear gene phylogenies. Some particular anomalies appear to be inherent to all mitogenomic analyses [26, 28, 39, 41], regardless of either taxon sampling or the phylogenetic methods employed. Rather than a monophyletic Primates, as revealed by nuclear genes, our analyses as well as previous mitochondrial phylogenies indicate a paraphyletic Primates with the order Dermoptera (flying lemurs) sister to anthropoid primates (monkeys, lesser and great apes) to the exclusion of the other primate lineages such as tarsiers and prosimians (lemurs). Monophyly of the insectivore group Eulipotyphla, containing the families Erinaceidae, Soricidae, and Talpidae, is supported by nuclear gene phylogenies [32–34, 61] but not by mitochondrial data, which in our case indicates eulipotyphlan diphyly with the Erinaceidae (hedgehogs) at the base of the Laurasiatheria clade. The order Scandentia (tree shrews) is generally considered sister to either Dermoptera or Primates based on recent molecular and morphological data [10, 33, 34, 50], whereas mitogenomic analyses place scandentians at the base of Euarchontoglires. Additionally, mitochondrial data support a monophyletic Tethytheria (elephants and manatees), whereas the more recent nuclear studies [34] do not, and although recent molecular data [62] place marsupial moles (Notoryctes) as part of a monophyletic group (Australidelphia) confined to Australia, our analysis places them basal to other lineages of Metatheria.

Persistent incongruence between mitochondrial and nuclear gene phylogenies relative to the placement of some mammalian lineages may have more than one explanation. Long-branch attraction is often used as an explanation for misplacement of taxa [63, 64], and many of the ambiguous placements involve lineages with longer branches (Fig. 1). As indicated by Bergsten [63], outgroups can often influence placement of ingroup taxa, which may be the case for the position of the marsupial mole. Increased taxon sampling and the incorporation of maximum likelihood models for mitogenomic analyses [63] did remove the Erinaceidae from a basal position in the placental phylogeny to one associated with the Laurasiatheria. Nevertheless, these modifications do not result in a monophyletic Eulipotyphla, as suggested by nuclear genes. In the case of the placement of Dermoptera, there is no apparent reason to consider this as the result of either long branches or branch support from character partitions in the higher rate classes. Schmitz et al. [28] suggested an association between demopteran and anthropoid primate mitochondrial sequences being the result of similarities in nucleotide and amino acid composition. However, Hudelot et al. [31] recovered a monophyletic primates with their doublet model, with the flying lemur as its sister taxon, despite similarities in nucleotide composition at third positions between the flying lemur and Anthropoidea. Finally, if these areas of incongruence are the result of similarities in base composition, covariotide/covarion effects, or some other source of heterogeneity [64], it may very well be that no existing model adequately corrects for all anomalies observed for the mammalian mitochondrial genome.

Conclusion

Although some incongruence still remains between phylogenies derived from mitochondrial and nuclear sequences, our results indicate that the exclusion of data is not necessary for an effective reconstruction of eutherian relationships (although we still excluded the control region and unalignable RNA sites). Rather, selection of an appropriate model that accommodates rate heterogeneity across data partitions and proper treatment of RNA genes can yield information highly congruent with more extensive nuclear sequences, even when addressing the deepest nodes of the eutherian phylogeny. And while we are using "expected" clades to support our conclusions, we note that we are not using phylogenetic expectations as a rationale to exclude data, as is often the case, but rather to retain data. Arguments to retain data should be met with a lower burden of proof than arguments to exclude data.

Methods

Mitochondrial genomes were downloaded from GenBank. A Nexus file was constructed, with each block in the file corresponding to either one gene or a block of data between 100–150 nucleotides for manually aligned rRNAs (the number of nucleotides that are visible on one computer screen without scrolling). Nucleotides between genes were manually aligned, and unaligned regions were placed between brackets (which eliminates them from the dataset, while retaining them for visual inspection). Ribosomal RNAs and tRNAs were aligned manually with reference to secondary structure, according to recommendations of Kjer [65] and Gutell et al. [66]. Models for rRNA secondary structure came from the Comparative RNA Web (CRW) Site [67]. The control region was eliminated. All other genes and codon positions were included. Genes coded in the reverse strand were reversed and complemented.

A site specific rate model was constructed according to Kjer et al. [48]. Briefly, a fast heuristic bootstrap analysis, with 1000 replicates, was completed in PAUP, having saved one tree per replicate. The characters were then separated into 6 discrete rate classes by first selecting the "reweight characters" option in PAUP, according to the "best" CI from among the 1000 bootstrap-generated trees. By selecting "view character weights," and editing the resultant output, we constructed a file in Microsoft Excel that was sorted according to the weights, and then re-imported into the Nexus file to construct 6 partitions or "charsets" from fastest to slowest. These charsets were then used in a partitioned Bayesian analysis, with each partition free to vary according to its own GTR + gamma model.

Each Bayesian analysis was performed with 3 hot and one cold chain. Burnin periods were graphically visualized from the .p files from MrBayes and viewed in Excel. The first set of two independent Bayesian analyses was run for 7.5 million generations in MrBayes 3.0 [68]. Since the likelihood scores from these two chains were not the same, another pair of analyses was conducted in MrBayes 3.1 [68]. This analysis was terminated with a power-failure after 5 million replicates. However, these runs had stabilized on the same likelihood plateau, which was the same as the better of two earlier runs of 7.5 million. Therefore, after discarding the burnin, trees from all three optimal analyses were pooled into a single tree file, from which a majority rule consensus was used to visualize posterior probability values. The best tree was visualized with Treeview [69], and the likelihood phylogram was exported as a pict file for modification.

References

  1. Osborn HF: The Age of Mammals in Europe, Asia and North America. 1910, New York: MacMillan

    Book  Google Scholar 

  2. Simpson GG: The principles of classification and a classification of mammals. Amer Mus Nat Hist Bull. 1945, 85: 1-350.

    Google Scholar 

  3. Wilson DE, Reeder DM: Mammal Species of the World: A Taxonomic and Geographic Reference. 1993, Washington DC: Smithsonian Institution Press

    Google Scholar 

  4. Simpson GG: Tempo and Mode in Evolution. 1944, New York: Columbia University Press

    Google Scholar 

  5. Li W-H, Ellsworth DL, Krushkal J, Chang BH-J, Hewett-Emmett D: Rates of nucleotide substitution in primates and rodents and the generation-time effect hypothesis. Mol Phylogenet Evol. 1996, 5: 182-187. 10.1006/mpev.1996.0012.

    Article  CAS  PubMed  Google Scholar 

  6. Martin AP, Palumbi SR: Body size, metabolic rate, generation time and the molecular clock. Proc Natl Acad of Sci USA. 1993, 90: 4087-4091. 10.1073/pnas.90.9.4087.

    Article  CAS  Google Scholar 

  7. Springer MS: Molecular clocks and the timing of the placental and marsupial radiations in relation to the Cretaceous-Tertiary boundary. J Mammal Evol. 1997, 4: 285-302. 10.1023/A:1027378615412.

    Article  Google Scholar 

  8. McKenna MC, Bell SK: Classification of Mammals: Above the Species Level. 1997, New York: Columbia University Press

    Google Scholar 

  9. Novacek MJ, Wyss AR, McKenna MC: The major groups of eutherian mammals. The Phylogeny and Classification of the Tetrapods. Edited by: Benton MJ. 1988, Oxford: Clarendon Press, 31-71.

    Google Scholar 

  10. Novacek MJ: Mammal phylogenies: shaking the tree. Nature. 1992, 356: 121-125. 10.1038/356121a0.

    Article  CAS  PubMed  Google Scholar 

  11. de Jong WW: Molecules remodel the mammalian tree. Trends Ecol Evol. 1998, 13: 270-275. 10.1016/S0169-5347(98)01335-4.

    Article  CAS  PubMed  Google Scholar 

  12. Honeycutt RL, Adkins RM: Higher level systematics of eutherian mammals: an assessment of molecular characters and phylogenetic hypotheses. Ann Rev Ecol Syst. 1993, 24: 279-305. 10.1146/annurev.es.24.110193.001431.

    Article  Google Scholar 

  13. Springer MS, Stanhope MJ, Madsen O, de Jong WW: Molecules consolidate the placental mammal tree. Trends Ecol Evol. 2004, 19: 430-438. 10.1016/j.tree.2004.05.006.

    Article  PubMed  Google Scholar 

  14. Luckett W, Hartenberger J-L: Monophyly or polyphyly of the order Rodentia: possible conflict between morphological and molecular interpretations. J Mammal Evol. 1993, 1: 127-147. 10.1007/BF01041591.

    Article  Google Scholar 

  15. Graur D, Hide W, Li W-H: Is the guinea-pig a rodent?. Nature. 1991, 351: 649-652. 10.1038/351649a0.

    Article  CAS  PubMed  Google Scholar 

  16. Graur D, Hide W, Zharkikh AA, Li W-H: The biochemical phylogeny of guinea pigs and gundis and the paraphyly of the order Rodentia. Comp Biochem Physiol B. 1992, 101: 495-498. 10.1016/0305-0491(92)90327-N.

    CAS  PubMed  Google Scholar 

  17. Li W-H, Hide WA, Zharkikh A, Ma D-P, Graur D: The molecular taxonomy and evolution of the guinea pig. J Heredity. 1992, 83: 174-181.

    CAS  Google Scholar 

  18. D'Erchia AM, Gissi C, Pesole G, Saccone C, Arnason U: The guinea-pig is not a rodent. Nature. 1996, 381: 597-600. 10.1038/381597a0.

    Article  PubMed  Google Scholar 

  19. Reyes A, Pesole G, Saccone C: Complete mitochondrial DNA sequence of the fat dormouse, Glis glis: further evidence of rodent paraphyly. Mol Biol Evol. 1998, 15: 499-505.

    Article  CAS  PubMed  Google Scholar 

  20. Reyes A, Gissi C, Pesole G, Catzeflis FM, Saccone C: Where do rodents fit? Evidence from the complete mitochondrial genome of Sciurus vulgaris. Mol Biol Evol. 2000, 17: 979-983.

    Article  CAS  PubMed  Google Scholar 

  21. Novacek MJ: Cranial evidence for rodent affinities. Evolutionary Relationships Among Rodents: A Multidisciplinary Analysis. Edited by: Luckett WP, Hartenberger JL. 1985, New York: Plenum Press, 59-81.

    Chapter  Google Scholar 

  22. Graur D, Duret L, Guoy M: Phylogenetic position of the order Lagomorpha (rabbits, hares and allies). Nature. 1996, 379: 333-335. 10.1038/379333a0.

    Article  CAS  PubMed  Google Scholar 

  23. Mouchaty SK, Gullberg A, Janke A, Arnason U: The phylogenetic position of the Talpidae within Eutheria based on analysis of complete mitochondrial sequences. Mol Biol Evol. 2000, 17: 60-67.

    Article  CAS  PubMed  Google Scholar 

  24. Arnason U, Gullberg A, Janke A: Phylogenetic analyses of mitochondrial DNA suggest a sister group relationship between Xenarthra (Edentata) and ferungulates. Mol Biol Evol. 1997, 14: 762-768.

    Article  CAS  PubMed  Google Scholar 

  25. Janke A, Xu X, Arnason U: The complete mitochondrial genome of the wallaroo (Macropus robustus) and the phylogenetic relationship among Montremata, Marsupialia, and Eutheria. Proc Natl Acad Sci USA. 1997, 94: 1276-1281. 10.1073/pnas.94.4.1276.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  26. Arnason U, Adegoke JA, Bodin K, Born EW, Esa YB, Gullberg A, Nilsson M, Short RV, Xu X, Janke A: Mammalian mitogenomic relationships and the root of the eutherian tree. Proc Natl Acad Sci USA. 2002, 99: 8151-8156. 10.1073/pnas.102164299.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  27. Allard MW, Honeycutt RL, Novacek MJ: Advances in higher level mammalian relationships. Cladistics. 1999, 15: 213-219. 10.1111/j.1096-0031.1999.tb00264.x.

    Article  Google Scholar 

  28. Schmitz J, Ohme M, Suryobroto B, Zischler H: The colugo (Cynocephalus variegates, Dermoptera): the primates'gliding sister?. Mol Biol Evol. 2002, 19: 2308-2312.

    Article  CAS  PubMed  Google Scholar 

  29. Schmitz J, Ohme M, Zischler H: The complete mitochondrial sequence of Tarsius bancanus: evidence for an extensive nucleotide compositional plasticity of primate mitochondrial DNA. Mol Biol Evol. 2002, 19: 544-553.

    Article  CAS  PubMed  Google Scholar 

  30. Douzery EJP, Huchon D: Rabbits, if anything, are likely Glires. Mol Phylogenet Evol. 2004, 33: 922-935. 10.1016/j.ympev.2004.07.014.

    Article  CAS  PubMed  Google Scholar 

  31. Hudelot C, Gowri-Shankar V, Jow H, Rattray M, Higgs PG: RNA-based phylogenetic methods: application to mammalian RNA sequences. Mol Phylogenet Evol. 2003, 28: 241-252. 10.1016/S1055-7903(03)00061-7.

    Article  CAS  PubMed  Google Scholar 

  32. Madsen O, Scally M, Douady CJ, Kao DJ, DeBry RW, Adkins R, Amrine HM, Stanhope MJ, de Jong WW, Springer MS: Parallel adaptive radiations in two major clades of placental mammals. Nature. 2001, 409: 610-614. 10.1038/35054544.

    Article  CAS  PubMed  Google Scholar 

  33. Murphy WJ, Eizirik E, Johnson WE, Zhang YP, Ryder OA, O'Brien SJ: Molecular phylogenetics and the origins of placental mammals. Nature. 2001, 409: 614-618. 10.1038/35054550.

    Article  CAS  PubMed  Google Scholar 

  34. Murphy WJ, Eizirik E, O'Brien SJ, Madsen O, Scally M, Douady CJ, Teeling E, Ryder OA, Stanhope MJ, de Jong WW, Springer MS: Resolution of the early placental mammal radiation using Bayesian phylogenetics. Science. 2001, 294: 2348-2351. 10.1126/science.1067179.

    Article  CAS  PubMed  Google Scholar 

  35. Springer MS, Cleven GC, O Madsen O, de Jong WW, Waddell VG, Amrine HM, Stanhope MJ: Endemic African mammals shake the phylogenetic tree. Nature. 1997, 388: 61-64. 10.1038/40386.

    Article  CAS  PubMed  Google Scholar 

  36. Springer MS, DeBry RW, Douady C, Amrine HM, Madsen O, de Jong WW, Stanhope MJ: Mitochondrial versus nuclear gene sequences in deep-level mammalian phylogeny reconstruction. Mol Biol Evol. 2001, 18: 132-143.

    Article  CAS  PubMed  Google Scholar 

  37. Stanhope MJ, Waddell VG, Madsen O, de Jong WW, Hedges SB: Molecular evidence for multiple origins of Insectivora and for a new order of endemic African insectivore mammals. Proc Natl Acad Sci USA. 1998, 95: 9967-9972. 10.1073/pnas.95.17.9967.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  38. Waddell PJ, Shelley S: Evaluating placental inter-ordinal phylogenies with novel sequences including RAG1,?-fibrinogen, ND6, and mt-tRNA, pluse MCMC-driven nucleotide, amino acid, and codon models. Mol Phylogenet Evol. 2003, 28: 197-224. 10.1016/S1055-7903(03)00115-5.

    Article  CAS  PubMed  Google Scholar 

  39. Gibson A, Gowri-Shankar V, Higgs PG, Rattray M: A comprehensive analysis of mammalian mitochondrial genome base composition and improved phylogenetic methods. Mol Biol Evol. 2005, 22: 251-264. 10.1093/molbev/msi012.

    Article  CAS  PubMed  Google Scholar 

  40. Penny D, Hasegawa M: The platypus put in its place. Nature. 1997, 387: 549-550. 10.1038/42352.

    Article  CAS  PubMed  Google Scholar 

  41. Reyes A, Gissi C, Catzeflis F, Nevo E, Pesole G, Saccone C: Congruent mammalian trees from mitochondrial and nuclear genes using Bayesian methods. Mol Biol Evol. 2004, 21: 397-403. 10.1093/molbev/msh033.

    Article  CAS  PubMed  Google Scholar 

  42. Sullivan J, Swofford DL: Are guinea pigs rodents? The importance of adequate models in molecular phylogenetics. J Mammal Evol. 1997, 4: 77-86. 10.1023/A:1027314112438.

    Article  Google Scholar 

  43. Phillips MJ, Penny D: The root of the mammalian tree inferred from whole mitochondrial genomes. Mol Phylogenet Evol. 2003, 28: 171-185. 10.1016/S1055-7903(03)00057-5.

    Article  CAS  PubMed  Google Scholar 

  44. Delsuc F, Scally M, Madsen O, Stanhope MJ, de Jong WW, Catzeflis FM, Springer MS, Douzery EJP: Molecular phylogeny of living xenarthrans and the impact of character and taxon sampling on the placental tree rooting. Mol Biol Evol. 2002, 19: 1656-1671.

    Article  CAS  PubMed  Google Scholar 

  45. Halanych KM: Lagomorphs misplaced by more characters and fewer taxa. Syst Biol. 1998, 47: 138-146. 10.1080/106351598261085.

    Article  CAS  PubMed  Google Scholar 

  46. Lin YH, McLenachan PA, Gore AR, Phillips MJ, Ota R, Hendy MD, Penny D: Four new mitochondrial genomes and the increased stability of evolutionary trees of mammals from improved taxon sampling. Mol Biol Evol. 2002, 19: 2060-2070.

    Article  CAS  PubMed  Google Scholar 

  47. Lin YH, Waddell P, Penny D: Pika and vole mitochondrial genomes increase support for both rodent monophyly and Glires. Gene. 2002, 294: 119-129. 10.1016/S0378-1119(02)00695-9.

    Article  CAS  PubMed  Google Scholar 

  48. Kjer KM, Blahnik RJ, Holzenthal RW: Phylogeny of Trichoptera (Caddisflies): characterization of signal and noise within multiple datasets. Syst Biol. 2001, 50: 781-816. 10.1080/106351501753462812.

    Article  CAS  PubMed  Google Scholar 

  49. Phylogenetic Datasets. [http://www.rci.rutgers.edu/~insects/pdata.htm]

  50. Springer MS, Stanhope MJ, Madsen O, de Jong WW: Molecules consolidate the placental mammal tree. Trends Ecol Evol. 2005, 19: 430-438. 10.1016/j.tree.2004.05.006.

    Article  Google Scholar 

  51. Asher RJ, Novacek MJ, Geisler JH: Relationships of endemic African mammals and their fossil relatives based on morphological and molecular evidence. J Mammal Evol. 2003, 10: 131-194. 10.1023/A:1025504124129.

    Article  Google Scholar 

  52. Ursing BM, Arnason U: Analyses of mitochondrial genomes strongly support a hippopotamus-whale clade. Proc R Soc London [Biol]. 1998, 265: 2251-2255. 10.1098/rspb.1998.0567.

    Article  CAS  Google Scholar 

  53. Geisler JH, Uhen MD: Morphological support for a close relationship between hippos and whales. J Vertebrate Paleont. 2003, 23: 991-996. 10.1671/32.

    Article  Google Scholar 

  54. O'Leary MA, Geisler JH: The position of Cetacea within Mammalia: phylogenetic analysis of morphological data from extinct and extant taxa. Syst Biol. 1999, 48: 455-490. 10.1080/106351599260102.

    Article  PubMed  Google Scholar 

  55. Theodor JM: Molecular clock divergence estimates and the fossil record of Cetartiodactyla. J Paleont. 2004, 78: 39-44. 10.1666/0022-3360(2004)078<0039:MCDEAT>2.0.CO;2.

    Article  Google Scholar 

  56. Springer MS, Teeling EC, Madsen O, Stanhope MJ, de Jong WW: Integrated fossil and molecular data reconstruct bat echolocation. Proc Natl Acad Sci USA. 2001, 98: 6241-6246. 10.1073/pnas.111551998.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  57. Teeling EC, Madsen O, van den Bussche RA, de Jong WW, Stanhope M: Microbat paraphyly and the convergent evolution of a key innovation in Old World rhinolophoid microbats. Proc Natl Acad Sci USA. 2002, 99: 1431-1436. 10.1073/pnas.022477199.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  58. Teeling EC, Springer MS, Madsen O, Bates P, O'Brien SJ: A molecular phylogeny for bats illuminates biogeography and the fossil record. Science. 2005, 307: 580-584. 10.1126/science.1105113.

    Article  CAS  PubMed  Google Scholar 

  59. Kriegs JO, Churakov G, Kiefmann M, Jordan U, Brosius J, Schmitz J: Retrotransposed elements as archives for the evolutionary history of placental mammals. PloS Biology. 2006, 4: e91-10.1371/journal.pbio.0040091.

    Article  PubMed Central  PubMed  Google Scholar 

  60. Lewis PO, Holder MT, Holsinger KE: Polytomies and Bayesian phylogenetic inference. Syst Biol. 2005, 54: 241-53. 10.1080/10635150590924208.

    Article  PubMed  Google Scholar 

  61. Amerine-Madsen H, Koepfli K-P, Wayne RK, Springer MS: A new phylogenetic marker, apolipoprotein B, provides compelling evidence for eutherian relationships. Mol Phylogenet Evol. 2003, 28: 225-240. 10.1016/S1055-7903(03)00118-0.

    Article  Google Scholar 

  62. Amrine-Madsen H, Scally M, Westerman M, Stanhope MJ, Krajewski C, Springer MS: Nuclear gene sequences provide evidence for the monophyly of australidelphian marsupials. Mol Phylogenet Evol. 2003, 28: 186-196. 10.1016/S1055-7903(03)00122-2.

    Article  CAS  PubMed  Google Scholar 

  63. Bergsten J: A review of long-branch attraction. Cladistics. 2005, 21: 163-193. 10.1111/j.1096-0031.2005.00059.x.

    Article  Google Scholar 

  64. Sanderson MJ, Shaffer HB: Troubleshooting molecular phylogenetic analyses. Ann Rev Ecol Syst. 2002, 33: 49-72. 10.1146/annurev.ecolsys.33.010802.150509.

    Article  Google Scholar 

  65. Kjer KM: Use of rRNA secondary structure in phylogenetic studies to identify homologous positions: an example of alignment and data presentation from the frogs. Mol Phylogenet Evol. 1995, 4: 314-330. 10.1006/mpev.1995.1028.

    Article  CAS  PubMed  Google Scholar 

  66. Gutell RR, Larsen N, Woese CR: Lessons from an evolving rRNA: 16S and 23S rRNA structures from a comparative perspective. Microbiol Rev. 1994, 58: 10-26.

    PubMed Central  CAS  PubMed  Google Scholar 

  67. Cannone JJ, Subramanian S, Schnare MN, Collett JR, D'Souza LM, Du Y, Feng B, Lin N, Madabusi LV, Müller KM, Pande N, Shang Z, Yu N, Gutell RR: The comparative RNA web (CRW) site: an online database of comparative sequence and structure information for ribosomal, intron, and other RNAs. BMC Bioinformatics. 2002, 3: 2-10.1186/1471-2105-3-2. [Correction:BMC Bioinformatics 2002, 3:15.]

    Article  PubMed Central  PubMed  Google Scholar 

  68. Ronquist F, Huelsenbeck JP: MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics. 2003, 19: 1572-1574. 10.1093/bioinformatics/btg180.

    Article  CAS  PubMed  Google Scholar 

  69. Page RD: TreeView: an application to display phylogenetic trees on personal computers. Comp Appl BioScience. 1996, 12: 357-358.

    CAS  Google Scholar 

Download references

Acknowledgements

We thank Kenneth (Tripp) MacDonald, William J. Murphy, and two anonymous reviewers for helpful comments on the manuscript. KMK acknowledges financial support from NSF DEB 0423834 and the New Jersey Agricultural Experiment Station, and RLH thanks Pepperdine University for defraying costs of publication.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Rodney L Honeycutt.

Additional information

Authors' contributions

KMK collected genome sequences from GenBank, aligned sequences, and performed initial analyses. RLH provided a detailed comparison of the new phylogeny to previous phylogenetic hypotheses for mammalian relationships and interpreted results relative to ideas concerning the evolution of mammals.

Authors’ original submitted files for images

Below are the links to the authors’ original submitted files for images.

Authors’ original file for figure 1

Authors’ original file for figure 2

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

Kjer, K.M., Honeycutt, R.L. Site specific rates of mitochondrial genomes and the phylogeny of eutheria. BMC Evol Biol 7, 8 (2007). https://doi.org/10.1186/1471-2148-7-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1471-2148-7-8

Keywords