Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Jun 1.
Published in final edited form as: Mol Phylogenet Evol. 2014 Feb 28;75:165–183. doi: 10.1016/j.ympev.2014.02.023

Primate phylogenetic relationships and divergence dates inferred from complete mitochondrial genomes

Luca Pozzi a,b,c,, Jason A Hodgson a,b,d, Andrew S Burrell a, Kirstin N Sterner e, Ryan L Raaum b,f, Todd R Disotell a,b
PMCID: PMC4059600  NIHMSID: NIHMS571603  PMID: 24583291

Abstract

The origins and the divergence times of the most basal lineages within primates have been difficult to resolve mainly due to the incomplete sampling of early fossil taxa. The main source of contention is related to the discordance between molecular and fossil estimates: while there are no crown primate fossils older than 56 Ma, most molecule-based estimates extend the origins of crown primates into the Cretaceous. Here we present a comprehensive mitogenomic study of primates. We assembled 87 mammalian mitochondrial genomes, including 62 primate species representing all the families of the order. We newly sequenced eleven mitochondrial genomes, including eight Old World monkeys and three strepsirrhines. Phylogenetic analyses support a strong topology, confirming the monophyly for all the major primate clades. In contrast to previous mitogenomic studies, the positions of tarsiers and colugos relative to strepsirrhines and anthropoids are well resolved. In order to improve our understanding of how fossil calibrations affect age estimates within primates, we explore the effect of seventeen fossil calibrations across primates and other mammalian groups and we select a subset of calibrations to date our mitogenomic tree. The divergence date estimates of the Strepsirrhine/Haplorhine split support an origin of crown primates in the Late Cretaceous, at around 74 Ma. This result supports a short fuse model of primate origins, whereby relatively little time passed between the origin of the order and the diversification of its major clades. It also suggests that the early primate fossil record is likely poorly sampled.

Keywords: Mitogenomics, Molecular clock, Divergence date, Fossil calibration, Strepsirrhini, Haplorhini

1. Introduction

The order Primates, including 420 species and 73 genera (IUCN 2013.2 IUCN Red List of Threatened Species), is a major mammalian radiation with a rich diversity of morphological, behavioral, and ecological adaptations. Today the non-human members of this order are mainly restricted to tropical or subtropical regions of Asia, Central and South America, and Africa. The fossil record shows that primates were more widely distributed in the past, including in North America and Europe (Covert, 2002). Despite the great interest of evolutionary biologists in this radiation, numerous issues about their origin and diversification remain unclear.

The traditional view of primate origins is that crown primates or euprimates (i.e., all descendants of the last common ancestor of the living species) originated and diversified soon after the Cretaceous-Paleogene (K-Pg) boundary (explosive model). According to this model, the abrupt extinction of dinosaurs and the radiation of flowering plants (angiosperms) opened a new set of ecological niches that placental mammals (Eutheria) and birds (Neoaves) could exploit (Conroy, 1990; Rose, 2006). As a consequence, those groups of animals that remained fairly small and cryptic during the Cretaceous underwent a rapid adaptive radiation soon after the mass extinction 65 millions year ago (Rose, 2006; Wible et al., 2007; O’Leary et al., 2013).

The main line of evidence supporting the explosive model is the absence of euprimate fossils prior to the K-Pg boundary. The earliest unambiguous fossil euprimates date from the beginning of the Eocene epoch and are no older than 56 Ma (Miller et al. 2005; Smith et al. 2006; Williams et al. 2010). The best-known representatives of these early euprimates are the two groups Omomyoidea (e.g., Teilhardina and Melanerimia) and Adapoidea (e.g., Cantius and Donrussellia), possibly related to the living haplorhines and strepsirrhines, respectively (Gingerich, 1986; Martin, 1993; Ni et al., 2004; Miller et al., 2005; Smith et al., 2006; Rasmussen, 2007). The North African Altiatlasius koulchii (Sigé et al., 1990) is slightly earlier, dated to approximately 57 Ma in the late Paleocene, but is not unambiguously a euprimate (Rasmussen, 2007; Williams et al., 2010). Finally, the plesiadapiforms, which have in the past been included in the order Primates, are found in the fossil record between the latest Cretaceous and the early Eocene, with most specimens recovered from deposits dating to the Paleocene (56–66 Ma). However, in contemporary taxonomies, none of these fossils are included in the euprimates and are thought to represent either a stem-group that diverged prior to the origin of euprimates (Bloch and Silcox, 2001; Bloch and Boyer, 2002; Bloch et al., 2007; Silcox et al., 2007) or a non-primate mammalian group possibly related to primates or colugos (Martin, 1968; Cartmill, 1972; Wible and Covert, 1987; Beard and Houde, 1989; Beard, 1990, 1993; Kay and Thorington, 1990; Kay et al., 1992).

Analyses of molecular data have provided a quite different scenario for the origin and diversification of primates. Molecular studies have yielded divergence dates older than any known euprimate fossil, extending the origins of crown primates before the K-Pg boundary (63.7–95.0 Ma; Springer et al., 2003, 2012; Chatterjee et al., 2009; Fabre et al., 2009; Meredith et al., 2011; Perelman et al., 2011; Jameson et al., 2011; Wilkinson et al., 2011). This view of primate origins agrees with a number of recent molecular studies on mammal diversification suggesting a deep Cretaceous origin for most of the modern placental lineages (Madsen et al., 2001; Murphy et al., 2001; Springer et al., 2003; Bininda-Emonds et al., 2007; Meredith et al., 2011), despite the lack of fossils showing clear morphological traits of the crown group Placentalia (Wible et al., 2007; O’Leary et al., 2013). In contrast to the explosive model, these studies place the origin of eutherian orders in the Cretaceous, supporting either a long-fuse model (the origin of the orders extend deep into the Cretaceous, with the origin of the main intraordinal lineages near the K-Pg boundary) or a short-fuse model (both interordinal and intraordinal diversity originating deep in the Cretaceous) (Springer et al., 2003; Bininda-Emonds et al., 2007; Meredith et al., 2011).

While it is expected that molecules would overestimate and fossils underestimate divergence times (Steiper and Young, 2008), the early lineage origin dates inferred from molecular studies suggest that we are missing a substantial proportion of primate evolutionary history (~20–25 million years). A possible source of the discrepancy between paleontological and molecular models of primate origins lies in the calibration of molecular clocks. In order to convert molecular evolutionary distance estimates to chronological time, a set of well-dated fossils with well-supported phylogenetic placement is required. These calibration points are critical in molecular dating since they are the only source of information about absolute time (Steiper and Young, 2008; Ho and Phillips, 2009; Parham et al., 2012). Three main problems can affect the use of fossil information in estimating divergence times using molecular sequences: 1) the selection of inappropriate fossils; 2) the incompleteness of the fossil record in some areas of the tree; and 3) the phylogenetic misplacement of a fossil (e.g., placing a stem form in a crown group). Though these problems have long been recognized as critical in dating a molecular tree (Raaum et al. 2005; Rutschmann et al., 2007; Steiper and Young, 2008; Ho and Phillips, 2009), the difficulty of accurately choosing proper calibration points is still underestimated (Pyron, 2010; Dornburg et al., 2011; Parham et al., 2012). Many molecular studies suffer from a non-rigorous use of the paleontological record that may easily include considerable error in fossil calibration and provide misleading divergence estimates (Graur and Martin, 2004).

Here, we present one of the largest time-calibrated phylogenies of primates estimated using complete mitochondrial genome sequences. Mitogenomic analyses have been shown to be useful for phylogenetic reconstruction and divergence time estimation at different taxonomic levels and in different groups, including amphibians (Zardoya et al., 2003; Zhang et al., 2005; Zhang and Wake, 2009), birds (Pereira and Baker, 2006; Slack et al., 2007; Pacheco et al., 2011), fish (Inoue et al., 2003, 2010; Miya et al., 2003), and mammals (Arnason and Janke, 2002; Arnason et al., 2002, 2004, 2008). Mitochondrial genomes have traditionally been used for phylogenetic and phylogeographic analyses of animal taxa. The use of mitochondrial DNA (mtDNA) has several advantages including: the lack of recombination, a rapid coalescence time, relatively high substitution rates, high copy number, and haploidy. For these reasons, it has been suggested that mitochondrial gene trees are more likely to approximate the history of divergence among species than other loci (Moore, 1995). Today, multilocus phylogenies are very common (Jameson et al., 2011; Perelman et al., 2011; Springer et al., 2012), but the nuclear loci used in these studies are usually quite short (<1500 bp) compared to the large size of mitochondrial genomes (>16,000 bp). Mitochondrial genomes also show higher information content per base than nuclear DNA (Anderson et al., 1982; Cummings et al., 1995). Finally, the high number of copies of mitochondria in each cell makes mitogenomic data easy to obtain and sequence, especially in low quality samples such as museum and fossil specimens (Briggs et al., 2009; Krause et al., 2010; Rowe et al., 2011; Mason et al., 2011; Guschanski et al., 2013). Therefore, a good knowledge of the mitochondrial evolutionary history is not only valuable per se, but it can also provide an important comparative dataset in studies for which mtDNA is the only genetic marker available.

Within primates, mitochondrial genomes have been commonly used in phylogenetic studies at the family or genus level (Raaum et al., 2005; Sterner et al., 2006; Hodgson et al., 2009; Chan et al., 2010; Chiou et al., 2011; Guschanski et al., 2013; Zinner et al., 2013), but to date only two comprehensive mitogenomic studies including all main primate lineages have been performed (Matsui et al., 2009; Finstermeier et al., 2013). Here we assembled whole mitochondrial genome sequences from sixty-two primate species (including eleven new sequences) representing all sixteen living families within the order Primates. The goals of this study are (i) to infer the evolutionary relationships of primates using whole mitochondrial genome sequences, (ii) to investigate how different fossil calibrations selected within and outside primates affect estimates of times of divergence within the order, and (iii) to develop a suite of congruent fossil calibrations for use in future studies.

2. Material and methods

2.1 Taxon sampling

We sequenced complete mitochondrial genomes for eleven species representing three strepsirrhines (Daubentonia madagascarensis, Mirza coquereli, and Otolemur crassicaudatus) and eight Old World monkeys (Allenopithecus nigroviridis, Cercopithecus diana, C. lhoesti, C. mitis, Erythrocebus patas, Cercocebus torquatus, Lophocebus aterrimus, and Mandrillus sphynx) (Table 1). Complete genome sequences from these specimens are deposited in GenBank (http://www.ncbi.nlm.nih.gov/) under accession numbers KJ158462-KJ158463 and KJ434955-KJ434963 (Table 1).

Table 1.

Sequences used in this study.

Scientific name Common name NCBI Accession No Reference
Primates
 Strepsirrhini
  Lorisiformes
   Lorisidae
    Loris tardigradus Slender loris NC_012763 Matsui et al., 2009
    Nycticebus coucang Slow loris NC_002765 Arnason et al., 2000
    Perodicticus potto Potto NC_012764 Matsui et al., 2009
   Galagidae
    Galago senegalensis Senegal galago NC_012761 Matsui et al., 2009
    Otolemur crassicaudatusa Thick-tailed galago KJ434961 This study
  Chiromyiformes
   Daubentoniidae
    Daubentonia madagascariensisa Aye-aye KJ158462 This study
  Lemuriformes
   Lemuridae
    Eulemur mongoz Mongoose lemur NC_010300 Arnason et al., 2008
    Eulemur fulvus fulvus Brown lemur NC_012766 Matsui et al., 2009
    Eulemur fulvus mayottensis Brown lemur NC_012769 Matsui et al., 2009
    Eulemur macaco macaco Black lemur NC_012771 Matsui et al., 2009
    Lemur catta Ring-tailed lemur NC_004025 Arnason et al., 2002
    Varecia variegata Black and white ruffled lemur NC_012773 Matsui et al., 2009
   Cheirogaleidae
    Mirza coquerelia Coquereli’s giant mouse lemur KJ158463 This study
   Lepilemuridae
    Lepilemur hubbardorum Hubbard’s sportive lemur NC_014453 Lei at al., 2010
   Indridae
    Propithecus coquereli Coquerel’s sifaka AB286049 Matsui et al., 2007
 Haplorrhini
  Tarsiformes
   Tarsiidae
    Tarsius bancanus Horsfield’s tarsier NC_002811 Schmitz et al., 2002b
    Tarsius syrichta Philippine tarsier NC_012774 Matsui et al., 2009
 Anthropoidea
  Platyrrhini
   Cebidae
    Aotus lemurinus Lemurine night monkey FJ785421 Hodgson et al., 2009
    Cebus albifrons White-fronted capuchin AJ309866 Arnason et al., 2000
    Saimiri sciureus Common squirrel monkey FJ785425 Hodgson et al., 2009
    Saguinus oedipus Cotton-top tamarin FJ785424 Hodgson et al., 2009
   Pithecidae
    Callicebus donacophilus Bolivian titi FJ785423 Hodgson et al., 2009
   Atelidae
    Ateles belzebuth White-bellied spider monkey FJ785422 Hodgson et al., 2009
 Catarrhini
  Cercopithecoidea
   Colobinae
    Colobus guereza Guereza AY863427 Raaum et al., 2005
    Nasalis larvatus Proboscis monkey DQ355298 Sterner et al., 2006
    Pilocolobus badius Western red colobus DQ355301 Sterner et al., 2006
    Presbytis melalophos Mitred leaf monkey DQ355299 Sterner et al., 2006
    Pygathrix nemaeus Douc langur DQ355302 Sterner et al., 2006
    Rhinopithecus roxellana Golden snub-nosed monkey DQ355300 Sterner et al., 2006
    Semnopithecus entellus Hanuman langur DQ355297 Sterner et al., 2006
    Trachypithecus obscurus Dusky leaf monkey NC_006900 Raaum et al., 2005
   Cercopithecinae
    Allenopithecus nigroviridisa Allen’s swamp monkey KJ434962 This study
    Cercocebus torquatusa Collared mangabey KJ434959 This study
    Cercopithecus dianaa Diana monkey KJ434958 This study
    Cercopithecus lhoestia L’Hoest’s monkey KJ434957 This study
    Cercopithecus mitisa Blue monkey KJ434956 This study
    Chlorocebus aethiops Grivet NC_007009 Raaum et al., 2005
    Chlorocebus pygerythrus Vervet monkey NC_009747 Wertheim & Worobey, 2007
    Chlorocebus sabaeus Green monkey EF597503 Wertheim & Worobey, 2007
    Chlorocebus tantalus Tantalus monkey NC_009748 Wertheim & Worobey, 2007
    Erythrocebus patasa Patas monkey KJ434955 This study
    Lophocebus aterrimusa Black crested mangabey KJ434960 This study
    Macaca sylvanus Barbary ape NC 002764 Arnason et al., 2000
    Macaca mulatta Rhesus monkey NC_005943 Gokey et al., 2004
    Macaca fascicularis Crab-eating macaque NC_012670 Yi et al., unpub.
    Macaca thibetana Pere David’s macaque NC_011519 Li et al., 2009
    Mandrillus sphinxa Mandrill KJ434963 This study
    Papio hamadryas Hamadryas baboon NC_001992 Arnason et al., 1998
    Theropithecus gelada Gelada FJ785426 Hodgson et al., 2009
  Hominoidea
    Gorilla gorilla Western gorilla NC_001645 Horai et al., 1995
    Homo neanderthalensis Neanderthal AM948965 Green et al., 2008
    Homo sapiens Human EF061150 Friedlaender et al., 2007
    Homo sp. (Denisovan) Denisova hominin FR695060 Krause et al., 2010
    Hylobates agilis Agile gibbon NC_014042 Matsudaira & Ishida, 2010
    Hylobates lar White-handed gibbon NC_002082 Arnason et al., 1996
    Hylobates pileatus Pileated gibbon NC_014045 Matsudaira & Ishida, 2010
    Nomascus siki Southern white-cheeked gibbon NC_014051 Matsudaira & Ishida, 2010
    Pan paniscus Pygmy chimpanzee D38116 Horai et al., 1995
    Pan troglodytes Common chimpanzee EU095335 Flynn & Carr, 2007
    Pongo abelii Sumatran orangutan X97707 Xu & Arnason, 1996
    Pongo pygmaeus Bornean orangutan NC_001646 Horai et al., 1995
    Symphalangus syndactylus Siamang NC_014047 Matsudaira & Ishida, 2010
Non-primate mammals
Euarchontoglires
 Dermoptera
  Galeopterus variegatus Sunda flying lemur NC_004031 Arnason et al., 2002
  Galeopterus variegatus Sunda flying lemur AF460846 Schmitz et al., 2002a
 Rodentia
  Cavia porcellus Guinea pig AJ222767 D’Erchia et al., 1996
  Cricetulus griseus Chinese hamster NC_007936 Partridge et al., 2007
  Glis glis Fat dormouse NC_001892 Reyes et al., 1998
  Mus musculus Mouse NC_005089 Bayona-Bafaluy et al., 2003
  Rattus norvegicus Brown rat AJ428514 Nilsson et al., 2003
  Sciurus vulgaris Squirrel AJ238588 Reyes et al., 2000
 Lagomorpha
  Lepus europaeus European hare AJ421471 Arnason et al., 2002
  Ochotona collaris Collared pika NC_003033 Lin et al., 2002
  Oryctolagus cuniculus Rabbit NC_001913 Gissi et al., 1998
Laurasiatheria
 Carnivora
  Canis familiaris Dog NC_002008 Kim et al., 1998
  Felis catus Domestic cat U20753 Lopez et al., 1996
 Perissodactyla
  Equus caballus Horse NC_001640 Xu & Arnason, 1994
  Rhinoceros unicornis Greater Indian rhinoceros NC_001779 Xu et al., 1996
 Cetartiodactyla
  Balaenoptera musculus Blue whale NC_001601 Arnason et al., 1993
  Balaenoptera physalus Finback whale X61145 Arnason et al., 1991
  Bos taurus Cow NC_006853 Chung,and Ha, unpub.
  Hippopotamus amphibius Hippopotamus AJ010957 Ursing & Arnason, 1998b
  Lama pacos Alpaca NC_002504 Ursing et al., 2000
  Ovis aries Sheep AF010406 Hiendleder et al., 1998
  Physeter catodon Sperm whale NC_002503 Arnason et al., 2000
  Sus scrofa Pig AJ002189 Ursing & Arnason, 1998a
Afrotheria
 Proboscidea
  Elephas maximus Asiatic elephant NC_005129 Rogaev et al., 2006
  Loxodonta africana African elephant NC_000934 Hauf et al., 2000
a

in bold new sequences generated for this study

These genomes were added to a dataset including mitochondrial genomes of 51 other primates, in total representing all major taxonomic lineages within the order, as well as mitogenomes of 23 other species representing major lineages within Boreoeutheria. Rooting was done using two Afrotheria species, Elephas maximum and Loxodonta africanus, commonly accepted as external to Boreoeutheria in mammal phylogenies (Murphy et al., 2001; Arnason and Janke, 2002; Springer et al., 2003; Bininda-Emonds et al., 2007; Arnason et al., 2008; Meredith et al., 2011; Song et al., 2012). Overall 87 mammalian genomes were included in the phylogenetic analyses.

2.2 Sequencing

DNA was extracted and isolated from either blood or tissue samples using the QIAamp DNA Micro Kit (Qiagen, Inc.) following the protocol provided by the manufacturer. List of the samples with location, sample type, and GenBank accession number is reported in Supplementary Table A1. All the samples are currently stored at the Molecular Anthropology lab at New York University. Mitochondrial genomes were sequenced by long-template PCR, which minimizes the chance of amplifying mitochondrial pseudogenes in the nuclear genome (numts) (Thalmann et al., 2004; Raaum et al., 2005). We employed the same approach previously used in Raaum et al. (2005), Sterner et al. (2006), Hodgson et al. (2009), and Chiou et al. (2011). Briefly, independent sets of amplification primers that amplify the entire mtDNA genome in two or more overlapping long fragments were designed for each species. PCR reactions were performed employing a long-range PCR kit (Expand Long Template PCR System, Roche). The sequencing products were analyzed on a 3730 DNA Analysis System (Applied Biosystems, Foster City, CA). Sequences were then assembled and edited using Sequencher 4.1 (Gene Codes, Corp., Ann Arbor, MI). In addition to the eleven newly sequenced genomes, fifteen of the primate genomes used in this study were previously sequenced in the NYU Molecular Anthropology lab applying the technique described above (see Raaum et al. 2005; Sterner et al. 2006; Hodgson et al. 2009). All the remaining genomes (61) obtained from GenBank were analyzed to minimize the presence of nuclear copies of mitochondrial genes (numts) in the dataset. First, we translated all of the protein-coding mitochondrial genes into amino acids to check for the presence of premature stop codons and frame shifting insertions or deletions. Then, for each protein-coding gene and RNA sequence independently, we constructed a neighbor-joining tree to identify any sequence that displayed unusual phylogenetic placement.

2.3 Alignment

Heavy-strand protein-coding genes and RNA sequences (2 rRNAs and 22 tRNAs) were automatically extracted from complete genome sequences according to GenBank annotations using a Perl script written by JAH for this purpose. The D-loop was excluded from the dataset because of alignment difficulties due to its high variability. Also the ND6 gene was excluded from the final alignment because it is encoded on the mitochondrial L-strand which has a different nucleotide composition from the H-strand, and has been shown to have poor phylogenetic signal (Gissi et al., 2000; Miya and Nishida, 2000). Protein-coding genes were then aligned based on their corresponding amino acid translations using the software TranslatorX (Abascal et al., 2010). Since homology is best identified at the amino acid level due to the evolution of coding DNA as triplets of nucleotides and the redundancy of the genetic code, this software translates the DNA sequences into amino acids, aligns the amino acid sequences, and then back-translates the alignment to the nucleotide sequences. We allowed the software to automatically identify the most likely reading frame (minimizing the number of stop codons) and we used MUSCLE (Edgar, 2004) to perform the protein alignment. RNA sequences (rRNAs and tRNAs) were aligned with reference to their secondary structure. Due to their stem-and-loop structure, the identification of homologous regions is best identified when the secondary structure of the RNA sequences is taken into account. This approach has been advocated to improve not only the alignment itself but also the phylogenetic tree reconstruction (Kjer and Honeycutt, 2007; Stocsits et al., 2009; Letsch et al., 2010). We employed two different structure informed alignment methods for RNA sequences: RNAsalsa (Stocsits et al., 2009) and RAF (Do et al., 2006, 2008). RNAsalsa, specifically designed to align ribosomal RNAs, uses preexisting knowledge of RNA structure to simultaneously predict secondary structure in the RNA sequences and to align them based on the secondary structure information. We used secondary structures of Bos taurus as reported in the literature (Springer and Douzery, 1996; Burk and Douzery, 2002; Stocsits et al., 2009) to align the remaining 86 sequences in our dataset. RAF (RNA Alignment and Folding) uses an algorithm for simultaneous alignment and consensus folding of unaligned RNA sequences. In contrast to RNAsalsa, RAF does not require preexisting information about secondary structure but identifies sets of likely pairing and alignment candidates for each nucleotide to effectively obtain a simultaneous pairwise alignment and folding (Do et al., 2006, 2008) and then compares the results of the two candidate alignments to identify possible discrepancies.

The resulting alignments (12 protein coding genes, 12S and 16S rRNAs, and 22 tRNAs) were concatenated after removing problematic regions using Gblocks 0.91 (Talavera and Castresana, 2007) under a relaxed approach. Gblocks removes all poorly aligned regions in a dataset and it has been shown to be particularly effective in phylogenetic studies including very divergent sequences (Castresana, 2000; Talavera and Castresana, 2007). Gblocks selects blocks to be deleted following a reproducible set of requirements, including the lack of large segments of contiguous non-conserved positions, lack of gap positions and high conservation of flanking positions. Gblocks was run with the options “Minimum Length Of A Block” = 10 (5 for RNAs), and “Allowed Gap Positions” = “With Half”.

A total of 11,022 bp from the 12 protein-coding genes, 1,313 bp from 22 tRNA genes, and 1,708 bp from two rRNA genes were unambiguously aligned. The individual alignments were then concatenated using the software SequenceMatrix v1.7.6 (Vaidya et al., 2011) to create a master alignment of 14,043 bp total, equivalent to approximately 83% of the mitochondrial genome.

2.4 Data partitions

For both Bayesian and maximum likelihood (ML) analyses, we evaluated six different partition schemes of the mitochondrial genome: 1) no partition (MT0); 2) four partitions (protein-coding genes, 12S rRNA, 16rRNAs, and 22 tRNAs) (MT4); 3) five partitions, identical to the previous one but with the protein coding genes separated in two different partitions (i) codon position 1 and 2 and (ii) codon position 3 (MT5); 4) six partitions, consisting of three partitions for the two rRNAs (12S and 16S) and the 22 tRNAs, and three partitions for the three codon positions of the protein-coding genes (MT6); 5) fifteen partitions, where the 12 protein-coding genes were treated as independent partitions (MT15); and 6) 39 partitions, where the different codon positions of each protein-coding gene were treated as separate partition, plus the three RNA partitions (MT39). The nucleotide substitution model for each partition was selected among 24 models using the Akaike Information Criterion (AIC) (Posada and Buckley, 2004) as implemented in MrModeltest v2.3 (Nylander, 2004). For each of the data partitions, we evaluated the variable sites and the parsimony-informative sites across all taxa and within the ingroup (primates) only (Table 2).

Table 2.

Sequence characteristics and models chosen for each data partition

Partition No. sites All taxa Only primates Model

VS PIS VS PIS
All genes 14043 9080 8244 8396 7532 GTR+I+G
12S rRNA 691 349 308 286 244 GTR+I+G
16S rRNA 1017 489 389 419 313 GTR+I+G
all tRNAs 1313 777 643 652 514 GTR+I+G
All protein-coding genes 11022 7465 6904 7039 6461 GTR+I+G
 pos1 3674 2277 2033 2092 1827 GTR+I+G
 pos2 3674 1568 1270 1341 1075 GTR+I+G
 pos1+2 7348 3845 3303 3433 2902 GTR+I+G
 pos3 3674 3620 3601 3606 3559 GTR+I+G
Individual protein-coding genes
ATP6 681 499 465 477 434 GTR+I+G
  pos1 227 163 151 156 134 GTR+I+G
  pos2 227 112 92 98 80 GTR+I+G
  pos3 227 224 222 223 220 GTR+I+G
ATP8 225 198 190 193 176 GTR+G
  pos1 75 64 59 62 55 GTR+G
  pos2 75 64 62 61 54 GTR+G
  pos3 75 70 69 70 67 GTR+G
COI 1569 807 727 752 694 GTR+I+G
  pos1 523 198 153 164 135 GTR+I+G
  pos2 523 96 62 76 50 GTR+I+G
  pos3 523 513 512 512 509 GTR+I+G
COII 696 443 408 417 385 GTR+I+G
  pos1 232 134 121 124 111 GTR+I+G
  pos2 232 81 60 66 49 HKY+I+G
  pos3 232 228 227 227 225 GTR+I+G
COIII 804 463 426 443 406 GTR+I+G
  pos1 268 131 110 122 101 GTR+I+G
  pos2 268 72 58 62 50 GTR+I+G
  pos3 268 260 258 259 255 GTR+I+G
CYTB 1143 747 961 714 648 GTR+I+G
  pos1 381 221 200 209 182 GTR+I+G
  pos2 381 147 114 126 102 HKY+I+G
  pos3 381 379 377 379 364 GTR+I+G
NADH1 963 640 599 601 556 GTR+I+G
  pos1 321 203 180 176 160 GTR+I+G
  pos2 321 118 101 106 81 GTR+I+G
  pos3 321 319 318 319 315 GTR+G
NADH2 1050 848 478 788 729 GTR+I+G
  pos1 350 274 252 258 230 GTR+I+G
  pos2 350 226 192 185 159 GTR+I+G
  pos3 350 348 34 345 340 GTR+G
NADH3 360 248 231 234 219 GTR+I+G
  pos1 120 78 71 69 64 GTR+I+G
  pos2 120 54 45 50 42 GTR+I+G
  pos3 120 116 115 115 113 GTR+G
NADH4 1386 1002 911 951 864 GTR+I+G
  pos1 462 319 287 297 258 GTR+I+G
  pos2 462 225 179 197 157 GTR+I+G
  pos3 462 458 445 457 449 GTR+I+G
NADH4L 297 222 205 204 186 GTR+I+G
  pos1 99 74 68 68 60 GTR+I+G
  pos2 99 51 40 40 30 GTR+G
  pos3 99 97 97 96 96 GTR+I+G
NADH5 1848 1348 1253 1265 1154 GTR+I+G
  pos1 616 418 381 387 337 GTR+I+G
  pos2 616 322 265 274 221 GTR+I+G
  pos3 616 608 607 604 596 GTR+I+G

Note: VS = variable sites; PIS = parsimony-informative sites.

2.5 Phylogenetic Analysis

Maximum likelihood (ML)

ML analyses were run using a Randomized Accelerated Maximum Likelihood in RAxML version 7.2.6 (Stamatakis et al., 2005; Stamatakis, 2006). For each partition scheme, we ran 50 independent ML inferences (using 50 distinct randomized MP trees) with a GTRGAMMA model to estimate the best topology. We then performed a rapid (–f a -x option) with 1000 replications to assess support on different nodes (Stamatakis, 2006; Stamatakis et al., 2008). Maximum-likelihood bootstrap proportions (MLBS) ≥70% were considered strong support (Hillis and Bull, 1993; Wilcox et al., 2002).

Bayesian analyses (MB)

Bayesian phylogenetic analyses were conducted with MrBayes 3.2.1 (Huelsenbeck and Ronquist, 2001; Ronquist and Huelsenbeck, 2003) using the Metropolis coupled Markov Chain Monte Carlo (Metropolis coupled MCMC or MC3) algorithm. The most appropriate model of nucleotide evolution was estimated independently for each partition using the AIC test as implemented in MrModelTest 2.3 (Nylander, 2004) as reported in Table 2. Although the model selected in most partitions was GTR+G+I, we also run our analyses using a simplified model GTR+G. This was done because Bayesian analyses may run into problems of convergence of the MCMC when one tries to accommodate among-site substitution rate variation using a combination of proportion of invariant sites model (+I) and a gamma distributed rates model (+G). This conflict may induce two regions of high posterior probability creating clear problems of convergence in the Bayesian analysis (Brian Moore, pers. comm.; Stamatakis 2006). Four independent runs, each including four incrementally heated chains, were run for at least 20 million generations. Different partition schemes required different number of generations to properly reach convergence as reported in Table 3. The sample frequency was adjusted according to the number of generations (Table 3). We assessed convergence visually using Tracer v.1.5 (Rambaut and Drummond, 2007) to plot the likelihood versus generation number and estimate the effective sample size (ESS>200) of all parameters and to compare the performance of the four independent analyses. In addition, we used AWTY (Are We There Yet?, Nylander et al., 2008) to plot pairwise split frequencies for the four independent MCMC runs and to check the posterior probabilities of clades for non-overlapping samples of trees in the sample using the compare and slide commands, respectively. After checking for convergence we summarized the posterior distribution of trees removing the first 10% of generations as burn-in. Posterior probability (PP) support values higher than 0.95 were considered strong support for individual clades (e.g., Alfaro et al., 2003; Erixon et al., 2003; Huelsenbeck and Rannala, 2004).

Table 3.

Summary of model partition comparison (Maximum Likelihood and Bayesian analyses)

Model Maximum Likelihood Bayesian
Run Generations Burn-in Marginal Likelihood SD
MT0 −421145.742 1 20,000,000 2,000,000 −417245.466 0.239
2 20,000,000 2,000,000 −417245.834 0.287
MT4 −420115.955 1 20,000,000 2,000,000 −415805.630 0.329
2 20,000,000 2,000,000 −415804.224 0.280
MT5 −409906.041 1 20,000,000 2,000,000 −409740.282 0.295
2 20,000,000 2,000,000 −409740.561 0.329
MT6 −408934.350 1 20,000,000 2,000,000 −408453.105 0.300
2 20,000,000 2,000,000 −408453.716 0.297
MT15 −419192.485 1 30,000,000 3,000,000 −414803.673 0.311
2 30,000,000 3,000,000 −414814.001 0.381
MT39 −407068.508 1 40,000,000 4,000,000 −402921.612 0.318
2 40,000,000 4,000,000 −402921.590 0.364

2.6 Partition models comparison

In order to evaluate the relative support for competing partition models, we compared the marginal likelihood scores for both maximum likelihood and Bayesian analyses (Table 3). We also used Bayes factors (Kass and Raftery, 1995) to select the partition model better supported by Bayesian analyses. We calculated ln (Bayes factor) as the difference in estimated marginal likelihoods for pairs of models, as described in (Nylander et al., 2004). For Bayesian analyses, marginal likelihoods were estimated from the harmonic means using Tracer v1.5 (Rambaut and Drummond, 2007) by employing the weighted likelihood bootstrap estimator with standard error estimated using 1000 bootstrap pseudoreplicates (Newton, 1994; Suchard et al., 2001). The results for the partition model comparison reported here are based on combining the post–burn-in data from the two runs that achieved the greatest marginal likelihoods, as summarized in Table 3. The similarity of tree topologies resulting from the analyses using different partitions was assessed using the program TOPD-FMTS version 3.3 (Puigbò et al., 2007). Unrooted maximum likelihood and Bayesian trees (without branch lengths) for each of the six partitioning schemes were compared. The TOPD-FMTS analysis was performed under default conditions, with 100 simulated trees as a null model. Analyses were run using the nodal, the split, and the disagree options. The nodal method compares the number of nodes that separate each taxon from the other taxa in the tree and calculates pairwise distance matrices from two input trees using only those taxa that are common to both trees (Puigbò et al., 2007). The split option uses a crossover method that takes into account the minimum number of operations required to convert one tree into the other, following the algorithm proposed by Robinson and Foulds (1981). Finally, the disagree option compares two trees and returns the taxa that make the disagreement between those trees. It first removes one taxon and calculates the gain (reduction in the split distance) between two trees. The taxon that produces most gain is removed for the following iterations. Then this procedure is repeated until the split distance is zero (Puigbò et al., 2007).

2.7 Test for compositional biases

Research over the last decade has shown that phylogenetic inference can be affected by systematic error, such as non-phylogenetic signals that can drive phylogenetic results (Phillips et al., 2004; Phillips, 2009). Even complex models of nucleotide substitution may not be able to capture the substitution process well enough to reliably provide an accurate phylogenetic reconstruction (Phillips et al., 2004; Phillips, 2009). This problem might be even more relevant in large concatenated datasets in which “long sequences exacerbate the potential for biases (systematic error) to be positively misleading” (Phillips et al., 2004: pp. 1455).

The encoding of nucleotides as purines and pyrimidines (RY-coding) has been suggested as an effective approach to assess the effects of compositional biases and non-phylogenetic signal (multiple substitutions) (Phillips et al., 2004; Phillips, 2009). Previous studies on mitochondrial genomes have shown that RY-coding can improve the level of signal relative to compositional heterogeneity, by giving more weight to the slowest evolving sites in the mitochondrial genomes (Brinkmann and Philippe 1999; Phillips and Penny 2003; Delsuc, Phillips, and Penny 2003; Phillips et al., 2004). This approach can therefore increase the ratio between historical and nonhistorical signal, reducing the weight of systematic biases. In order to control for this issue, we compiled and analyzed three additional datasets: 1) third codon position in protein coding regions coded as RY [3RY], 2) positions with Cs and Ts replaced with Y [AGY], and 3) all nucleotides coded as purines (R) and pyrimidines (Y) [allRY]. For each dataset, relative support of internal nodes was assessed by bootstrap analyses using the bootstopping criteria autoMRE (Majority Rule Criterion) as implemented in RAxML 7.2.6 (Pattengale et al., 2009).

2.8 Taxon sampling effects on phylogenetic reconstruction

Incomplete taxon sampling has been demonstrated to be a major problem in phylogenetic reconstructions (Zwickl and Hillis, 2002; Hillis et al., 2003; Plazzi et al., 2010; Townsend and Leuenberger, 2011; Nabhan and Sarkar, 2012). Here, we explicitly test the importance of taxon representation in resolving the relationships between the major lineages within primates and colugos. We assembled four additional datasets, including different numbers of primate taxa in each: 1) only six primate taxa (one for each main lineage within primates: Lemuriformes, Chyromyiformes, Lorisiformes, Tarsiiformes, Platyrrhini, Catarrhini), 2) ten taxa as used by Schmitz et al. (2002a), 3) 14 taxa as used by Arnason et al. (2008), and 4) 26 taxa as in Matsui et al. (2009). We then compared the results obtained by these four datasets with our original dataset with 62 primate species. Phylogenetic analyses were run using RAxML 7.2.6 as fully described above.

2.9 Divergence date estimation and cross-validation analyses

Fossil constraints were carefully selected based on criteria for choosing appropriate calibration points reviewed elsewhere (Raaum et al., 2005; Ho and Phillips, 2009; Parham et al., 2012). We employed seventeen fossil constraints identified from the literature and previous studies performed on primate and mammal phylogeny. Details about these fossil calibrations, including age, phylogenetic position, minimum and (putative) maximum bounds, and references are reported in Table 4. The detailed procedure of fossil selection, with paleontological details, and assignment to alternative nodes for calibration of the molecular clock is fully reported in Supplementary Material B1.

Table 4.

Evolutionary rate calibration constraints (in millions of years).

Divergence Min. Max. Fossil Reference Age
Primate fossils
1. Homo 5.0 10.0 Ardipithecus Haile-Selassie et al., 2001 5.2
Orrorin Senut et al., 2001 6.0
Sahelanthropus Vignaud et al., 2002; Brunet et al., 2002 6.0–7.0
2. Gorilla 10.0 NA Chororapithecus abyssinicus Suwa et al., 2007 ≈10.0
3. Pongo 12.5 18.0 Sivapithecus Kelley, 2002 ≈12.5
4. Crown Catarrhini 21.0 33.9 Morotopithecus Gebo et al., 1997 >20.6
Victoriapithecus Pilbeam and Walker, 1968
Benefit and McCrossin, 2002
≈19.0
5. Theropithecus 3.5 6.5 Theropithecus Leakey, 1993
Frost, 2007
≈3.5
6. Macaca 5.5 NA Macaca libyca Delson 1980
Jablonski, 2002
≈5.5
7. Colobinae 10 NA Microcolobus tugenensis Benefit and Pickford, 1986; Jablonski, 2002 9.0–11.0
8. Aotus 12.1 NA Aotus dindensis Setoguchi and Rosenberger, 1987 ≈12.1
9. Saimiri 12.1 NA Neosaimiri Hartwig and Meldrum, 2002 ≈12.1
10. Crown Anthropoidea 31.5 NA Fayum catarrhines Rasmussen, 2002; Seiffert, 2005 31.5
11. Lorisoidea 36.9 NA Saharagalago
Karanisia
Seiffert et al., 2003
Seiffert, 2007
≈36.9
Non-primate fossils
12. Cetartyodactyla 48.6 NA Himalayacetus Bajpai And Gingerich, 1998 53.5
Pakicetus Gingerich and Russel 1981
Thewissen at al, 2001
>48.6
13. Mysticeti-Odontoceti 33.5 53.5 Llanocetus denticrenatus Mitchell, 1989 ≈3.5
14. Feliniforma-Caniformia 43.0 63.8 Tapocyon Benton and Donoghue, 2007 43.0–46.0
15. Bovidae 18.3 28.55 Eotragus noyi Solounias et al., 1995
Benton and Donoghue, 2007
≈18.3
16. Mus/Rattus 7.3 16.0 Progonomys hussaini Jacobs & Downs, 1994
Benton and Donoghue, 2007
11.1–11.5
17. Equus/Rhinoceros 54.0 58.0 Lambdotherium Rose, 2006 >50.3
*

see Supplementary Material B1 for the detailed procedure of fossil selection, with paleontological details, and assignment to alternative nodes for calibration of the molecular.

We evaluated the congruence of these 17 minimum-age constraints using fossil cross-calibration (Near and Sanderson, 2004; Near et al., 2005; Hugall et al., 2007; Rutschmann et al., 2007). Fossil cross-validation involves the use of one fossil at a time to generate age estimates for the nodes to which the other fossils are assigned; then it compares the inferred age with the fossil age at that node. This technique assesses the consistency of date estimates generated from different putative calibration points (Near and Sanderson, 2004; Near et al., 2005). We calculated two different statistics to evaluate the congruence of each calibration point. First, following the method proposed by Near et al., we calculate the mean deviation (D) between molecular (MA) and fossil age (FA) estimates for all nodes using a single fossil-dated node as a calibration point, and the SS values (SSx= Σi≠xD2) for a given fossil calibration node when it was used as the single calibration point. Second, we applied the statistics described by Hugall et al. (2007). We first calculate the mean deviation D as described above. Then, for each calibration point, we obtained the sum of the absolute values of differences between the estimated and proposed ages for each calibration node. This value (Σ|D|) is an indication of how congruent a specific fossil calibration is with the remainder: the smaller the value the more consistent. For these analyses, we used the topology and branch lengths of the MT39 model tree obtained in our Bayesian analyses. This topology was strongly favored in the Bayes factor comparison analysis explained above (see Table 3).

For all the analyses described above, we used the rate-autocorrelated Bayesian relaxed molecular clock method implemented in the MULTIDIVTIME program package (Thorne and Kishino, 2002). First, maximum likelihood parameters for the F84 + G model of evolution were estimated with baseml from the PAML package (Yang, 1997, 2007). Next, branch lengths and their variance-covariance matrix were calculated with estbranches. Finally divergence times were calculated with multidivtime using conservative priors and the fossil calibrations. Bayesian priors for multidivtime were chosen according to J. Thorne’s recommendations in the multidivtime manual. Root-to-tip mean was set at 100.0 Ma, with a 100.0 Ma standard deviation. The evolutionary rate at the root was set at 0.009 substitutions per nucleotide site per Myr, with a standard deviation of 0.009. This was calculated as the median root-to-tip branch length divided by the root-to-tip mean prior. The brownmean and its standard deviation were set to 0.020, such that the brownmean multiplied by the root-to-tip mean prior (100Ma) equals 2. After a burn-in period of 100,000 generations, MCMC chains were sampled every 100 generations until 10,000 samples were taken. To test the effect of the priors on the posteriors, several runs were performed with various prior choices. The results were extremely robust to changes in priors.

During the preliminary analyses fossil constraints were treated as point calibrations (minimum equal to maximum bounds). Based on the results of the preliminary analyses in which all 17 fossil calibrations were tested for congruence, a set of 16 constraints was used to obtain the final chronogram presented in the results. In this final analysis, for each fossil calibration we defined a minimum and a maximum constraint (if available) as reported in Table 4.

3. Results

3.1 Phylogeny

A comparison of the tree topologies that resulted from the analyses showed that varying partition scheme had little effect on the overall tree topology. For ML trees, between zero and six out of 168 possible conflicts were identified across partitions, representing a difference of only 0.00–3.57% (nodal distance: 0–0.474; split distance: 0–0.036). The differences were mainly represented by the position of two taxa: the genus Cavia within the rodents (either sister group of Cricetulus/(Mus/Rattus) or sister taxon of the dormouse, Glis glis) and the position of Cercopithecus diana in relation to the other cercopithecins. Within the MB trees, only 2 possible conflicts were found across different partitioning scheme (1.07%; nodal distance: 0–0.331; split distance: 0–0.0119). The position of C. diana was the only taxon with conflicting placements among different partitions.

Bayes factor comparisons strongly supported the model with 39 partitions (MT39), consisting of two partitions for the rRNAs (12S and 16S), one for the 22 tRNAs, and 36 partitions for the three codon positions of each protein-coding gene. Partition scheme MT39 was strongly favored over MT6 (lnBF = 5531.809), MT6 was favored over MT5 (lnBF = 1287.011), MT5 strongly favored over MT15 (lnBF = 5068.416), MT15 favored over MT4 (lnBF = 996.090), and MT4 favored over MT0 (lnBF = 1440.723) (Table 3). The same results were obtained using maximum likelihood (MT39<MT6<MT5<MT15<MT4<MT0). We therefore treat the MT39 ML and Bayesian trees as our “preferred tree”. Tree topology comparison analysis using TOPD-FMTS version 3.3 showed identical topologies between ML and Bayesian trees (0/168 nodes; Split Distance random [differents/possibles]: 0.997 +/− 0.007 [167.46 +/− 1.195/168 +/− 0.000]).

The Bayesian clade credibility values (PP) across the entire tree were high, with only two nodes below 0.95: the sister group relationship between C. diana and the clade including Erythrocebus patas and C.mitis/C.lhoesti (PP=0.91), and the sister group relationship between Erythrocebus patas and C.mitis/C.lhoesti (PP=0.80) (Figure 1). In contrast to the Bayesian results, the maximum likelihood bootstrap (BP) values were quite low in several nodes across the tree (Figure 1). Within primates several nodes were weakly supported in the ML analyses. For instance, within the platyrrhines, the sister relationship between Ateles and the clade including Aotus, Saimiri, Cebus, and Saguinus was below 70% bootstrap support Also, several nodes within cercopithecins (particularly, the position of E. patas and C. diana, and the sister group relationship between Nasalis and Pygathrix) showed bootstrap support below 70% (Figure 1). BP and PP values for all the nodes within primates are reported in Supplementary Material (Table C1).

Figure 1.

Figure 1

Single phylogenetic tree inferred from complete mitochondrial genome sequences based on maximum likelihood and Bayesian analyses. The tree topology remained constant across all analyses and branch lengths presented here follow the Bayesian analysis. The tree was rooted using Loxodonta africana and Elephas maximus. Maximum likelihood bootstrap values (BP) are shown with color coded inner circles (black: BP≥90%, grey 70%≤BP<90%; white: BP<70%); Bayesian posterior probability value (PP) are indicated by the outer ring (black: PP≥0.95, grey 0.75≤PP<0.95; white: PP<0.75).

In order to evaluate the influence of RNA sequences on the topology and support values, we also ran RAxML and MrBayes using only the 12 protein coding genes (11,022 bp). The topology of the reduced dataset did not differ from the one using all data. Similar to the ML results described above, BP values were generally lower than PP in Bayesian analyses (see Supplementary Figure D1–2).

All the analyses supported the same topology and the monophyly of the order Primates (Figure 1). Specifically, mitochondrial genome analyses supported the subdivision of primates into infraorders Strepsirrhini and Haplorhini, with tarsiers as sister group of Anthropoidea (BP>70% and PP>0.95). Moreover, the dermopterans (colugos, here represented by the genus Galeopterus) fell outside the primates (BP>70% and PP>0.95), and not the sister group of anthropoids as in many previous studies using mitochondrial DNA (Arnason and Janke, 2002; Arnason et al., 2002, 2008; Schmitz et al., 2002a).

We evaluated the effect of compositional biases and the level of non-phylogenetic signal using three additional datasets (3RY, AGY, and allRY). All analyses strongly supported (BP>70%) the monophyly of primates (with colugos outside primates), and two datasets (3RY and AGY) also strongly supported the sister group relationship between tarsiers and anthropoids. When all nucleotides were coded as RY (allRY), the monophyly of haplorhines (tarsiers+anthropoids) was only weakly supported (BP= 57%) (Supplementary Figures E1–3).

Relationships among the major lineages within primates were well supported by all the analyses, with lemuriformes and lorisoids within strepsirrhines, platyrrhines and catarrhines within anthropoids, and cercopithecoids (Cercopithecinae and Colobinae) and hominoids within catarrhines (Figure 1). Two major discrepancies between our results and widely accepted relationships within primates are found in the phylogenetic relationships within the lorisoids and the papionins. In the former clade, the division between galagids and lorisids was not supported by our mitogenomic data. The African lorisids were found to be the sister group to the clade composed of Asian lorisids and African galagids in all the analyses performed here. However, while Bayesian analyses strongly supported this topology (PP>0.95), bootstrap values were relatively low across all the analyses (BP<70%). Within the tribe Papionini, our results did not support the presence of an African clade (Theropithecus/Papio/Lophocebus/Mandrillus/Cercocebus). Instead, the group Mandrillus/Cercocebus clustered together with the genus Macaca (BP>70% and PP>0.95). Within these two primate groups (lorisoids and papionins), the same relationships were found when RY-coded datasets (3RY, AGY, and allRY) were analyzed.

3.2 Taxon sampling effects on phylogenetic reconstruction

In order to investigate the influence of taxon sampling in our dataset we analyzed four additional datasets including a subset of the primate taxa. Interestingly, all the datasets (6-taxa, 10-taxa, 14-taxa, and 26-taxa) showed no support for the monophyly of primates and the sister relationship between tarsiers and anthropoids. All reduced datasets supported instead the sister relationship between colugos and anthropoids and between strepsirrhines and tarsiers. However, the bootstrap values for both nodes were low (BP<70%) across all the analyses: 32–60% for the clade colugos-anthropoids and 45–68% for the clade tarsiers-strepsirrhines (Supplementary Figures F1–4).

3.3 Fossil congruence analyses

The cross-validation results (Table 5) reveal that calibration point 13 (C13: the split between Mysticeti and Odontoceti) is highly inconsistent with the other fossil calibrations. Specifically, when used as the only calibration point, C13 tended to overestimate all the other calibration points used in this study (Dx=23.52). Both statistics used in the study (SS and ΣD) supported this conclusion (Table 5). All the other fossil calibrations used in this study provide plausible age estimates that fall within the range of the other calibrations. We thus used 16 calibration points listed in Table 4 (excepting C13) for all dating analyses.

Table 5.

Cross-validation deviance of candidate calibration points.(Dx: average mean deviation between molecular and fossil age estimates for all available nodes based on the fossil calibration at node x; SS: sum of the squared differences between the molecular and fossil age estimates at all other fossil-dated nodes; ΣD: sum of the absolute values of differences between the estimated and proposed ages for each calibration node)

CP divergence age node
Dx SS ΣD
C1 C2 C3 C4 C5 C6 C7 C8 C9 C10 C11 C12 C13 C14 C15 C16 C17
C1 Homo 5.0 - 7.0 11.9 21.1 4.5 8.1 14.5 10.0 9.6 28.1 21.7 25.7 11.8 27.1 13.7 11.1 27.5 −5.60 2277.0 130.5
C2 Gorilla 10.0 7.2 - 17.0 30.4 6.5 11.6 20.9 14.4 13.8 40.6 31.4 37.1 17.0 39.2 19.8 16.0 39.7 −0.93 1099.6 110.94
C3 Pongo 12.5 5.4 7.5 - 21.9 4.8 8.5 15.2 10.4 10.0 29.0 22.3 26.2 11.9 27.7 13.9 11.3 28.1 −6.38 2181.2 127.77
C4 Crown Catarrhini 21.0 5.3 7.4 12.2 - 4.6 8.1 14.6 10.0 9.6 27.7 21.3 25.1 11.4 26.5 13.3 10.8 26.9 −5.21 2387.5 133.29
C5 Theropithecus 3.5 4.2 5.8 9.7 16.8 - 6.3 11.5 8.0 7.7 22.3 17.2 20.4 9.3 21.5 10.9 8.8 21.8 −9.75 3489.7 166.78
C6 Macaca 5.5 3.7 5.1 8.5 15.0 3.1 - 10.2 7.1 6.8 20.1 15.6 18.5 8.4 19.5 9.8 8.0 19.8 −11.56 4051.0 182.05
C7 Colobinae 10.0 3.6 5.0 8.4 14.5 3.1 5.6 - 6.9 6.6 19.2 14.8 17.5 7.9 18.5 9.3 7.5 18.7 −11.32 4336.2 188.3
C8 Aotus 12.1 6.6 9.2 15.2 26.3 5.7 10.2 18.3 - 11.7 35.0 27.1 32.2 14.8 33.9 17.2 13.9 34.4 −2.71 1373.5 110.84
C9 Saimiri 12.1 6.9 9.6 15.9 27.4 5.9 10.6 19.0 12.6 - 36.2 28.1 33.1 15.1 34.9 17.6 14.2 35.4 −1.88 1306.1 110.76
C10 Crown Anthropoidea 31.5 6.0 8.4 13.9 23.9 5.2 9.3 16.7 11.3 10.9 - 24.3 28.9 13.2 30.4 15.3 12.5 30.8 −4.92 1772.1 117.83
C11 Lorisoidea 36.9 9.6 13.4 22.1 37.9 8.3 14.8 26.5 18.0 17.4 49.7 - 44.8 20.5 47.3 23.8 19.3 48.1 4.81 1599.1 139.06
C12 Cetartyodactyla 48.6 10.8 15.0 24.7 42.4 9.3 16.5 29.7 20.3 19.5 55.4 42.2 - 22.1 52.0 25.7 21.4 52.7 8.05 2393.3 168.76
C13 Mysticeti-Odontoceti 33.5 15.5 21.5 35.6 61.5 13.4 23.8 42.7 29.2 28.1 81.1 62.3 73.5 - 77.6 39.1 31.9 78.7 23.52 11026.8 383.95
C14 Felis/Catus 43.0 8.8 12.3 20.3 35.0 7.6 13.6 24.4 16.7 16.0 45.9 35.1 41.1 18.8 - 21.9 17.9 43.8 8.06 1313.0 125.72
C15 Ovis/Bos 18.3 7.3 10.2 16.8 29.0 6.3 11.2 20.2 13.8 13.3 38.2 29.3 34.4 15.7 36.3 - 14.8 36.8 −0.58 1251.1 113.95
C16 Mus/Rattus 7.3 4.0 5.6 9.3 16.0 3.5 6.2 11.1 7.6 7.3 21.0 16.1 19.1 8.8 20.2 10.2 - 20.5 −10.29 3836.8 174.83
C17 Equus/Rhinoceros 54.0 11.1 15.4 25.4 43.6 9.6 17.0 30.5 20.8 20.0 57.1 43.5 51.1 23.4 53.5 27.2 22.3 - 9.21 2694.5 180.89

Because the relative position of fossil calibrations in the tree might influence the precision of the dating estimates for the node of interest, we tested whether the absolute values of the deviation between the estimated and proposed ages for each calibration node were correlated with the branch lengths (nodal distance) between the calibration point used in each run and the remaining 16 nodes. We performed a Mantel test to check the correlation between the matrix of the deviations |D| (index of the estimation error) and the matrix of nodal distances in term of branch length (R Development Core Team, 2009). The Mantel test based on 10,000 replicates produced a positive correlation (p-value: 0.038) with r=0.265. In order to evaluate the effect size of this correlation, we calculated the slope estimate from a linear regression. The analyses showed a positive relationship with β=4.7.

3.4 Divergence dates

Dating analyses were run with multidivtime using the MT39 partition strategy, as selected by the preliminary analyses described above. Estimates for the major nodes within primates are reported in Table 6, along with 95% HPD intervals. The chronogram estimated the origin of crown primates at 74.1 Ma (95% HPD=68.2–81.2; Figure 2 and Table 6). Both crown strepsirrhines and crown haplorhines are inferred to have originated in the Late Cretaceous, 66.3 Ma (95% HPD=61.1–72.8) and 70.0 Ma (95% HPD=64.3–76.7), respectively. The most recent common ancestor of the living platyrrhines was estimated to have lived 20.9 Ma (95% HPD=17.9–24.4), while the origin of crown catarrhines occurred considerably earlier, approximately 32.1 Ma (95% HPD=29.4–33.8). Homo sapiens and Neanderthals diverged at around 680,000 years ago (95% HPD=490–926,000), while the date of the most recent common mitochondrial ancestor between the Denisovans, Neanderthals and modern humans was estimated to be approximately 1.39 Ma (95% HPD=1.06–1.83). The estimates for all the nodes within primates are reported in the Supplementary Material (Table C1).

Table 6.

Comparison of divergence time estimates (posterior mean and 95% credible interval in in millions of years) between this study and recent studies on primates.

Divergence (node) This study Arnason et
al., 2008c, d
Matsui et
al., 2009d
Fabre et al.,
2009e
Chatterjee
et al., 2009 e
Perelman
et al., 2011f
Wilkinson
et al., 2011f
Jameson et
al., 2011f
Springer et
al., 2012e
Finstermeier
et al., 2013d
Crown Primates (61) 74.1 (68.2–81.2) 80.0 76.0 (69.382.5) 78.8 (69.988.4) 63.7 (58.368.7) 87.2 (75.998.6) 84.5 (69.2–103.5) 72.6 (69.676.9) 67.8 (61.076.7) 66.2 (59.673.6)
Crown Strepsirrhini (60) 66.3 (61.1–72.8) 70.0 64.5 (57.271.7) 67.1 (60.274.5) 51.6 (47.755.7) 68.7 (58.876.6) 49.8 (35.9–72.0) 52.4 (47.057.2) 54.2 (48.857.2) 56.9 (50.564.1)
Crown Lorisiformes (59) 40.3a (37.1–46.3) N/A 35.4 (28.543.1) 39.5 (38.041.8) 37.5 (36.938.7) 40.3 (35.245.6) N/A N/A 34.7 (27.938.2) 34.5 (30.239.0)
Lemuriformes-Chiromyiformes (55) 59.6 (54.3–65.9) 64.0 55.3 (47.763.0) 59.6 (53.366.7) 46.2 (41.350.8) 58.6 (38.676.8) N/A N/A 50.0 (45.253.8) 47.1 (40.953.6)
Crown Lemuriformes (54) 43.5 (37.5–50.1) N/A N/A 47.3 (37.950.8) 32.4 (28.633.6) 38.6 (26.450.0) 34.4 (23.549.1) N/A 31.8 (26.736.6) 35.5 (30.440.4)
Crown Haplorrhini (46) 70.0 (64.3–76.7) N/A 70.1 (63.276.7) N/A N/A 81.3 (69.595.8) N/A 68.6 (65.672.7) 61.2 (57.669.6) 63.1 (56.069.7)
Crown Anthropoidea (44) 46.7a (42.4–50.8) 55.0 45.3 (39.451.3) 37.7 (33.342.7) 42.8 (40.145.0) 43.5 (38.648.4) 47.2 (38.956.5) 40.0 (37.343.1) 40.6 (33.649.5) 45.3 (40.750.1)
Crown Platirrhini (43) 20.9 (17.9–24.4) N/A N/A 14.5 (9.719.9) 26.6 (23.530.0) 24.8 (20.629.3) 25.1 (20.1–31.0) N/A 23.3 (19.227.5) 22.0 (19.224.4)
Crown Catarrhini (38) 32.1b (29.4–33.8) 39.0 30.5 (28.835.3) 23.9 (23.125.9) 29.3 (28.030.0) 31.6 (25.737.9) 31.0 (25.1–37.7) 25.4 (23.727.6) 25.1 (19.732.8) 31.9 (28.335.7)
Crown Cercopithecoidea (37) 20.8a (18.6–22.9) N/A N/A 13.3 (11.614.7) 23.4 (21.924.9) 17.6 (13.921.5) 14.1 (11.0–17.7) N/A 13.2 (8.918.3) 22.8 (20.025.6)
Crown Colobinae (36) 14.1 (12.0–16.3) N/A N/A 8.7 (7.310.4) 18.4 (16.820.1) 12.3 (9.415.1) N/A N/A 8.9 (6.312.0) 15.5 (13.417.8)
Crown Cercopithecinae (29) 14.1 (12.2–15.8) N/A N/A 9.1 (7.710.0) 18.6 (16.520.5) 11.5 (9.213.9) 10.3 (8.113.0) N/A 8.4 (5.411.6) 14.9 (12.916.8)
Crown Hominoidea (12) 22.3 (20.5–23.9) N/A 19.9 (16.723.0) 18.6 (17.120.5) 21.5 (18.924.3) 20.3 (16.624.2) 19.2 (15.1–24.1) N/A 17.4 (12.423.9) 20.3 (17.423.5)
a

Lower bound defined as calibration point (see Table 4)

b

Lower and upper bound defined as calibration point (see Table 4)

c

Confidence intervals not provided

d

The study used entire mitochondrial genomes

e

The study used a supermatrix approach (mitochondrial and nuclear genes together)

f

The study used multiple nuclear loci (Perelman et al.: 54 loci ~35kb - Wilkinson et al.: 2 genomic region ~83kb – Jameson et al.: 1078 genes ~1.26 Mbp)

Figure 2.

Figure 2

Single chronogram with divergence date estimates from complete mitochondrial genome sequences. Only age estimates within primates are shown. Calibration points are depicted with filled circles; gray bars indicate constraint ranges. Mean node ages and 95% HPD intervals for major nodes within primates are presented in Table 6. Age estimates for all the nodes in the chronogram are reported in Supplementary Material (Table C1).

4. Discussion

In this study, we provided an assessment of primate phylogenetic relationships and divergence dates using complete mitochondrial genome data. We compiled one of the largest mitogenomic datasets for primates, and we applied a fossil cross-validation procedure to identify and select the most appropriate set of fossils to calibrate the molecular primate tree.

4.1 Phylogenetic conclusion

All major lineages within the order Primates were included in our analyses and were inferred to be monophyletic. In contrast to previous mitogenomic studies, our analyses posit a monophyletic Primates with colugos as their sister group within Euarchontoglires. Previous mitogenomic studies hypothesized a sister relationship between anthropoid primates and colugos (Dermoptera), making the order Primates paraphyletic (e.g., Arnason and Janke, 2002; Arnason et al., 2002; Schmitz et al., 2002a; Kjer and Honeycutt, 2007). Arnason et al. (2002) proposed the name Dermosimii for this assembly (Arnason and Janke, 2002; Arnason et al., 2002, 2008). However, this group has been suggested to be an artifact of similar nucleotide composition of mitochondrial genomes rather than reflecting the true evolutionary history (Schmitz et al., 2002a; Schmitz and Zischler, 2003; Raina et al., 2005; Zischler, 2007). In contrast to mitochondrial results, the monophyly of primates is in fact strongly supported not only by morphological studies (Silcox et al., 2007) but also by nuclear sequences (Jameson et al., 2011; Meredith et al., 2011; Perelman et al., 2011) and transposable elements (Schmitz et al., 2002a). In this study we suggest that this artifact might also be a consequence of taxon sampling and the small internode distance between Dermoptera and Primates. Incomplete or biased taxon sampling is a major concern in phylogenetic studies and can provide misleading results in reconstructing phylogenetic relationships (Zwickl and Hillis, 2002; Hillis et al., 2003; Plazzi et al., 2010; Townsend and Leuenberger, 2011; Nabhan and Sarkar, 2012), and increased sampling of taxa has been shown to be an effective way to improve overall phylogenetic accuracy (Zwickl and Hillis, 2002). Here, we show that a large taxon sampling within primates can recover phylogenetic relationships consistent with morphological and nuclear data, and supports the monophyly of primates and the invalidity of the taxon Dermosimii. When only a subset of primate taxa was included in the analyses, the position of colugos and tarsiers became poorly supported and all the analyses recovered a sister relationship between colugos and anthropoids similar to previous studies (e.g., Arnason and Janke, 2002; Arnason et al., 2002; Schmitz et al., 2002a; Kjer and Honeycutt, 2007), thus indicating that insufficient taxon sampling was likely the cause of the discrepancy found between mitochondria DNA and other lines of evidence (e.g., Schmitz et al., 2002a; Silcox et al., 2007; Jameson et al., 2011; Meredith et al., 2011; Perelman et al., 2011)

The difficulty of recovering primate monophyly using mitochondrial sequences is somewhat confirmed by lower support in the ML bootstrap analyses, compared to Bayeasian posterior probabilities. Theoretical and empirical studies showed that Bayesian posterior probability tends to be substantially higher than corresponding ML bootstrap frequencies. In particular, bootstrap analyses have been shown to be particularly sensitive to small numbers of characters and might underestimate confidence in tree topologies compared to Bayesian posterior probability in case of short internodes (Alfaro, 2003; Erixon et al., 2003; Rothfels et al., 2012). The low amount of phylogenetic signal in the mitochondrial genome for these nodes is demonstrated by the relatively short branch length of the internode between Dermoptera and Primates, and between Strepsirrhini and Haplorhini (Figure 1).

Within primates, we found a well-supported division between strepsirrhines (including lorises, galagids, and Malagasy lemurs) and haplorhines (including tarsiers and anthropoids). Overall, support values across the entire primate tree were high with only a few exceptions. Within cercopithecoids, the sister group relationship between C. diana and the clade including Erythrocebus patas and C.mitis/C.lhoesti and the sister group relationship between Erythrocebus patas and C.mitis/C.lhoesti were poorly supported in both Bayesian and ML analyses. Within the platyrrhines, most of the nodes showed low bootstrap support (but high Bayesian PP support), in agreement with previous studies that employed complete mitochondrial genomes (Hodgson et al., 2009; Chiou et al., 2011).

The phylogenetic position of the newly sequenced mitochondrial genomes was consistent with previous studies. The aye-aye (Daubentonia madagascariensis) was inferred as the most basal divergence within lemurs, while Mirza (member of the family Cheirogaleidae) was recovered as sister group of sportive lemurs (Lepilemur) (Perelman et al., 2011; Springer et al., 2012). Within Cercopithecoidea, the phylogenetic positions of the five guenon genomes (Allenopithecus nigroviridis, Cercopithecus diana, C. lhoesti, C. mitis, and Erythrocebus patas) were identical to the ones suggested in a larger mitogenomic study for guenons (Guschanski et al., 2013), but in contrast with nuclear data (Perelman et al., 2011; but see Guschanski et al., 2013).

Despite the large taxon representation in this study, we could not overcome some well-known problems affecting mitochondrial-based reconstructions of primate phylogenetic relationships. For example, two well-supported taxonomic groups, the African papionins and the lorisids, were not inferred to be monophyletic in our study. In the former group, our mitogenomic tree clusters Mandrillus and Cercocebus with Macaca instead of with the clade including Papio, Theropithecus, and Lophocebus. Within lorisoids, our mitogenomic tree places Asian lorisids as the sister group to the African galagids rather than to the African lorisid Perodicticus. Both the lorisids and the papionins are well supported by morphological data and nuclear sequences, and previous studies employing mitochondrial sequences have shown similar difficulty in recovering such clades (Papionini: Disotell et al., 1992; Harris, 2000; Finstermeier et al., 2013; Lorisidae: Roos et al., 2004; Masters et al., 2007; Matsui et al., 2009; Finstermeier et al., 2013; Pozzi et al., accepted). The mitochondrial genome represents a single locus and its phylogenetic history can differ from the species as a consequence of several events, including gene flow, hybridization, or incomplete lineage sorting (Maddison, 1997; Degnan and Rosenberg, 2006; Maddison and Knowles, 2006; Edwards, 2009). Although mitochondrial genomes are still a powerful resource to estimate phylogeny and phylogeography (Moore, 1995), conclusions based on a single locus should always be made with caution (Edwards, 2009; Ting and Sterner, 2013; Pozzi et al., 2014).

4.2 Divergence dates

To our knowledge, this is the first study that explores the application of fossil cross-validation within primates. The primate fossil record is extremely fragmented and incomplete, and some studies estimate that only 4–7% of all primate species are known from the fossil record (Tavaré et al., 2002; Soligo and Martin, 2006; Martin et al., 2007). The degree of incompleteness of the fossil record is particularly extreme in some lineages, most notably within the Malagasy lemurs in which no fossils are known with the possible – but controversial – exception of Bugtilemur mathesoni in Pakistan (see Marivaux et al., 2001, 2006). As a consequence, most of the molecular studies to date employed a small set of calibration points, representing only a few nodes within the primate tree (e.g., Fabre et al., 2009; Matsui et al., 2009). More recently, Wilkinson et al. (2011) tried to better integrate the current knowledge of the fossil record into molecular date estimates. They developed a new methodology to better account for the rates of fossil preservation and discovery when calibrating a molecule-based tree. This study dated the origin for the crown primates at ~84 Ma, confirming a discrepancy between molecular estimates and the known fossil record.

In this study, we present a cross validation analysis of the fossil record to select the most appropriate calibration points to be used with our data set. Our approach consisted of two different steps: 1) we selected a priori putative fossils suitable to be used as calibration points for our dataset, following the criteria proposed by Parham et al. (2012); 2) we performed a cross validation analyses aimed to detect possible outliers among our fossil selection. Since one of the main goals of this paper was to estimate the timing of primate origins we also included several calibration points within Boreoeutheria (Euarchontoglires and Laurasiatheria). Using only calibration points within primates might lead to problems in extrapolating dates, especially at the root of the tree (Raaum et al., 2005). In this study we selected sixteen fossils as calibration points, including eleven within and five outside primates. The original cross validation method proposed by Near et al. (2005) has been criticized by several authors because it could lead to several problems in selecting calibration points, including the possibility that the method will select an internally consistent set of fossils, excluding the most informative fossils because they are inconsistent with the majority of the remaining calibrations (Marshall, 2008; Ho and Phillips, 2009). Here, we used a conservative approach aimed to identify only possible outliers in the sets of fossils selected during the a priori phase. Although we acknowledge that the approach used here is not sufficient to address the complex challenges of calibration (Magallón, 2004; Near and Sanderson, 2004; Near et al., 2005; Hugall et al., 2007; Rutschmann et al., 2007; Pyron, 2010; Wilkinson et al., 2011), we believe that our approach allows researchers to better select fossils to be used in molecular phylogenetic analyses. For example, this approach might be beneficial in inferring the correct placement of fossils to specific nodes in a tree and excluding fossils that might be otherwise misused in calibrating molecular phylogenies. Moreover, new fossils can be tested against a good set of calibration points and alternative placements can be evaluated using the approach described here (see also Pyron, 2010 and Ward et al., 2010). This technique also allows researchers to include more fossils in their analyses, providing a better - and much needed - integration between molecular phylogenies and the paleontological record.

In order to identify how nodal distance can affect molecular estimates across the tree, we investigated the relationship between the deviation |D| and the nodal distance for each node selected as putative calibration point. Due to the incomplete nature of the fossil record, many areas of the primate tree have no available calibration point (e.g., the Malagasy lemurs). If the evolutionary distance from calibration points were a limiting factor in our ability to accurately estimate divergence times, then we would expect a strong correlation between nodal distance and the estimation error. The relationship that we found in our data set, although statistically significant, is particularly low. The Mantel test showed a weak positive correlation (r=0.265), suggesting that evolutionary distance from calibration points plays a very small role in the associated error. For instance, we found that the difference in estimated error of the crown Catarrhini divergence date between a near calibration point (Homo/Pan, branch length distance = 0.2588) and a more distant calibration point (Lorisidae/Galagidae, branch length = 0.7788) was only ±2.44 million years. For this reason, our confidence in divergences distant from calibrations will be similar to those closer to calibration points. This result is particularly important for some taxonomic groups in which the fossil record is particularly scarce or absent. Therefore, despite the negative relationship between nodal distance and accuracy, modern molecular dating techniques can successfully provide reliable estimates even in areas where no calibration points are available.

The molecular chronogram obtained in this study estimated the crown-group origin of the order Primates in the Late Cretaceous (around 74 Ma), in agreement with most recent molecular studies (Arnason et al., 2008; Fabre et al., 2009; Matsui et al., 2009; Jameson et al., 2011; Perelman et al., 2011; Wilkinson et al., 2011). Nuclear estimates seem to provide slightly older estimates than the ones obtained by mitogenomic data (Perelman et al., 2011; Wilkinson et al., 2011), however all molecular data to date seem to strongly suggest an origin of primates before the K-Pg boundary. This result agrees with estimates obtained for other group of mammals, where the majority of recent molecular studies on mammalian diversification place the origin for most orders in the Cretaceous (Springer et al., 2003; Bininda-Emonds et al., 2007; Meredith et al., 2011).

A recent study on mitochondrial genomes including over 80 species of primates, recovered more recent origins for the crown group, dated around 66Ma (59–73Ma) (Finstermeier et al., 2013). This discordance with our study is particularly interesting given the large overlap in genome sequences used in both studies. Two main differences may have contributed to the discrepant conclusions between the two studies. First, Finstermeier et al. (2013) only used calibration points within primates, making the age estimates at the root of the tree problematic due to extrapolation biases (see Raaum et al., 2005). The main differences in age estimates between Finstermeier et al.’s study and ours are limited to deeper nodes in the phylogeny. An additional source of discordance might be due to the use of controversial calibration points by Finstermeier et al. (2013). Out of the nine calibration points employed in that study, at least four were either secondary calibrations (i.e., molecular date estimates obtained from other molecular analyses) or based on misinterpretation of the fossil record (i.e., the use of stem fossils to calibrate a crown groups) (Graur and Martin 2004; Steiper and Young, 2008; Ho and Phillips, 2009; Pozzi et al., 2011). For instance, Finstermeier et al. (2013) calibrated the origins of crown catarrhines using the fossil Saadanius hijazensis (Zalmout et al., 2010). However, Saadanius is a stem catarrhine and cannot possibly inform the divergence between hominoids and catarrhines (see Pozzi et al., 2011).

Similarly to Finstermeier et al. (2013), a recent study conducted by Steiper and Seiffert (2012) also suggested a more recent origin for crown primates, closer to the K-Pg boundary or possibly even in the Paleocene. In their study, the authors showed a negative correlation between three different life history variables (body size, absolute endocranial volume and relative endocranial volume) and molecular rates in primates. While almost all divergence date estimates for placental mammals based on molecular data are broadly consistent with our results, it is possible that there are life history factors that all these studies have failed to take into account (Bromham, 2009, 2011; Lanfear et al., 2010; Steiper and Seiffert, 2012).

Finally, O’Leary et al. (2013) recently integrated molecular and morphological evidence to support an explosive model of evolution for mammals, with most the interordinal diversification dated just after the K-Pg boundary. Based on their analyses, the crown primates originated in the late Paleocene, around 55–56 Ma (O’Leary et al., 2013). The reconstruction proposed by O’Leary et al. (2013) is however based on phylogenetic analyses that fail to discriminate between homology and homoplasy and also implies extremely high acceleration in nucleotide substitution rate in early Paleocene mammals (Springer et al., 2013). One of the major shortcomings of O’Leary et al.’s study is the use of the earliest known fossil representative to date the origin of each group (‘ghost lineage’ analysis). As a consequence the origin of crown primates coincides with the earliest members of the group found in the fossil record at around 55–56 Ma (e.g., Teilhardina brandti) (O’Leary et al., 2013). However, the absence of evidence in the fossil record is not necessarily evidence of absence (Ho and Phillips, 2009) and the estimates obtained by O’Leary et al. (2013) can be considered only minimum divergence times based on ghost-lineage minimization rather than actual divergence times (Slater, 2013; Springer et al., 2013; Yoder, 2013; dos Reis et al., 2014).

The ancestor of the living catarrhines (apes and Old World monkeys) was estimated to have lived approximately 32 Ma. This result is in contrast to Zalmout et al. (2012) who suggested that the divergence for crown catarrhines dates after 29 Ma, based on the stem catarrhine Saadanius hijazensis. Inferring a date older than 29 Ma is not surprising because stem fossils do not provide any information about the divergence time of a crown group (Steiper and Young, 2008; Ho and Phillips, 2009); therefore, Saadanius does not inform the time of origin of crown catarrhines (Pozzi et al., 2011).

A recent study conducted by Langergraber et al. (2012) introduced a novel method to date divergence times within great apes without employing fossils as calibration points. This method relies on recent advances in the direct measurement of the mutation rate in humans (Awadalla et al., 2010; Roach et al., 2010; Sanders et al., 2012; Scally and Durban, 2012) and in the generation times within wild populations of chimpanzees and gorillas (Langergraber et al., 2012). Interestingly, the estimates obtained by Langergraber et al. are in line with the ones reported in this study. The only contrast with the above study is the age estimate for the divergence between bonobos (Pan paniscus) and common chimpanzee (Pan troglodytes) dated between 1.5 and 2.6 Ma by Langergraber et al. and 2.4–3.8 Ma in our study.

Finally, our estimates of 680 ka for Neanderthal and modern humans, and ~1.4 Ma for Denisovans and Homo sapiens are older - roughly 1.4x - than the ones obtained in the original paper by Krause et al. (2010) also using mitochondrial genomes. We identified the origin of this discrepancy in the different use of calibration points employed in the two studies. Krause et al. in fact assumed an average divergence of human and chimpanzee mtDNAs of 6 million years, while in our study we allowed this node to date between 5 and 10 Ma. We inferred this divergence to be at approximately 7.6 Ma (95% HPD=6.7–8.8), which is substantially older than the age used by Krause et al. (2010). When the split between Homo and Pan was set at 6 Ma, we obtained age estimates that agree with Krause et al. (H. sapiens-Neanderthal: 530 ka [397–693]; H. sapiens-Denisovans: 1.08 Ma [0.87–1.35]). We contend that the mitochondrial divergences obtained in our study are more realistic than the one presented by Krause at al. (2010), because the common ancestor of modern humans and chimpanzees is likely to be older than 6 Ma, based on both molecular estimates (Jameson et al., 2011; Wilkinson et al., 2011; Springer et al., 2012) and the fossil record (Haile-Selassie et al., 2001; Senut et al., 2001; Brunet et al., 2002; Vignaud et al., 2002).

4.3 Conclusion

Our study applied a fossil cross-validation procedure to identify a set of fossil calibrations to be used in dating a molecular phylogeny within primates. We used a two-step approach that includes an a priori selection of putative fossils and a posteriori cross validation analysis to exclude possible outliers. The set of calibrations obtained by such an analysis can then be used to estimate the ages of the nodes of interest. We therefore recommend the use of these fossil calibrations for future studies aimed to estimate divergence times within primates. Also, our study showed a weak correlation between nodal distance and accuracy, suggesting that distant calibration points can be reliably used for dating areas of the trees where the fossil record is particularly scant.

The results obtained in this study agree with recent studies of primate divergence dates, supporting an origin of the order in the Late Cretaceous. This suggests that the early primate fossil record is likely poorly sampled. Although primate evolution may involve a slowdown in molecular rates, most molecular studies to date suggest that it is highly unlikely that primates – together with most mammalian orders – originated after the K-Pg boundary. Based on the age estimated obtained in our study we suggest a short fuse model of primate origins, whereby relatively little time (<8Myr) passed between the origin of the order Primates and the diversification of its major clades, Strepsirrhini, Tarsiiformes, and Anthropoidea.

Supplementary Material

01

Supplementary Material Table A1. Details about the samples used in this study.

02. Supplementary Material Table B1. List of the fossils used as calibration points in this study.

For each node, reference fossil, type specimens, dates, and references are reported.

03

Supplementary Material Table C1. List of bootstrap values (maximum likelihood), posterior probabilities (MrBayes), mean divergence dates, and 95% composite credibility intervals in millions of years for all nodes within primates. Nodes are reported in Figure 2.

04. Supplementary Material Figure D1–2. Phylogenetic tree inferred from the 12 protein coding genes only.

1) Maximum likelihood analyses performed using RAxML 7.2.6. Bootstrap values (PP) are shown for each node. 2) Bayesian analysis obtained using MrBayes 3.2.1. Posterior probability (PP) values are shown for each node.

05. Supplementary Material Figure E1–3. Test for compositional biases.

Maximum likelihood trees with bootstrap values (the bootstopping criteria autoMRE (Majority Rule Criterion) as implemented in RAxML 7.2.6 (Pattengale et al., 2009) for three datasets: 1) third codon position in protein coding regions coded as RY [3RY], 2) positions with Cs and Ts replaced with Y [AGY], and 3) all nucleotides coded as purines (R) and pyrimidines (Y) [allRY].

06. Supplementary Material Figure F1–4. Test for taxon sampling effects.

Maximum likelihood trees with bootstrap values (the bootstopping criteria autoMRE [Majority Rule Criterion] as implemented in RAxML 7.2.6 [Pattengale et al., 2009]) for four datasets: 1) only six primate taxa (one for each main lineage within primates: Lemuriformes, Chyromyiformes, Lorisiformes, Tarsiiformes, Platyrrhini, Catarrhini), 2) ten taxa as used by Schmitz et al. (2002a), 3) 14 taxa as used by Arnason et al. (2008), and 4) 26 taxa as in Matsui et al. (2009).

07
08

HIGHLIGHTS.

  • We present one of the largest phylogenies of primates using complete mitochondrial genomes.

  • We included 62 primate species (representing all families) and 25 mammal species.

  • We recovered all the major clades within primates.

  • A cross-validation procedure was used to identify a set of fossil calibrations.

  • Crown primates diversified in the late Cretaceous supporting a short fuse model.

Acknowledgments

We thank Terry Harrison, Eric Delson, Stephen Frost, and Christina Bergey for helpful comments and suggestions. We also want to thank three anonymous reviewers and the editor for their comments that strongly improve the quality of our study. This research was partly supported by National Institutes of Health Grants R01-GM060760 and R24-GM65580.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Contributor Information

Luca Pozzi, Email: lpozzi@dpz.eu.

Jason A. Hodgson, Email: j.hodgson@imperial.ac.uk.

Andrew S. Burrell, Email: andrew.burrell@nyu.edu.

Kirstin N. Sterner, Email: ksterner@uoregon.edu.

Ryan L. Raaum, Email: ryan.raaum@lehman.cuny.edu.

Todd R. Disotell, Email: todd.disotell@nyu.edu.

References

  1. Abascal F, Zardoya R, Telford MJ. TranslatorX: multiple alignment of nucleotide sequences guided by amino acid translations. Nucleic acids research. 2010;38:W7–W13. doi: 10.1093/nar/gkq291. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Alfaro ME. Bayes or Bootstrap? A Simulation Study Comparing the Performance of Bayesian Markov Chain Monte Carlo Sampling and Bootstrapping in Assessing Phylogenetic Confidence. Molecular Biology and Evolution. 2003;20:255–266. doi: 10.1093/molbev/msg028. [DOI] [PubMed] [Google Scholar]
  3. Anderson S, De Bruijn MHL, Coulson AR, Eperon IC, Sanger F, Young IG. Complete sequence of bovine mitochondrial DNA conserved features of the mammalian mitochondrial genome. Journal of Molecular Biology. 1982;156:683–717. doi: 10.1016/0022-2836(82)90137-1. [DOI] [PubMed] [Google Scholar]
  4. 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. Proceedings of the National Academy of Sciences. 2002;99(12):8151–8156. doi: 10.1073/pnas.102164299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Arnason U, Adegoke JA, Gullberg A, Harley EH, Janke A, Kullberg M. Mitogenomic relationships of placental mammals and molecular estimates of their divergences. Gene. 2008;421:37–51. doi: 10.1016/j.gene.2008.05.024. [DOI] [PubMed] [Google Scholar]
  6. Arnason U, Gullberg A. Comparison between the complete mtDNA sequences of the blue and the fin whale, two species that can hybridize in nature. Journal of Molecular Evolution. 1993;37:312–322. doi: 10.1007/BF00178861. [DOI] [PubMed] [Google Scholar]
  7. Arnason U, Gullberg A, Gretarsdottir S, Ursing B. The mitochondrial genome of the sperm whale and a new molecular reference for estimating eutherian divergence dates. Journal of Molecular Evolution. 2000;50:569–578. doi: 10.1007/s002390010060. [DOI] [PubMed] [Google Scholar]
  8. Arnason U, Gullberg A, Janke A. Mitogenomic analyses provide new insights into cetacean origin and evolution. Gene. 2004;333:27–34. doi: 10.1016/j.gene.2004.02.010. [DOI] [PubMed] [Google Scholar]
  9. Arnason U, Gullberg A, Widegren B. The complete nucleotide sequence of the mitochondrial DNA of the fin whale, Balaenoptera physalus. Journal of Molecular Evolution. 1991;33:556–568. doi: 10.1007/BF02102808. [DOI] [PubMed] [Google Scholar]
  10. Arnason U, Janke A. Mitogenomic analyses of eutherian relationships. Cytogenetic and genome research. 2002;96:20–32. doi: 10.1159/000063023. [DOI] [PubMed] [Google Scholar]
  11. Awadalla P, Gauthier J, Myers R, Casals F, Hamdan FF, Griffing AR, Côté M, Henrion E, Spiegelman D, Tarabeux J, Piton A, Yang Y, Boyko A, Bustamante C, Xiong L, Rapoport JL, Addington AM, DeLisi JLE, Krebs M-O, Joober R, Millet B, Fombonne E, Mottron L, Zilversmit M, Keebler J, Daoud H, Marineau C, Roy-Gagnon M-H, Dubé M-P, Eyre-Walker A, Drapeau P, Stone E, Lafrenière RG, Rouleau G. Direct measure of the de novo mutation rate in autism and schizophrenia cohorts. American Journal of Human Genetics. 2010;87:316–24. doi: 10.1016/j.ajhg.2010.07.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Bajpai S, Gingerich P. A new Eocene archaeocete (Mammalia, Cetacea) from India and the time of origin of whales. Proceedings of the National Academy of Science. 1998;95:15464–15468. doi: 10.1073/pnas.95.26.15464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Bayona-Bafaluy MP. Revisiting the mouse mitochondrial DNA sequence. Nucleic Acids Research. 2003;31:5349–5355. doi: 10.1093/nar/gkg739. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Beard K, Houde P. An unusual assemblage of diminutive plesiadapiforms (Mammalia, ?Primates) from the early Eocene of the Clark’s Fork Basin, Wyoming. Journal of Vertebrate Paleontology. 1989;9:37–41. [Google Scholar]
  15. Beard KC. Gliding behaviour and palaeoecology of the alleged primate family Paromomyidae (Mammalia, Dermoptera) Nature. 1990;345:340–341. [Google Scholar]
  16. Beard KC. Origin and evolution of gliding in Early Cenozoic Dermoptera (Mammalia, Primatomorpha) In: MacPhee RDE, editor. Primates and their Relatives in Phylogenetic Perspective. Plenum; 1993. pp. 63–90. [Google Scholar]
  17. Benefit BR, McCrossin ML. The Victoriapithecidae, Cercopithecoidea. In: Hartwig WC, editor. The Primate Fossil Record. Cambridge University Press; Cambridge, UK: 2002. pp. 241–253. [Google Scholar]
  18. Benefit B, Pickford M. Miocene fossil cercopithecoids from Kenya. American Journal of Physical Anthropology. 1986;464:441–464. [Google Scholar]
  19. Benton MJ, Donoghue PCJ. Paleontological Evidence to Date the Tree of Life. Molecular Biology and Evolution. 2007;24:26–53. doi: 10.1093/molbev/msl150. [DOI] [PubMed] [Google Scholar]
  20. Bininda-Emonds ORP, Cardillo M, Jones KE, MacPhee RDE, Beck RMD, Grenyer R, Price SA, Vos RA, Gittleman JL, Purvis A. The delayed rise of present-day mammals. Nature. 2007;446:507–12. doi: 10.1038/nature05634. [DOI] [PubMed] [Google Scholar]
  21. Bloch JI, Boyer DM. Grasping primate origins. Science. 2002;298:1606–1610. doi: 10.1126/science.1078249. [DOI] [PubMed] [Google Scholar]
  22. Bloch JI, Silcox MT, Boyer DM, Sargis EJ. New Paleocene skeletons and the relationship of plesiadapiforms to crown-clade primates. Proceedings of the National Academy of Sciences of the United States of America. 2007;104:1159–64. doi: 10.1073/pnas.0610579104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Bloch JI, Silcox MT. New basicrania of Paleocene-Eocene Ignacius: re-evaluation of the Plesiadapiform-Dermopteran link. American Journal of Physical Anthropology. 2001;116:184–98. doi: 10.1002/ajpa.1114. [DOI] [PubMed] [Google Scholar]
  24. Briggs AW, Good JM, Green RE, Krause J, Maricic T, Stenzel U, Lalueza-Fox C, Rudan P, Brajkovic D, Kucan Z, Gusic I, Schmitz R, Doronichev VB, Golovanova LV, De la Rasilla M, Fortea J, Rosas A, Pääbo S. Targeted retrieval and analysis of five Neandertal mtDNA genomes. Science. 2009;325:318–21. doi: 10.1126/science.1174462. [DOI] [PubMed] [Google Scholar]
  25. Brinkmann H, Philippe H. Archaea sister group of Bacteria? Indications from tree reconstruction artifacts in ancient phylogenies. Molecular Biology and Evolution. 1999;16:817–825. doi: 10.1093/oxfordjournals.molbev.a026166. [DOI] [PubMed] [Google Scholar]
  26. Bromham L. Why do species vary in their rate of molecular evolution? Biology Letters. 2009;5:401–4. doi: 10.1098/rsbl.2009.0136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Bromham L. The genome as a life-history character: why rate of molecular evolution varies between mammal species. Philosophical transactions of the Royal Society of London. Series B, Biological sciences. 2011;366:2503–13. doi: 10.1098/rstb.2011.0014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Brunet M, Guy F, Pilbeam D, Mackaye HT, Likius A, Ahounta D, Beauvilain A, Blondel C, Bocherens H, Boisserie J-R, De Bonis L, Coppens Y, Dejax J, Denys C, Duringer P, Eisenmann V, Fanone G, Fronty P, Geraads D, Lehmann T, Lihoreau F, Louchart A, Mahamat A, Merceron G, Mouchelin G, Otero O, Pelaez Campomanes P, Ponce De Leon M, Rage J-C, Sapanet M, Schuster M, Sudre J, Tassy P, Valentin X, Vignaud P, Viriot L, Zazzo A, Zollikofer C. A new hominid from the Upper Miocene of Chad, Central Africa. Nature. 2002;418:145–51. doi: 10.1038/nature00879. [DOI] [PubMed] [Google Scholar]
  29. Burk A, Douzery E. The secondary structure of mammalian mitochondrial 16S rRNA molecules: refinements based on a comparative phylogenetic approach. Journal of Mammalian Evolution. 2002;9:225–252. [Google Scholar]
  30. Cartmill M. Arboreal adaptations and the origin of the order Primates. In: Tuttle RH, editor. The functional and evolutionary biology of primtes. Aldine-Atherton; 1972. pp. 97–122. [Google Scholar]
  31. Castresana J. Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Molecular Biology and Evolution. 2000:540–552. doi: 10.1093/oxfordjournals.molbev.a026334. [DOI] [PubMed] [Google Scholar]
  32. Chan Y-C, Roos C, Inoue-Murayama M, Inoue E, Shih C-C, Pei KJ-C, Vigilant L. Mitochondrial genome sequences effectively reveal the phylogeny of Hylobates gibbons. PloS One. 2010;5:e14419. doi: 10.1371/journal.pone.0014419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Chatterjee HJ, Ho SYW, Barnes I, Groves C. Estimating the phylogeny and divergence times of primates using a supermatrix approach. BMC Evolutionary Biology. 2009;9:259. doi: 10.1186/1471-2148-9-259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Chiou KL, Pozzi L, Lynch Alfaro JW, Di Fiore A. Pleistocene diversification of living squirrel monkeys (Saimiri spp.) inferred from complete mitochondrial genome sequences. Molecular Phylogenetics and Evolution. 2011;59:736–745. doi: 10.1016/j.ympev.2011.03.025. [DOI] [PubMed] [Google Scholar]
  35. Conroy GC. Primate Evolution. New York: W.W. Norton and Co; 1990. p. 492. [Google Scholar]
  36. Covert HH. The earliest fossil primates and the evolution of prosimians. In: Hartwig WC, editor. The Primate Fossil Record. Cambridge University Press; Cambridge, UK: 2002. pp. 13–20. [Google Scholar]
  37. Cummings MP, Otto SP, Wakeley J. Sampling properties of DNA sequence data in phylogenetic analysis. Molecular Biology and Evolution. 1995;12:814–22. doi: 10.1093/oxfordjournals.molbev.a040258. [DOI] [PubMed] [Google Scholar]
  38. Degnan JH, Rosenberg NA. Discordance of species trees with their most likely gene trees. PLoS Genetics. 2006;2:e68. doi: 10.1371/journal.pgen.0020068. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Delson E. Fossil macaques, phyletic relationships and a scenario of deployment. In: Lindburg DG, editor. The Macaques: Studies in Ecology, Behavior and Evolution. Van Nostrand; New York: 1980. pp. 10–30. [Google Scholar]
  40. Delsuc F, Phillips MJ, Penny D. Comment on “Hexapod Origins: Monophyletic or Paraphyletic?”. Science. 2003;301:1482. doi: 10.1126/science.1086558. [DOI] [PubMed] [Google Scholar]
  41. D’Erchia AM, Gissi C, Pesole G, Saccone C, Arnason U. The guinea-pig is not a rodent. Nature. 1996;381:597–600. doi: 10.1038/381597a0. [DOI] [PubMed] [Google Scholar]
  42. Disotell TR, Honeycutt RL, Ruvolo M. Mitochondrial DNA Phylogeny of the Old-World monkeytribe papionini. Molecular Biology and Evolution. 1992;9:1–13. doi: 10.1093/oxfordjournals.molbev.a040700. [DOI] [PubMed] [Google Scholar]
  43. Do CB, Foo C, Batzoglou S. A max-margin model for efficient simultaneous alignment and folding of RNA sequences. Bioinformatics. 2008;24:i68–76. doi: 10.1093/bioinformatics/btn177. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Do CB, Woods DA, Batzoglou S. CONTRAfold: RNA secondary structure prediction without physics-based models. Bioinformatics. 2006;22:e90–8. doi: 10.1093/bioinformatics/btl246. [DOI] [PubMed] [Google Scholar]
  45. Dornburg A, Beaulieu JM, Oliver JC, Near TJ. Integrating fossil preservation biases in the selection of calibrations for molecular divergence time estimation. Systematic Biology. 2011;60:519–27. doi: 10.1093/sysbio/syr019. [DOI] [PubMed] [Google Scholar]
  46. dos Reis M, Donoghue PCJ, Yang Z. Neither phylogenomic nor palaeontological data support a Palaeogene origin of placental mammals. Biol Lett. 2014;10:20131003. doi: 10.1098/rsbl.2013.1003. http://dx.doi.org/10.1098/rsbl.2013.1003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Edgar RC. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic acids research. 2004;32:1792–7. doi: 10.1093/nar/gkh340. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Edwards SV. Is a new and general theory of molecular systematics emerging? Evolution. 2009;63:1–19. doi: 10.1111/j.1558-5646.2008.00549.x. [DOI] [PubMed] [Google Scholar]
  49. Erixon P, Svennblad B, Britton T, Oxelman B. Reliability of Bayesian Posterior Probabilities and Bootstrap Frequencies in Phylogenetics. Systematic Biology. 2003;52:665–673. doi: 10.1080/10635150390235485. [DOI] [PubMed] [Google Scholar]
  50. Fabre P-H, Rodrigues A, Douzery EJP. Patterns of macroevolution among Primates inferred from a supermatrix of mitochondrial and nuclear DNA. Molecular Phylogenetics and Evolution. 2009;53:808–25. doi: 10.1016/j.ympev.2009.08.004. [DOI] [PubMed] [Google Scholar]
  51. Finstermeier K, Zinner D, Brameier M, Meyer M, Kreuz E, Hofreiter M, Roos C. A mitogenomic phylogeny of living primates. PloS one. 2013;8(7):e69504. doi: 10.1371/journal.pone.0069504. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Flynn SMC, Carr SM. Interspecies hybridization on DNA resequencing microarrays: efficiency of sequence recovery and accuracy of SNP detection in human, ape, and codfish mitochondrial DNA genomes sequenced on a human-specific MitoChip. BMC Genomics. 2007;8:339. doi: 10.1186/1471-2164-8-339. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Friedlaender JS, Friedlaender FR, Hodgson JA, Stoltz M, Koki G, Horvat G, Zhadanov S, Schurr TG, Merriwether DA. Melanesian mtDNA complexity. PloS One. 2007;2:e248. doi: 10.1371/journal.pone.0000248. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Frost S. African Pliocene and Pleistocene cercopithecid evolution and global climatic change. In: Bobe R, Alemseged Z, editors. Hominin Environments in the East African Pliocene: An assessment of the Faunal Evidence. New York: Springer; 2007. pp. 51–76. [Google Scholar]
  55. Gebo DL, MacLatchy L, Kityo R, Deino A, Kingston J, Pilbeam D. A hominoid genus from the early Miocene of Uganda. Science. 1997;276:401–404. doi: 10.1126/science.276.5311.401. [DOI] [PubMed] [Google Scholar]
  56. Gingerich PD. Early Eocene Cantius torresi—oldest primate of modern aspect from North America. Nature. 1986;319:319–321. [Google Scholar]
  57. Gingerich PD, Russell DE. Contributions from the Museum of Paleontology. 11. Vol. 25. University of Michigan; 1981. Pakicetus inachus, a new archaeocete (Mammalia, Cetacea) from the early-middle Eocene Kuldana Formation of Kohat (Pakistan) pp. 235–246. [Google Scholar]
  58. Gissi C, Gullberg A, Arnason U. The Complete Mitochondrial DNA Sequence of the Rabbit, Oryctolagus cuniculus. Genomics. 1998;169:161–169. doi: 10.1006/geno.1998.5282. [DOI] [PubMed] [Google Scholar]
  59. Gissi C, Reyes A, Pesole G, Saccone C. Lineage-specific evolutionary rate in mammalian mtDNA. Molecular Biology and Evolution. 2000;17:1022–31. doi: 10.1093/oxfordjournals.molbev.a026383. [DOI] [PubMed] [Google Scholar]
  60. Gokey NG, Cao Z, Pak JW, Lee D, McKiernan SH, McKenzie D, Weindruch R, Aiken JM. Molecular analyses of mtDNA deletion mutations in microdissected skeletal muscle fibers from aged rhesus monkeys. Aging cell. 2004;3:319–26. doi: 10.1111/j.1474-9728.2004.00122.x. [DOI] [PubMed] [Google Scholar]
  61. Graur D, Martin W. Reading the entrails of chickens: molecular timescales of evolution and the illusion of precision. Trends in Genetics. 2004;20:80–86. doi: 10.1016/j.tig.2003.12.003. [DOI] [PubMed] [Google Scholar]
  62. Green RE, Malaspinas A, Krause J, Briggs AW, Johnson PLF, Uhler C, Meyer M, Good JM, Maricic T, Stenzel U, Prüfer K, Siebauer M, Burbano HA, Ronan M, Rothberg JM, Egholm M, Rudan P, Brajković D, Kućan Z, Gusić I, Wikström M, Laakkonen L, Kelso J, Slatkin M, Pääbo S. A complete Neandertal mitochondrial genome sequence determined by high-throughput sequencing. Cell. 2008;134:416–26. doi: 10.1016/j.cell.2008.06.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Guschanski K, Krause J, Sawyer S, Valente LM, Bailey S, Finstermeier K, Sabin R, Gilissen E, Sonet G, Nagy ZT, Lenglet G, Mayer F, Savolainen V. Next-Generation Museomics Disentangles One of the Largest Primate Radiations. Systematic biology. 2013;62(4):539–554. doi: 10.1093/sysbio/syt018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Haile-Selassie Y. Late Miocene hominids from the middle Awash, Ethiopia. Nature. 2001;412:178–181. doi: 10.1038/35084063. [DOI] [PubMed] [Google Scholar]
  65. Harris EE. Molecular systematics of the old world monkey tribe papionini: analysis of the total available genetic sequences. Journal of Human Evolution. 2000;38:235–56. doi: 10.1006/jhev.1999.0318. [DOI] [PubMed] [Google Scholar]
  66. Hartwig WC, Meldrum DJ. Miocene platyrrhines of the northern Neotropics. In: Hartwig W, editor. The Primate Fossil Record. Cambridge University Press; Cambridge, UK: 2002. pp. 175–189. [Google Scholar]
  67. Hauf J, Waddell P, Chalwatzis N, Joger U, Zimmermann F. The complete mitochondrial genome sequence of the African elephant (Loxodonta africana), phylogenetic relationships of Proboscidae to other mammals, and D-loop heteroplasmy. Zoology. 2000;102:184–195. [Google Scholar]
  68. Hiendleder S, Mainz K. Mitochondrial DNA indicates that domestic sheep are derived from two different ancestral maternal sources: no evidence for contributions from urial and argali sheep. Journal of Heredity. 1998;89:113–120. doi: 10.1093/jhered/89.2.113. [DOI] [PubMed] [Google Scholar]
  69. Hillis D, Bull J. An Empirical Test of Bootstrapping as a Method for Assessing Confidence in Phylogenetic Analysis. Systematic biology. 1993;42:182–192. [Google Scholar]
  70. Hillis DM, Pollock DD, McGuire JA, Zwickl DJ. Is Sparse Taxon Sampling a Problem for Phylogenetic Inference? Systematic Biology. 2003;52:124–126. doi: 10.1080/10635150390132911. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Ho SYW, Phillips MJ. Accounting for Calibration Uncertainty in Phylogenetic Estimation of Evolutionary Divergence Times. Systematic Biology. 2009;58:367–380. doi: 10.1093/sysbio/syp035. [DOI] [PubMed] [Google Scholar]
  72. Hodgson JA, Sterner KN, Matthews LJ, Burrell AS, Rachana AJ, Raaum RL, Stewart C-B, Disotell TR. Successive radiations, not stasis, in the South American primate fauna. Proceedings of the National Academy of Sciences. 2009;106:5534–5539. doi: 10.1073/pnas.0810346106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Horai S, Hayasaka K, Kondo R, Tsugane K, Takahata N. Recent African origin of modern humans revealed by complete sequences of hominoid mitochondrial DNAs. Proceedings of the National Academy of Science. 1995;92:532–536. doi: 10.1073/pnas.92.2.532. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Huelsenbeck JP, Rannala B. Frequentist properties of Bayesian posterior probabilities of phylogenetic trees under simple and complex substitution models. Systematic Biology. 2004;53:904–13. doi: 10.1080/10635150490522629. [DOI] [PubMed] [Google Scholar]
  75. Huelsenbeck JP, Ronquist F. MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics. 2001;17:754–755. doi: 10.1093/bioinformatics/17.8.754. [DOI] [PubMed] [Google Scholar]
  76. Hugall AF, Foster R, Lee MSY. Calibration choice, rate smoothing, and the pattern of tetrapod diversification according to the long nuclear gene RAG-1. Systematic Biology. 2007;56:543–63. doi: 10.1080/10635150701477825. [DOI] [PubMed] [Google Scholar]
  77. Inoue JG, Miya M, Lam K, Tay B-H, Danks JA, Bell J, Walker TI, Venkatesh B. Evolutionary origin and phylogeny of the modern holocephalans (Chondrichthyes: Chimaeriformes): a mitogenomic perspective. Molecular Biology and Evolution. 2010;27:2576–86. doi: 10.1093/molbev/msq147. [DOI] [PubMed] [Google Scholar]
  78. Inoue JG, Miya M, Tsukamoto K, Nishida M. Evolution of the deep-sea gulper eel mitochondrial genomes: large-scale gene rearrangements originated within the eels. Molecular Biology and Evolution. 2003;20:1917–24. doi: 10.1093/molbev/msg206. [DOI] [PubMed] [Google Scholar]
  79. IUCN. IUCN Red List of Threatened Species. Version 2013.2. 2013 www.iucnredlist.org.
  80. Jablonski NG. Fossil Old World monkeys: the late Neogene radiation. In: Hartwig WC, editor. The Primate Fossil Record. Cambridge University Press; Cambridge, UK: 2002. pp. 255–299. [Google Scholar]
  81. Jacobs LL, Downs WR. The evolution of murine rodents in Asia. In: Tomida Y, Li CK, Setoguschi T, editors. Rodent and lagomorph families of Asian origins and their diversification. National Science Museum Monograph; Tokyo: 1994. pp. 149–156. [Google Scholar]
  82. Jameson NM, Hou Z-C, Sterner KN, Weckle A, Goodman M, Steiper ME, Wildman DE. Genomic data reject the hypothesis of a prosimian primate clade. Journal of Human Evolution. 2011;61:295–305. doi: 10.1016/j.jhevol.2011.04.004. [DOI] [PubMed] [Google Scholar]
  83. Kass RE, Raftery AE. Bayes Factors. Journal of American Statistical Association. 1995;90:773–795. [Google Scholar]
  84. Kay RF, Thorington RW. Eocene plesiadapiform shows affinities with flying lemurs not primates. Nature. 1990;345:342–344. [Google Scholar]
  85. Kay RF, Thewissen JGM, Yoder AD. Cranial Anatomy of lgnacius graybullianus and the Affinities of the Plesiadapiformes. American Journal of Physical Anthropology. 1992;89:477–498. [Google Scholar]
  86. Kelley J. The hominoid radiation in Asia. In: Hartwig W, editor. The Primate Fossil Record. Cambridge University Press; Cambridge: 2002. pp. 369–384. [Google Scholar]
  87. Kim K, Lee S, Jeong H, Ha J. The complete nucleotide sequence of the domestic dog (Canis familiaris) mitochondrial genome. MolecularPhylogenetics and Evolution. 1998;10:210–220. doi: 10.1006/mpev.1998.0513. [DOI] [PubMed] [Google Scholar]
  88. Kjer KM, Honeycutt RL. Site specific rates of mitochondrial genomes and the phylogeny of eutheria. BMC Evolutionary Biology. 2007;7:8. doi: 10.1186/1471-2148-7-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Krause J, Fu Q, Good JM, Viola B, Shunkov MV, Derevianko AP, Pääbo S. The complete mitochondrial DNA genome of an unknown hominin from southern Siberia. Nature. 2010;464:894–7. doi: 10.1038/nature08976. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Lanfear R, Welch JJ, Bromham L. Watching the clock: studying variation in rates of molecular evolution between species. Trends in Ecology & Evolution. 2010;25:495–503. doi: 10.1016/j.tree.2010.06.007. [DOI] [PubMed] [Google Scholar]
  91. Langergraber KE, Prüfer K, Rowney C, Boesch C, Crockford C, Fawcett K, Inoue E, Inoue-Muruyama M, Mitani JC, Muller MN, Robbins MM, Schubert G, Stoinski TS, Viola B, Watts D, Wittig RM, Wrangham RW, Zuberbühler K, Pääbo S, Vigilant L, Prufer K, Zuberbuhler K. Generation times in wild chimpanzees and gorillas suggest earlier divergence times in great ape and human evolution. Proceedings of the National Academy of Sciences of the United States of America. 2012;109:15716–21. doi: 10.1073/pnas.1211740109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Leakey M. Evolution of Theropithecus in the Turkana Basin. In: JNG, editor. Theropithecus: The Rise and Fall of a Primate Genus. Cambridge,UK: CambridgeUniv Press; 1993. pp. 85–123. [Google Scholar]
  93. Lei R, Shore GD, Brenneman RA, Engberg SE, Sitzmann BD, Bailey CA, Kimmel LM, Randriamampionona R, Ranaivoarisoa JF, Louis EE. Complete sequence and gene organization of the mitochondrial genome for Hubbard’s sportive lemur (Lepilemur hubbardorum) Gene. 2010;464:44–9. doi: 10.1016/j.gene.2010.06.001. [DOI] [PubMed] [Google Scholar]
  94. Letsch HO, Kück P, Stocsits RR, Misof B. The impact of rRNA secondary structure consideration in alignment and tree reconstruction: simulated data and a case study on the phylogeny of hexapods. Molecular Biology and Evolution. 2010;27:2507–21. doi: 10.1093/molbev/msq140. [DOI] [PubMed] [Google Scholar]
  95. Li D, Fan L, Zeng B, Yin H, Zou F, Wang H, Meng Y, King E, Yue B. The complete mitochondrial genome of Macaca thibetana and a novel nuclear mitochondrial pseudogene. Gene. 2009;429:31–6. doi: 10.1016/j.gene.2008.10.010. [DOI] [PubMed] [Google Scholar]
  96. Lin Y, Waddell P, Penny D. Pika and vole mitochondrial genomes increase support for both rodent monophyly and glires. Gene. 2002;294:119–129. doi: 10.1016/s0378-1119(02)00695-9. [DOI] [PubMed] [Google Scholar]
  97. Lopez J, Cevario S, O’Brien S. Complete nucleotide sequences of the domestic cat (Felis catus) mitochondrial genome and a transposed mtDNA tandem repeat (Numt) in the nuclear genome. Genomics. 1996;246:229–246. doi: 10.1006/geno.1996.0188. [DOI] [PubMed] [Google Scholar]
  98. Maddison WP, Knowles LL. Inferring phylogeny despite incomplete lineage sorting. Systematic Biology. 2006;55:21. doi: 10.1080/10635150500354928. [DOI] [PubMed] [Google Scholar]
  99. Maddison WP. Gene Trees in Species Trees. Systematic Biology. 1997;46:523. [Google Scholar]
  100. 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. doi: 10.1038/35054544. [DOI] [PubMed] [Google Scholar]
  101. Magallón S. Dating lineages: molecular and paleontological approaches to the temporal framework of clades. International Journal of Plant Sciences. 2004;165 [Google Scholar]
  102. Marivaux L, Chaimanee Y, Tafforeau P, Jaeger J. New strepsirrhine primate from the late Eocene of Peninsular Thailand (Krabi Basin) American Journal of Physical Anthropology. 2006;130:425–34. doi: 10.1002/ajpa.20376. [DOI] [PubMed] [Google Scholar]
  103. Marivaux L, Welcomme JL, Antoine PO, Métais G, Baloch IM, Benammi M, Chaimanee Y, Ducrocq S, Jaeger JJ. A fossil lemur from the Oligocene of Pakistan. Science (New York, NY) 2001;294:587–91. doi: 10.1126/science.1065257. [DOI] [PubMed] [Google Scholar]
  104. Marshall CR. A simple method for bracketing absolute divergence times on molecular phylogenies using multiple fossil calibration points. The American Naturalist. 2008;171:726–42. doi: 10.1086/587523. [DOI] [PubMed] [Google Scholar]
  105. Martin R. Primate origins: plugging the gaps. Nature. 1993;363:223–234. doi: 10.1038/363223a0. [DOI] [PubMed] [Google Scholar]
  106. Martin RD, Soligo C, Tavaré S. Primate origins: implications of a cretaceous ancestry. Folia Primatologica. 2007;78:277–96. doi: 10.1159/000105145. [DOI] [PubMed] [Google Scholar]
  107. Martin RD. Towards a new definition of Primates. Man New Series. 1968;3:377–401. [Google Scholar]
  108. Mason VC, Li G, Helgen KM, Murphy WJ. Efficient cross-species capture hybridization and next-generation sequencing of mitochondrial genomes from noninvasively sampled museum specimens. Genome Research. 2011;21:1695–704. doi: 10.1101/gr.120196.111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Masters JC, Boniotto M, Crovella S, Roos C, Pozzi L, Delpero M. Phylogenetic relationships among the Lorisoidea as indicated by craniodental morphology and mitochondrial sequence data. American Journal of Primatology. 2007;69:6–15. doi: 10.1002/ajp.20322. [DOI] [PubMed] [Google Scholar]
  110. Matsudaira K, Ishida T. Phylogenetic relationships and divergence dates of the whole mitochondrial genome sequences among three gibbon genera. Molecular Phylogenetics and Evolution. 2010;55:454–9. doi: 10.1016/j.ympev.2010.01.032. [DOI] [PubMed] [Google Scholar]
  111. Matsui A, Rakotondraparany F, Munechika I, Hasegawa M, Horai S. Molecular phylogeny and evolution of prosimians based on complete sequences of mitochondrial DNAs. Gene. 2009;441:53–66. doi: 10.1016/j.gene.2008.08.024. [DOI] [PubMed] [Google Scholar]
  112. Meredith RW, Janečka JE, Gatesy J, Ryder O, Fisher C, Teeling EC, Goodbla A, Eizirik E, Simão TLL, Stadler T, Rabosky DL, Honeycutt RL, Flynn JJ, Ingram CM, Steiner C, Williams TL, Robinson TJ, Burk-Herrick A, Westerman M, Ayoub N, Springer MS, Murphy WJ. Impacts of the Cretaceous Terrestrial Revolution and KPg extinction on mammal diversification. Science (New York, NY) 2011;334:521–4. doi: 10.1126/science.1211028. [DOI] [PubMed] [Google Scholar]
  113. Miller ER, Gunnell GF, Martin RD. Deep time and the search for anthropoid origins. American Journal of Physical Anthropology Suppl. 2005;41:60–95. doi: 10.1002/ajpa.20352. [DOI] [PubMed] [Google Scholar]
  114. Mitchell ED. A new cetacean from the late Eocene La Meseta Formation, Seymour Island, Antarctic Peninsula. Canadian Journal of Fisheries and Aquatic Sciences. 1989;46(12):2219–2235. [Google Scholar]
  115. Miya M, Nishida M. Use of mitogenomic information in teleostean molecular phylogenetics: a tree-based exploration under the maximum-parsimony optimality criterion. Molecular Phylogenetics and Evolution. 2000;17:437–55. doi: 10.1006/mpev.2000.0839. [DOI] [PubMed] [Google Scholar]
  116. Miya M, Takeshima H, Endo H, Ishiguro NB, Inoue JG, Mukai T, Satoh TP, Yamaguchi M, Kawaguchi A, Mabuchi K, Shirai SM, Nishida M. Major patterns of higher teleostean phylogenies: a new perspective based on 100 complete mitochondrial DNA sequences. Molecular Phylogenetics and Evolution. 2003;26:121–38. doi: 10.1016/s1055-7903(02)00332-9. [DOI] [PubMed] [Google Scholar]
  117. Moore W. Inferring Phylogenies from mtDNA Variation : Mitochondrial-Gene Trees Versus Nuclear-Gene Trees. Evolution. 1995;49:718–726. doi: 10.1111/j.1558-5646.1995.tb02308.x. [DOI] [PubMed] [Google Scholar]
  118. 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. doi: 10.1038/35054550. [DOI] [PubMed] [Google Scholar]
  119. Nabhan AR, Sarkar IN. The impact of taxon sampling on phylogenetic inference: a review of two decades of controversy. Briefings in Bioinformatics. 2012;13:122–34. doi: 10.1093/bib/bbr014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Near TJ, Meylan P, Shaffer H. Assessing concordance of fossil calibration points in molecular clock studies: an example using turtles. American Naturalist. 2005;165 doi: 10.1086/427734. [DOI] [PubMed] [Google Scholar]
  121. Near TJ, Sanderson MJ. Assessing the quality of molecular divergence time estimates by fossil calibrations and fossil-based model selection. Philosophical transactions of the Royal Society of London. Series B, Biological sciences. 2004;359:1477–83. doi: 10.1098/rstb.2004.1523. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Newton M. Approximate Bayesian Inference with the Weighted Likelihood Bootstrap. Journal of the Royal Statistical Society. Series B. 1994;56:3–48. [Google Scholar]
  123. Ni X, Wang Y, Hu Y, Li C. A euprimate skull from the early Eocene of China. Nature. 2004;427:65–8. doi: 10.1038/nature02126. [DOI] [PubMed] [Google Scholar]
  124. Nilsson MA, Gullberg A, Spotorno AE, Arnason U, Janke A. Radiation of extant marsupials after the K/T boundary: evidence from complete mitochondrial genomes. Journal of Molecular Evolution. 2003;57(Suppl 1):S3–12. doi: 10.1007/s00239-003-0001-8. [DOI] [PubMed] [Google Scholar]
  125. Nylander J, Wilgenbusch JC, Warren DL, Swofford DL. AWTY (are we there yet?): a system for graphical exploration of MCMC convergence in Bayesian phylogenetics. Bioinformatics. 2008;24:581–3. doi: 10.1093/bioinformatics/btm388. [DOI] [PubMed] [Google Scholar]
  126. Nylander J, Ronquist F, Huelsenbeck J, Nieves-Aldrey J. Bayesian Phylogenetic Analysis of Combined Data. Systematic Biology. 2004;53:47–67. doi: 10.1080/10635150490264699. [DOI] [PubMed] [Google Scholar]
  127. Nylander J. MrModeltest v2. Program distributed by the author. Evolutionary Biology Centre, Uppsala University; 2004. [Google Scholar]
  128. O’Leary MA, Bloch JI, Flynn JJ, Gaudin TJ, Giallombardo A, Giannini NP, Goldberg SL, Kraatz BP, Luo Z-X, Meng J, Ni X, Novacek MJ, Perini FA, Randall ZS, Rougier GW, Sargis EJ, Silcox MT, Simmons NB, Spaulding M, Velazco PM, Weksler M, Wible JR, Cirranello AL. The Placental Mammal Ancestor and the Post-K-Pg Radiation of Placentals. Science. 2013;339:662–667. doi: 10.1126/science.1229237. [DOI] [PubMed] [Google Scholar]
  129. Pacheco MAM, Battistuzzi FFU, Lentino M, Aguilar R, Kumar S, Escalante AA. Evolution of modern birds revealed by Mitogenomics: timing the radiation and origin of major orders. Molecular Biology and Evolution. 2011 doi: 10.1093/molbev/msr014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Parham JF, Donoghue PCJ, Bell CJ, Calway TD, Head JJ, Holroyd PA, Inoue JG, Irmis RB, Joyce WG, Ksepka DT, Patané JSL, Smith ND, Tarver JE, Van Tuinen M, Yang Z, Angielczyk KD, Greenwood JM, Hipsley CA, Jacobs L, Makovicky PJ, Müller J, Smith KT, Theodor JM, Warnock RCM, Benton MJ. Best practices for justifying fossil calibrations. Systematic Biology. 2012;61:346–59. doi: 10.1093/sysbio/syr107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Partridge MA, Davidson MM, Hei TK. The complete nucleotide sequence of Chinese hamster (Cricetulus griseus) mitochondrial DNA. DNA sequence : the journal of DNA sequencing and mapping. 2007;18:341–6. doi: 10.1080/10425170601101287. [DOI] [PubMed] [Google Scholar]
  132. Pattengale ND, Alipour M, Bininda-Emonds ORP, Moret BME, Stamatakis A. Research in Computational Molecular Biology. Springer; Berlin Heidelberg: 2009. How many bootstrap replicates are necessary? pp. 184–200. [DOI] [PubMed] [Google Scholar]
  133. Pereira SL, Baker AJ. A mitogenomic timescale for birds detects variable phylogenetic rates of molecular evolution and refutes the standard molecular clock. Molecular Biology and Evolution. 2006;23:1731–40. doi: 10.1093/molbev/msl038. [DOI] [PubMed] [Google Scholar]
  134. Perelman P, Johnson WE, Roos C, Seuánez HN, Horvath JE, Moreira MAM, Kessing B, Pontius J, Roelke M, Rumpler Y, Schneider MPC, Silva A, O’Brien SJ, Pecon-Slattery J. A molecular phylogeny of living primates. PLoS Genetics. 2011;7:e1001342. doi: 10.1371/journal.pgen.1001342. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Phillips MJ. Branch-length estimation bias misleads molecular dating for a vertebrate mitochondrial phylogeny. Gene. 2009;441(1):132–140. doi: 10.1016/j.gene.2008.08.017. [DOI] [PubMed] [Google Scholar]
  136. Phillips MJ, Penny D. The root of the mammalian tree inferred from whole mitochondrial genomes. Molecular Phylogenetics and Evolution. 2003;28:171–185. doi: 10.1016/s1055-7903(03)00057-5. [DOI] [PubMed] [Google Scholar]
  137. Phillips MJ, Delsuc F, Penny D. Genome-scale phylogeny and the detection of systematic biases. Molecular biology and evolution. 2004;21:1455–1458. doi: 10.1093/molbev/msh137. [DOI] [PubMed] [Google Scholar]
  138. Pilbeam DR, Walker AC. Fossil monkeys from the Miocene of Napak, Northeast Uganda. Nature. 1968;220 (5168):657–60. doi: 10.1038/220657a0. [DOI] [PubMed] [Google Scholar]
  139. Plazzi F, Ferrucci RR, Passamonti M. Phylogenetic representativeness: a new method for evaluating taxon sampling in evolutionary studies. BMC Bioinformatics. 2010;11:209. doi: 10.1186/1471-2105-11-209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Posada D, Buckley TR. Model selection and model averaging in phylogenetics: advantages of akaike information criterion and bayesian approaches over likelihood ratio tests. Systematic Biology. 2004;53:793–808. doi: 10.1080/10635150490522304. [DOI] [PubMed] [Google Scholar]
  141. Pozzi L, Bergey CM, Burrell AS. The use (and misuse) of phylogenetic trees in comparative behavioral analyses. International Journal of Primatology. 2014;35:32–54. [Google Scholar]
  142. Pozzi L, Disotell TR, Masters JC. A multilocus phylogeny reveals deep lineages within African galagids (Primates: Galagidae) Accepted pending minor revisions in BMC Evolutionary Biology. doi: 10.1186/1471-2148-14-72. (accepted) [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Pozzi L, Hodgson JA, Burrell AS, Disotell TR. The stem catarrhine Saadanius does not inform the timing of the origin of crown catarrhines. Journal of Human Evolution. 2011;61(2):209–10. doi: 10.1016/j.jhevol.2011.02.008. [DOI] [PubMed] [Google Scholar]
  144. Puigbò P, Garcia-Vallvé S, McInerney JO. TOPD/FMTS: a new software to compare phylogenetic trees. Bioinformatics. 2007;23:1556–8. doi: 10.1093/bioinformatics/btm135. [DOI] [PubMed] [Google Scholar]
  145. Pyron RA. A likelihood method for assessing molecular divergence time estimates and the placement of fossil calibrations. Systematic Biology. 2010;59:185–94. doi: 10.1093/sysbio/syp090. [DOI] [PubMed] [Google Scholar]
  146. R Development Core Team. R: A language and environment for statistical computing. R Foundation for Statistical Computing; Vienna, Austria: 2009. URL http://www.R-project.org. [Google Scholar]
  147. Raaum RL, Sterner KN, Noviello CM, Stewart C-B, Disotell TR. Catarrhine primate divergence dates estimated from complete mitochondrial genomes: concordance with fossil and nuclear DNA evidence. Journal of Human Evolution. 2005;48:237–57. doi: 10.1016/j.jhevol.2004.11.007. [DOI] [PubMed] [Google Scholar]
  148. Raina SZ, Faith JJ, Disotell TR, Seligmann H, Stewart C, Pollock DD. Evolution of base-substitution gradients in primate mitochondrial genomes. Genome Research. 2005;15:665–73. doi: 10.1101/gr.3128605. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Rambaut A, Drummond AJ. Tracer v1.4. 2007 Available from http://beast.bio.ed.ac.uk/Tracer.
  150. Rasmussen DT. Early catarrhines of the African Eocene and Oligocene. In: Hartwig WC, editor. The Primate Fossil Record. Cambridge University Press; Cambridge, UK: 2002. pp. 203–220. [Google Scholar]
  151. Rasmussen DT. Fossil Record of the Primates from the Paleocene to the Oligocene. In: Henke W, Tattersall I, editors. Handbook of paleoanthropology. Springer; Berlin: 2007. pp. 889–920. [Google Scholar]
  152. Reyes A, Gissi C, Pesole G. Where Do Rodents Fit? Evidence from the Complete Mitochondrial Genome of Sciurus vulgaris. Molecular Biology and Evolution. 2000;17:979–983. doi: 10.1093/oxfordjournals.molbev.a026379. [DOI] [PubMed] [Google Scholar]
  153. Reyes A, Pesole G, Saccone C. Complete Mitochondrial DNA Sequence of the Fat Dormouse, Glis glis : Further Evidence of Rodent Paraphyly. Molecular Biology and Evolution. 1998;15:499–505. doi: 10.1093/oxfordjournals.molbev.a025949. [DOI] [PubMed] [Google Scholar]
  154. Roach JC, Glusman G, Smit AFA, Huff CD, Hubley R, Shannon PT, Rowen L, Pant KP, Goodman N, Bamshad M, Shendure J, Drmanac R, Jorde LB, Hood L, Galas DJ. Analysis of genetic inheritance in a family quartet by whole-genome sequencing. Science (New York, NY) 2010;328:636–9. doi: 10.1126/science.1186802. [DOI] [PMC free article] [PubMed] [Google Scholar]
  155. Robinson D, Foulds L. Comparison of phylogenetic trees. Mathematical Biosciences. 1981;141:131–141. [Google Scholar]
  156. Rogaev EI, Moliaka YK, Malyarchuk BA, Kondrashov FA, Derenko MV, Chumakov I, Grigorenko AP. Complete mitochondrial genome and phylogeny of Pleistocene mammoth Mammuthus primigenius. PLoS Biology. 2006;4:e73. doi: 10.1371/journal.pbio.0040073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Ronquist F, Huelsenbeck JP. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics. 2003;19:1572–1574. doi: 10.1093/bioinformatics/btg180. [DOI] [PubMed] [Google Scholar]
  158. Roos C, Schmitz J, Zischler H. Primate jumping genes elucidate strepsirrhine phylogeny. Proceedings of the National Academy of Sciences of the United States of America. 2004;101:10650. doi: 10.1073/pnas.0403852101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  159. Rose KD. The Beginning of the Age of Mammals. JHU Press; 2006. p. 428. [Google Scholar]
  160. Rothfels CJ, Larsson A, Kuo L-Y, Korall P, Chiou W-L, Pryer KM. Overcoming deep roots, fast rates, and short internodes to resolve the ancient rapid radiation of eupolypod II ferns. Systematic Biology. 2012;61:490–509. doi: 10.1093/sysbio/sys001. [DOI] [PubMed] [Google Scholar]
  161. Rowe KC, Singhal S, Macmanes MD, Ayroles JF, Morelli TL, Rubidge EM, Bi K, Moritz CC. Museum genomics: low-cost and high-accuracy genetic data from historical specimens. Molecular Bcology Resources. 2011;11:1082–92. doi: 10.1111/j.1755-0998.2011.03052.x. [DOI] [PubMed] [Google Scholar]
  162. Rutschmann F, Eriksson T, Salim KA, Conti E. Assessing calibration uncertainty in molecular dating: the assignment of fossils to alternative calibration points. Systematic Biology. 2007;56:591–608. doi: 10.1080/10635150701491156. [DOI] [PubMed] [Google Scholar]
  163. Sanders SJ, Murtha MT, Gupta AR, Murdoch JD, Raubeson MJ, Willsey AJ, Ercan-Sencicek AG, DiLullo NM, Parikshak NN, Stein JL, Walker MF, Ober GT, Teran NA, Song Y, El-Fishawy P, Murtha RC, Choi M, Overton JD, Bjornson RD, Carriero NJ, Meyer KA, Bilguvar K, Mane SM, Sestan N, Lifton RP, Günel M, Roeder K, Geschwind DH, Devlin B, State MW. De novo mutations revealed by whole-exome sequencing are strongly associated with autism. Nature. 2012;485:237–41. doi: 10.1038/nature10945. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Scally A, Durbin R. Revising the human mutation rate: implications for understanding human evolution. Nature reviews. Genetics. 2012;13:745–53. doi: 10.1038/nrg3295. [DOI] [PubMed] [Google Scholar]
  165. Schmitz J, Ohme M, Suryobroto B, Zischler H. The Colugo (Cynocephalus variegatus, Dermoptera): The Primates ’ Gliding Sister? Molecular Biology and Evolution. 2002a;19:2308–2312. doi: 10.1093/oxfordjournals.molbev.a004054. [DOI] [PubMed] [Google Scholar]
  166. Schmitz J, Ohme M, Zischler H. The complete mitochondrial sequence of Tarsius bancanus: evidence for an extensive nucleotide compositional plasticity of primate mitochondrial DNA. Molecular Biology and Evolution. 2002b;19:544–53. doi: 10.1093/oxfordjournals.molbev.a004110. [DOI] [PubMed] [Google Scholar]
  167. Schmitz J, Zischler H. A novel family of tRNA-derived SINEs in the colugo and two new retrotransposable markers separating dermopterans from primates. Molecular Phylogenetics and Evolution. 2003;28:341–349. doi: 10.1016/s1055-7903(03)00060-5. [DOI] [PubMed] [Google Scholar]
  168. Seiffert ER. Early evolution and biogeography of Lorisiform strepsirrhines. American Journal of Primatology. 2007;69:27–35. doi: 10.1002/ajp.20324. [DOI] [PubMed] [Google Scholar]
  169. Seiffert ER, Simons EL, Clyde WC, Rossie JB, Attia Y, Bown TM, Chatrath P, Mathison ME. Basal anthropoids from Egypt and the antiquity of Africa’s higher primate radiation. Science (New York, NY) 2005;310:300–4. doi: 10.1126/science.1116569. [DOI] [PubMed] [Google Scholar]
  170. Seiffert ER, Simons EL, Attia Y. Fossil evidence for an ancient divergence of lorises and galagos. Nature. 2003;422:421–424. doi: 10.1038/nature01489. [DOI] [PubMed] [Google Scholar]
  171. Senut B, Pickford M, Gommery D, Mein P, Cheboi K, Coppens Y. First hominid from the Miocene (Lukeino Formation, Kenya) Comptes Rendus de l’Académie des Sciences - Series IIA - Earth and Planetary Science. 2001;332:137–144. [Google Scholar]
  172. Setoguchi T, Rosenberger AL. A fossil owl monkey from La Venta, Colombia. Nature. 1987;326:692–694. doi: 10.1038/326692a0. [DOI] [PubMed] [Google Scholar]
  173. Sigé B, Jaeger J-J, Sudre J, Vianey Liaud M. Altiatlasius koulchii n. gen. et sp., primate omomyidé du Paléocène supérieur du Maroc, et les origines des euprimates. Palaeontographica. 1990;216:31–56. [Google Scholar]
  174. Silcox MT, Sargis EJ, Bloch JI, Boyer DM. Primate Origins and Supraordinal Relationships: Morphological Evidence. In: Henke W, Tattersall I, editors. Handbook of paleoanthropology. Springer; New York: 2007. pp. 831–859. [Google Scholar]
  175. Slack KE, Delsuc F, McLenachan P, Arnason U, Penny D. Resolving the root of the avian mitogenomic tree by breaking up long branches. Molecular Phylogenetics and Evolution. 2007;42:1–13. doi: 10.1016/j.ympev.2006.06.002. [DOI] [PubMed] [Google Scholar]
  176. Slater GJ. Phylogenetic evidence for a shift in the mode of mammalian body size evolution at the Cretaceous-Palaeogene boundary. Methods in Ecology and Evolution. 2013;4(8):734–744. [Google Scholar]
  177. Smith T, Rose KD, Gingerich PD. Rapid Asia-Europe-North America geographic dispersal of earliest Eocene primate Teilhardina during the Paleocene-Eocene Thermal Maximum. Proceedings of the National Academy of Sciences of the United States of America. 2006;103:11223–7. doi: 10.1073/pnas.0511296103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  178. Soligo C, Martin RD. Adaptive origins of primates revisited. Journal of Human Evolution. 2006;50:414–30. doi: 10.1016/j.jhevol.2005.11.001. [DOI] [PubMed] [Google Scholar]
  179. Solounias N, Barry JC, Bernor RL, Lindsay EH, Raza SM. The oldest bovid from the Siwaliks, Pakistan. Journal of Vertebrate Paleontology. 1995;15:806–814. [Google Scholar]
  180. Song S, Liu L, Edwards SV, Wu S. Resolving conflict in eutherian mammal phylogeny using phylogenomics and the multispecies coalescent model. Proceedings of the National Academy of Sciences of the United States of America. 2012;109:14942–7. doi: 10.1073/pnas.1211733109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Springer MS, Douzery E. Secondary Structure and Patterns of Evolution Among Mammalian Mitochondrial 12S rRNA Molecules. Journal of Molecular Evolution. 1996;43:357–373. doi: 10.1007/BF02339010. [DOI] [PubMed] [Google Scholar]
  182. Springer MS, Meredith RW, Gatesy J, Emerling C, Park J, Rabosky DL, Stadler T, Steiner C, Ryder O, Janečka JE, Fisher C, Murphy WJ. Macroevolutionary dynamics and historical biogeography of primate diversification inferred from a species supermatrix. PloS One. 2012;7:e49521. doi: 10.1371/journal.pone.0049521. [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Springer MS, Meredith RW, Teeling EC, Murphy WJ. Technical Comment on “The Placental Mammal Ancestor and the Post–K-Pg Radiation of Placentals”. Science. 2013;341(6146):613–613. doi: 10.1126/science.1238025. [DOI] [PubMed] [Google Scholar]
  184. Springer MS, Murphy WJ, Eizirik E, O’Brien SJ. Placental mammal diversification and the Cretaceous–Tertiary boundary. Proceedings of the National Academy of Sciences of the United States of America. 2003;100:1056. doi: 10.1073/pnas.0334222100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Stamatakis A, Hoover P, Rougemont J. A rapid bootstrap algorithm for the RAxML Web servers. Systematic Biology. 2008;57:758–71. doi: 10.1080/10635150802429642. [DOI] [PubMed] [Google Scholar]
  186. Stamatakis A, Ludwig T, Meier H. RAxML-III: a fast program for maximum likelihood-based inference of large phylogenetic trees. Bioinformatics. 2005;21:456–63. doi: 10.1093/bioinformatics/bti191. [DOI] [PubMed] [Google Scholar]
  187. Stamatakis A. RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics. 2006;22:2688–90. doi: 10.1093/bioinformatics/btl446. [DOI] [PubMed] [Google Scholar]
  188. Steiper ME, Seiffert ER. Evidence for a convergent slowdown in primate molecular rates and its implications for the timing of early primate evolution. Proceedings of the National Academy of Sciences of the United States of America. 2012;109:6006–11. doi: 10.1073/pnas.1119506109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  189. Steiper ME, Young NM. Timing primate evolution: Lessons from the discordance between molecular and paleontological estimates. Evolutionary Anthropology: Issues, News, and Reviews. 2008;17:179–188. [Google Scholar]
  190. Sterner KN, Raaum RL, Zhang Y-P, Stewart C-B, Disotell TR. Mitochondrial data support an odd-nosed colobine clade. Molecular Phylogenetics and Evolution. 2006;40:1–7. doi: 10.1016/j.ympev.2006.01.017. [DOI] [PubMed] [Google Scholar]
  191. Stocsits RR, Letsch H, Hertel J, Misof B, Stadler PF. Accurate and efficient reconstruction of deep phylogenies from structured RNAs. Nucleic Acids Research. 2009;37:6184–6193. doi: 10.1093/nar/gkp600. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Suchard MA, Weiss RE, Sinsheimer JS. Bayesian selection of continuous-time Markov chain evolutionary models. Molecular Biology and Evolution. 2001;18:1001–13. doi: 10.1093/oxfordjournals.molbev.a003872. [DOI] [PubMed] [Google Scholar]
  193. Suwa G, Kono RT, Katoh S, Asfaw B, Beyene Y. A new species of great ape from the late Miocene epoch in Ethiopia. Nature. 2007;448:921–924. doi: 10.1038/nature06113. [DOI] [PubMed] [Google Scholar]
  194. Talavera G, Castresana J. Improvement of phylogenies after removing divergent and ambiguously aligned blocks from protein sequence alignments. Systematic Biology. 2007;56:564–77. doi: 10.1080/10635150701472164. [DOI] [PubMed] [Google Scholar]
  195. Tavaré S, Marshall CR, Will O, Soligo C, Martin RD. Using the fossil record to estimate the age of the last common ancestor of extant primates. Nature. 2002;416:726–729. doi: 10.1038/416726a. [DOI] [PubMed] [Google Scholar]
  196. Thalmann O, Hebler J, Poinar HN, Pääbo S, Vigilant L. Unreliable mtDNA data due to nuclear insertions: a cautionary tale from analysis of humans and other great apes. Molecular Ecology. 2004;13:321–335. doi: 10.1046/j.1365-294x.2003.02070.x. [DOI] [PubMed] [Google Scholar]
  197. Thewissen JGM, Williams EM, Hussain ST. Skeletons of terrestrial cetaceans and the relationship of whales to artiodactyls. Nature. 2001;413:277–281. doi: 10.1038/35095005. [DOI] [PubMed] [Google Scholar]
  198. Thorne JL, Kishino H. Divergence time and evolutionary rate estimation with multilocus data. Systematic Biology. 2002;51:689–702. doi: 10.1080/10635150290102456. [DOI] [PubMed] [Google Scholar]
  199. Ting N, Sterner KN. Primate molecular phylogenetics in a genomic era. Molecular Phylogenetics and Evolution. 2013;66:565–568. doi: 10.1016/j.ympev.2012.08.021. [DOI] [PubMed] [Google Scholar]
  200. Townsend JP, Leuenberger C. Taxon sampling and the optimal rates of evolution for phylogenetic inference. Systematic Biology. 2011;60:358–65. doi: 10.1093/sysbio/syq097. [DOI] [PubMed] [Google Scholar]
  201. Ursing BM, Arnason U. The complete mitochondrial DNA sequence of the pig (Sus scrofa) Journal of Molecular Evolution. 1998a;47:302–306. doi: 10.1007/pl00006388. [DOI] [PubMed] [Google Scholar]
  202. Ursing BM, Arnason U. Analyses of mitochondrial genomes strongly support a hippopotamus-whale clade. Proceedings Biological sciences/The Royal Society. 1998b;265:2251–5. doi: 10.1098/rspb.1998.0567. [DOI] [PMC free article] [PubMed] [Google Scholar]
  203. Ursing BM, Slack KE, Arnason U. Subordinal artiodactyl relationships in light of phylogenetic analysis of 12 mitochondrial protein-coding genes. Zoologica Scripta. 2000;29:83–88. [Google Scholar]
  204. Vaidya G, Lohman DJ, Meier R. SequenceMatrix: concatenation software for the fast assembly of multi-gene datasets with character set and codon information. Cladistics. 2011;27:171–180. doi: 10.1111/j.1096-0031.2010.00329.x. [DOI] [PubMed] [Google Scholar]
  205. Vignaud P, Duringer P, Mackaye HT, Likius A, Blondel C, Boisserie J-R, De Bonis L, Eisenmann V, Etienne M-E, Geraads D, Guy F, Lehmann T, Lihoreau F, Lopez-Martinez N, Mourer-Chauviré C, Otero O, Rage J-C, Schuster M, Viriot L, Zazzo A, Brunet M. Geology and palaeontology of the Upper Miocene Toros-Menalla hominid locality, Chad. Nature. 2002;418:152–5. doi: 10.1038/nature00880. [DOI] [PubMed] [Google Scholar]
  206. Ward PS, Brady SG, Fisher BL, Schultz TR. Phylogeny and biogeography of dolichoderine ants: effects of data partitioning and relict taxa on historical inference. Systematic Biology. 2010;59:342–62. doi: 10.1093/sysbio/syq012. [DOI] [PubMed] [Google Scholar]
  207. Wertheim JO, Worobey M. A challenge to the ancient origin of SIVagm based on African green monkey mitochondrial genomes. PLoS Pathogens. 2007;3:e95. doi: 10.1371/journal.ppat.0030095. [DOI] [PMC free article] [PubMed] [Google Scholar]
  208. Wible J, Covert HH. Primates: cladistic diagnosis and relationships. Journal of Human Evolution. 1987;16:1–22. [Google Scholar]
  209. Wible JR, Rougier GW, Novacek MJ, Asher RJ. Cretaceous eutherians and Laurasian origin for placental mammals near the K/T boundary. Nature. 2007;447:1003–6. doi: 10.1038/nature05854. [DOI] [PubMed] [Google Scholar]
  210. Wilcox TP, Zwickl DJ, Heath TA, Hillis DM. Phylogenetic relationships of the dwarf boas and a comparison of Bayesian and bootstrap measures of phylogenetic support. Molecular phylogenetics and evolution. 2002;25:361–71. doi: 10.1016/s1055-7903(02)00244-0. [DOI] [PubMed] [Google Scholar]
  211. Wilkinson RD, Steiper ME, Soligo C, Martin RD, Yang Z, Tavaré S. Dating primate divergences through an integrated analysis of palaeontological and molecular data. Systematic Biology. 2011;60:16–31. doi: 10.1093/sysbio/syq054. [DOI] [PMC free article] [PubMed] [Google Scholar]
  212. Williams BA, Kay RF, Kirk EC. New perspectives on anthropoid origins. Proceedings of the National Academy of Sciences of the United States of America. 2010;107:4797–804. doi: 10.1073/pnas.0908320107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  213. Xu X, Arnason U. The complete mitochondrial DNA sequence of the horse, Equus caballus: extensive heteroplasmy of the control region. Gene. 1994;148:357–362. doi: 10.1016/0378-1119(94)90713-7. [DOI] [PubMed] [Google Scholar]
  214. Xu X, Arnason U. The Mitochondrial DNA Molecule of Sumatran Orangutan and a Molecular Proposal for Two (Bornean and Sumatran) Species of Orangutan. Journal of Molecular Evolution. 1996;43:431–437. doi: 10.1007/BF02337514. [DOI] [PubMed] [Google Scholar]
  215. Xu X, Janke A, Arnason U. The complete mitochondrial DNA sequence of the greater Indian rhinoceros, Rhinoceros unicornis, and the Phylogenetic relationship among Carnivora. Molecular Biology and Evolution. 1996;13:1167–1173. doi: 10.1093/oxfordjournals.molbev.a025681. [DOI] [PubMed] [Google Scholar]
  216. Yang Z. PAML: a program package for phylogenetic analysis by maximum likelihood. Computer applications in the biosciences: CABIOS. 1997;13:555–6. doi: 10.1093/bioinformatics/13.5.555. [DOI] [PubMed] [Google Scholar]
  217. Yang Z. PAML 4: phylogenetic analysis by maximum likelihood. Molecular Biology and Evolution. 2007;24:1586–91. doi: 10.1093/molbev/msm088. [DOI] [PubMed] [Google Scholar]
  218. Yoder AD. Fossils versus clocks. Science. 2013;339:656. doi: 10.1126/science.1233999. [DOI] [PubMed] [Google Scholar]
  219. Zalmout IS, Sanders WJ, Maclatchy LM, Gunnell GF, Al-Mufarreh YA, Ali MA, Nasser A-AH, Al-Masari AM, Al-Sobhi SA, Nadhra AO, Matari AH, Wilson JA, Gingerich PD. New Oligocene primate from Saudi Arabia and the divergence of apes and Old World monkeys. Nature. 2010;466:360–4. doi: 10.1038/nature09094. [DOI] [PubMed] [Google Scholar]
  220. Zardoya R, Malaga-Trillo E, Veith M, Meyer A. Complete nucleotide sequence of the mitochondrial genome of a salamander, Mertensiella luschani. Gene. 2003;317:17–27. doi: 10.1016/s0378-1119(03)00655-3. [DOI] [PubMed] [Google Scholar]
  221. Zhang P, Wake DB. Higher-level salamander relationships and divergence dates inferred from complete mitochondrial genomes. Molecular phylogenetics and Evolution. 2009;53:492–508. doi: 10.1016/j.ympev.2009.07.010. [DOI] [PubMed] [Google Scholar]
  222. Zhang P, Zhou H, Liang D, Liu Y-F, Chen Y-Q, Qu L-H. The complete mitochondrial genome of a tree frog, Polypedates megacephalus (Amphibia: Anura: Rhacophoridae), and a novel gene organization in living amphibians. Gene. 2005;346:133–43. doi: 10.1016/j.gene.2004.10.012. [DOI] [PubMed] [Google Scholar]
  223. Zinner D, Wertheimer J, Liedigk R, Groeneveld LF, Roos C. Baboon phylogeny as inferred from complete mitochondrial genomes. American Journal of Physical Anthropology. 2013;150(1):133–40. doi: 10.1002/ajpa.22185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  224. Zischler H. Molecular evidence on primate origins and evolution. In: Henke W, Tattersall I, editors. Handbook of paleoanthropology. Springer; New York: 2007. pp. 861–888. [Google Scholar]
  225. Zwickl DJ, Hillis DM. Increased taxon sampling greatly reduces phylogenetic error. Systematic Biology. 2002;51:588–98. doi: 10.1080/10635150290102339. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

01

Supplementary Material Table A1. Details about the samples used in this study.

02. Supplementary Material Table B1. List of the fossils used as calibration points in this study.

For each node, reference fossil, type specimens, dates, and references are reported.

03

Supplementary Material Table C1. List of bootstrap values (maximum likelihood), posterior probabilities (MrBayes), mean divergence dates, and 95% composite credibility intervals in millions of years for all nodes within primates. Nodes are reported in Figure 2.

04. Supplementary Material Figure D1–2. Phylogenetic tree inferred from the 12 protein coding genes only.

1) Maximum likelihood analyses performed using RAxML 7.2.6. Bootstrap values (PP) are shown for each node. 2) Bayesian analysis obtained using MrBayes 3.2.1. Posterior probability (PP) values are shown for each node.

05. Supplementary Material Figure E1–3. Test for compositional biases.

Maximum likelihood trees with bootstrap values (the bootstopping criteria autoMRE (Majority Rule Criterion) as implemented in RAxML 7.2.6 (Pattengale et al., 2009) for three datasets: 1) third codon position in protein coding regions coded as RY [3RY], 2) positions with Cs and Ts replaced with Y [AGY], and 3) all nucleotides coded as purines (R) and pyrimidines (Y) [allRY].

06. Supplementary Material Figure F1–4. Test for taxon sampling effects.

Maximum likelihood trees with bootstrap values (the bootstopping criteria autoMRE [Majority Rule Criterion] as implemented in RAxML 7.2.6 [Pattengale et al., 2009]) for four datasets: 1) only six primate taxa (one for each main lineage within primates: Lemuriformes, Chyromyiformes, Lorisiformes, Tarsiiformes, Platyrrhini, Catarrhini), 2) ten taxa as used by Schmitz et al. (2002a), 3) 14 taxa as used by Arnason et al. (2008), and 4) 26 taxa as in Matsui et al. (2009).

07
08

RESOURCES