Abstract
In phylogenetic inference, we commonly use models of substitution which assume that sequence evolution is stationary, reversible, and homogeneous (SRH). Although the use of such models is often criticized, the extent of SRH violations and their effects on phylogenetic inference of tree topologies and edge lengths are not well understood. Here, we introduce and apply the maximal matched-pairs tests of homogeneity to assess the scale and impact of SRH model violations on 3,572 partitions from 35 published phylogenetic data sets. We show that roughly one-quarter of all the partitions we analyzed (23.5%) reject the SRH assumptions, and that for 25% of data sets, tree topologies inferred from all partitions differ significantly from topologies inferred using the subset of partitions that do not reject the SRH assumptions. This proportion increases when comparing trees inferred using the subset of partitions that rejects the SRH assumptions, to those inferred from partitions that do not reject the SRH assumptions. These results suggest that the extent and effects of model violation in phylogenetics may be substantial. They highlight the importance of testing for model violations and possibly excluding partitions that violate models prior to tree reconstruction. Our results also suggest that further effort in developing models that do not require SRH assumptions could lead to large improvements in the accuracy of phylogenomic inference. The scripts necessary to perform the analysis are available in https://github.com/roblanf/SRHtests, and the new tests we describe are available as a new option in IQ-TREE (http://www.iqtree.org).
Keywords: model violations, phylogenetic inference, test of symmetry, systematic bias
Introduction
Phylogenetics is an essential tool for inferring evolutionary relationships between individuals, species, genes, and genomes. Moreover, phylogenetic trees form the basis of a huge range of other inferences in evolutionary biology, from gene function prediction to drug development and forensics (Eisen 1998; Farrell et al. 2000; Mäser et al. 2001; Gardner et al. 2002; Yao et al. 2003, 2004; Grenfell et al. 2004; Salipante and Horwitz 2006; Gray et al. 2009; Brady and Salzberg 2011; Dunn et al. 2011).
Most phylogenetic studies use models of sequence evolution which assume that the evolutionary process follows stationary, reversible, and homogeneous (SRH) conditions. Stationarity implies that the marginal frequencies of the nucleotides or amino acids are constant over time, reversibility implies that the evolutionary process is stationary and undirected (substitution rates between nucleotides or amino acids are equal in both directions), and homogeneity implies that the instantaneous substitution rates are constant along the tree or over an edge (Felsenstein 2004; Yang and Rannala 2012; Jermiin et al. 2017). However, these simplifying assumptions are often violated by real data (Foster and Hickey 1999; Tarrío et al. 2001; Paton et al. 2002; Goremykin and Hellwig 2005; Murray et al. 2005; Bourlat et al. 2006; Hyman et al. 2007; Sheffield et al. 2009; Nesnidal et al. 2010; Nabholz et al. 2011; Martijn et al. 2018). Such model violation may lead to systematic error that, unlike stochastic error, cannot be remedied simply by increasing the size of a data set (Felsenstein 2004; Ho and Jermiin 2004; Jermiin et al. 2004; Philippe et al. 2005; Sullivan and Joyce 2005; Kumar et al. 2012; Brown and Thomson 2017; Duchene et al. 2017). As phylogenetic data sets are steadily growing in terms of taxonomic and site sampling, it is vital that we develop and employ methods to measure and understand the extent to which systematic error affects phylogenetic inference (systematic bias), and explore ways of mitigating this systematic bias in empirical studies.
One approach to accommodate data that have evolved under non-SRH conditions is to employ models that relax the SRH assumptions. A number of non-SRH models have been implemented in a variety of software packages (Foster 2004; Lartillot and Philippe 2004; Blanquart and Lartillot 2006; Boussau and Gouy 2006; Jayaswal et al. 2007, 2011, 2014; Knight et al. 2007; Dutheil and Boussau 2008; Sumner et al. 2012; Zou et al. 2012; Groussin et al. 2013; Nguyen et al. 2015; Woodhams et al. 2015). However, such models remain infrequently used as searching for optimal phylogenetic trees under these models is computationally demanding (Betancur-r et al. 2013) and the implementations are often not easy to use. As a result, the vast majority of empirical phylogenetic inferences rely on models that assume sequences have evolved under SRH conditions, such as the general time reversible family of models implemented in many of the most widely used phylogenetics software packages (Swofford 2001; Drummond and Rambaut 2007; Guindon et al. 2010; Ronquist et al. 2012; Bazinet et al. 2014; Bouckaert et al. 2014; Stamatakis 2014; Nguyen et al. 2015; Höhna et al. 2016).
Another approach to accounting for data that may have evolved under non-SRH conditions is to test for model violations prior to tree reconstruction. Here, one first screens data sets or parts of data sets, and reconstructs trees exclusively from data that do not reject SRH conditions. A number of methods have been proposed to test for violation of SRH conditions in aligned sequences prior to estimating trees (Bowker 1948; Stuart 1955; Rzhetsky and Nei 1995; Kumar and Gadagkar 2001; Weiss and von Haeseler 2003; Ababneh et al. 2006; Ho et al. 2006), and there are also a posteriori tests for absolute model adequacy which are employed after trees have been estimated (Goldman 1993; Bollback 2002; Brown and ElDabaje 2009; Brown 2014; Duchene et al. 2017; Brown and Thomson 2018).
Allowing the data to reject the model when the assumptions of the model are violated is an important approach to reducing systematic bias in phylogenetic inference (Philippe et al. 2005; Brown 2014). Knowing in advance which sequences and loci are inconsistent with the SRH assumptions will allow us to choose more complex models or to omit some of these sequences and loci from downstream analyses (Kumar and Gadagkar 2001). The need for methods that assess the evolutionary process prior to phylogenetic inference becomes more important as the number of sequences and sites per data set increases, because systematic bias has an increasing effect on inferences from larger phylogenetic data sets (Ho and Jermiin 2004; Jermiin et al. 2004; Phillips et al. 2004; Delsuc et al. 2005).
In this article, we evaluate the extent and effect of model violation due to non-SRH evolution using 35 empirical data sets with a total of 3,572 partitions. We determine if the SRH assumptions are violated by extending and applying the matched-pairs tests of homogeneity (Jermiin et al. 2017) to each partition. We then compare the phylogenetic trees for each data set estimated from all of the partitions, the partitions that reject the SRH assumptions, and the partitions that do not reject the SRH assumptions, in order to evaluate the effect violating SRH conditions on phylogenetic inference. Our results suggest that violating SRH assumptions can have substantial impacts on phylogenetic inference.
Materials and Methods
Empirical Data Sets
In order to assess the impact of model violation in phylogenetics, we first gathered a representative sample of 35 partitioned empirical data sets that had been used for phylogenetic analysis in recent studies (table 1). Within the constraints of selecting data that were publicly available and suitably annotated, that is, such that all loci and all codon positions within protein-coding loci could be identified, we selected the data sets to provide as representative a sample as possible of the data types, taxa, and genomic regions most commonly used to infer bifurcating phylogenetic trees from concatenated alignments. These data sets include nucleotide sequences from nuclear, mitochondrial, plastid, and virus genomes, and include protein-coding DNA, introns, intergenic spacers, tRNA, rRNA, and ultraconserved elements. The number of taxa and sites in these data sets range from 27 to 355 and from 699 to 1,079,052, respectively. The clades represented in these data sets include animals, plants, and viruses. We partitioned all data sets to the maximum possible extent based on the biological properties of the data, that is, we divided every locus and every codon position within each protein-coding locus into a separate partition. All partitioning information is available at the github repository (https://github.com/roblanf/SRHtests/tree/master/datasets), and the full details of every data set are provided in table 1 and in supplementary extended table 5, Supplementary Material online.
Table 1.
Data Set | Study References | Data Set References | Clade | Taxa | Sites |
---|---|---|---|---|---|
Anderson_2013 | Anderson et al. (2014) | Anderson et al. (2013) | Loliginids | 145 | 3,037 |
Bergsten_2013 | Bergsten et al. (2013) | Bergsten et al. (2013) | Dytiscidae | 38 | 2,111 |
Broughton_2013 | Broughton et al. (2013) | Broughton et al. (2013) | Osteichthyes | 61 | 19,997 |
Brown_2012 | Brown et al. (2012) | Brown et al. (2012) | Ptychozoon | 41 | 1,665 |
Cannon_2016a | Cannon et al. (2016) | Cannon et al. (2016) | Metazoa | 78 | 89,792 |
Cognato_2001 | Cognato and Vogler (2001) | Cognato and Vogler (2001) | Coleoptera: Scolytinae | 44 | 1,897 |
Day_2013 | Day et al. (2013) | Day et al. (2013) | Synodontis | 152 | 3,586 |
Devitt_2013 | Devitt et al. (2013) | Devitt et al. (2013) | Ensatina eschscholtzii klauberi | 69 | 823 |
Dornburg_2012 | Dornburg et al. (2012) | Dornburg et al. (2012) | Teleostei: Beryciformes: Holocentridae | 44 | 5,919 |
Faircloth_2013 | Faircloth et al. (2013) | Faircloth et al. (2013) | Actinopterygii | 27 | 149,366 |
Fong_2012 | Fong et al. (2012) | Fong et al. (2012) | Vertebrata | 110 | 25,919 |
Horn_2014 | Horn et al. (2014) | Horn et al. (2014) | Euphorbia | 197 | 11,587 |
Kawahara_2013 | Kawahara and Rubinoff (2013) | Kawahara and Rubinoff (2013) | Hyposmocoma | 70 | 2,238 |
Lartillot_2012 | Lartillot and Delsuc (2012) | Lartillot and Delsuc (2012) | Eutheria | 78 | 15,117 |
McCormack_2013 | McCormack et al. (2013) | McCormack et al. (2013) | Neoaves | 33 | 1,079,052 |
Moyle_2016 | Moyle et al. (2016) | Moyle et al. (2016) | Oscines | 106 | 375,172 |
Murray_2013 | Murray et al. (2013) | Murray et al. (2013) | Eucharitidae | 237 | 3,111 |
Oaks_2011 | Oaks (2011) | Oaks (2011) | Crocodylia | 79 | 7,282 |
Rightmyer_2013 | Rightmyer et al. (2013) | Rightmyer et al. (2013) | Hymenoptera: Megachilidae | 94 | 3,692 |
Sauquet_2011 | Sauquet et al. (2012) | Sauquet et al. (2011) | Nothofagus | 51 | 5,444 |
Seago_2011 | Seago et al. (2011) | Seago et al. (2011) | Coccinellidae | 97 | 2,253 |
Sharanowski_2011 | Sharanowski et al. (2011) | Sharanowski et al. (2011) | Braconidae | 139 | 3,982 |
Siler_2013 | Siler et al. (2013) | Siler et al. (2013) | Lycodon | 61 | 2,697 |
Tolley_2013 | Tolley et al. (2013) | Tolley et al. (2013) | Chamaeleonidae | 203 | 5,054 |
Unmack_2013 | Unmack et al. (2013) | Unmack et al. (2013) | Melanotaeniidae | 139 | 6,827 |
Wainwright_2012 | Wainwright et al. (2012) | Wainwright et al. (2012) | Acanthomorpha | 188 | 8,439 |
Wood_2012 | Wood et al. (2013) | Wood et al. (2012) | Archaeidae | 37 | 5,185 |
Worobey_2014a | Worobey et al. (2014) | Worobey et al. (2014) | Influenzavirus A | 146 | 3,432 |
Worobey_2014b | Worobey et al. (2014) | Worobey et al. (2014) | Influenzavirus A | 327 | 759 |
Worobey_2014c | Worobey et al. (2014) | Worobey et al. (2014) | Influenzavirus A | 92 | 1,416 |
Worobey_2014d | Worobey et al. (2014) | Worobey et al. (2014) | Influenzavirus A | 355 | 1,497 |
Worobey_2014e | Worobey et al. (2014) | Worobey et al. (2014) | Influenzavirus A | 340 | 699 |
Worobey_2014f | Worobey et al. (2014) | Worobey et al. (2014) | Influenzavirus A | 332 | 2,151 |
Worobey_2014g | Worobey et al. (2014) | Worobey et al. (2014) | Influenzavirus A | 326 | 2,274 |
Worobey_2014h | Worobey et al. (2014) | Worobey et al. (2014) | Influenzavirus A | 351 | 2,280 |
Workflow Summary
Figure 1 outlines the workflow. For each partition in each data set, we used a new approach based on the three matched-pairs tests of homogeneity to ask whether the evolution of the aligned sequences in the partition rejects the SRH assumptions. The three matched-pairs tests of homogeneity, described in more detail below, test three slightly different assumptions about the historical process that generated each aligned pair of sequences in a given partition. A significant result from any test suggests that the nature of the evolutionary process required to explain the aligned sequences violates at least one of the three SRH conditions (Jermiin et al. 2017). For each test, we classify each partition as pass if the result of the test is nonsignificant or fail if the result of the test is significant. We then denote the original data set as Dall, while the concatenation of pass partitions is denoted Dpass and the concatenation of fail partitions as Dfail (fig. 1).
To investigate the impact of model violation on phylogenetic inference, we infer and compare three phylogenetic trees, Tall, Tpass, and Tfail, estimated from Dall, Dpass, and Dfail, respectively.
Matched-Pairs Tests of Homogeneity
The three matched-pairs tests of homogeneity that are applied to pairs of sequences are: the MPTS (matched-pairs test of symmetry), MPTMS (matched-pairs test of marginal symmetry), and MPTIS (matched-pairs test of internal symmetry). The statistics are computed on an m-by-m (m is 4 for nucleotides and 20 for amino acids) divergence matrix with elements , where is the number of alignment sites having nucleotide (or amino acid) in the first sequence and nucleotide (or amino acid) in the second sequence.
The MPTS tests the symmetry of by computing the Bowker’s (1948) test statistic as the χ2 distance between and its transpose:
where . A P value is then obtained by a χ2 test with degrees of freedom, where is the number of pairs for which . A small P value (e.g., <0.05) indicates that the assumption of symmetry is rejected at that significance level, suggesting that evolution is nonstationary, nonhomogeneous, or both (Jermiin et al. 2017).
The MPTMS tests the equality of nucleotide or amino acid composition between two sequences. To do so, MPTMS computes the Stuart’s test statistic using the difference between nucleotide or amino acid frequencies of two sequences, , and its variance–covariance matrix, . In detail, is given by where is the sum of over j, is the sum of over i, and, k = m−1. , the estimated variance–covariance matrix of u under the assumption of marginal symmetry, is defined elementwise by:
A P value is obtained by a χ2 test with m−1 degrees of freedom. A small P value (<0.05) indicates that the stationarity assumption is rejected. Note that when is not invertible, the Stuart’s statistic is ill-defined and the MPTMS is not applicable.
The MPTIS uses the test statistic as the difference between Bowker’s and Stuart’s statistic:
. is χ2 distributed with degrees of freedom. A small P value (<0.05) indicates that the homogeneity assumption is rejected.
The MPTS, MPTMS, and MPTIS test different aspects of the symmetry with which differences accumulate between pairs of sequences due to the substitution process. The MPTS is a comprehensive and sufficient test to determine whether the data comply with the SRH assumptions (Jermiin et al. 2017), but it cannot provide any information about the source of this violation. Some information on the underlying source of model violation may be obtained by performing the other two tests of symmetry: the MPTMS and the MPTIS. If the violation of the SRH assumptions stems from differences in base composition between the sequences, this should affect the marginal symmetry of the sequence pair, which can in principle be detected by the MPTMS. If the violation of the SRH assumptions stems from changes in the relative substitution rates over time, this should affect the internal symmetry of the sequence pair, which can in principle be detected by the MPTIS. However, even after performing all three tests, it is difficult to ascertain which of the three SRH assumptions is violated during the evolutionary process because the relationships between the SRH conditions and the three matched-pair tests is neither bijective nor injective, that is, there is not a one-to-one correspondence between the three tests and violation of the three SRH conditions (Jermiin et al. 2017).
The three matched-pairs tests of homogeneity are appropriate to test for SRH assumptions as they consider the alignment on a site-by-site basis. The basic intuition that underlies these tests is that two sequences diverging under SRH conditions should accumulate differences symmetrically (e.g., both sequences are equally likely to accumulate at a C to T change at a site in which both originally shared a C). This symmetry of accumulation is reflected by symmetries in the resulting difference matrix, violations of which can be assessed statistically. However, these tests were designed to ask whether any single pair of sequences rejects the SRH conditions (Jermiin et al. 2017). To ask whether a given partition rejects SRH conditions, we developed an approach to extend the matched-pairs tests of homogeneity to accommodate data sets with more than two sequences.
Maximum Symmetry Test
In order to determine whether a given multiple sequence alignment rejects SRH conditions, we consider only the pair of taxa with the maximum divergence. In order to find the maximum divergent pair, we sum the off-diagonal elements of the divergence matrix and divide by the sum of all elements. We then randomly choose one pair from all the pairs with the maximum divergence score (if there is more than one pair). By using the most divergent sequence pair, we maximize our power to detect model violations without a priori knowledge of the underlying tree topology and the dependencies that it induces in the data. For the maximum divergent pair, we then apply the matched-pair tests of homogeneity and calculate their χ2P values. If the obtained P value is <0.05, then we consider that the null hypothesis of SRH evolution is rejected for the corresponding partition and we add it to the Dfail data set. Otherwise, we add it to the Dpass data set. We denote our applications of the MPTS, MPTMS, and MPTIS based on the as MaxSymTest, MaxSymTestmar, and MaxSymTestint, respectively.
Phylogenetic Inference
We used IQ-TREE (Nguyen et al. 2015) to infer up to seven phylogenetic trees for every data set: Tall (all partitions from the original data set; Dall); and Tpass and Tfail based on the Dpass and Dfail data sets from each of the three tests (MaxSymTest, MaxSymTestmar, MaxSymTestint), provided that there was at least one partition in each category. We ran IQ-TREE using the default settings with the best-fit fully partitioned model (Chernomor et al. 2016), which allows each partition to have its own evolutionary model and edge-linked rate determined by ModelFinder (Kalyaanamoorthy et al. 2017) followed 1,000 ultrafast bootstrap replicates (Hoang et al. 2018).
Distance between Trees
For each of the three tests (MPTS, MPTMS, MPTIS) we calculated the Normalized Path-Difference (NPD) and quartet distance (QD) (Steel and Penny 1993; Sand et al. 2014) between all three possible pairs of trees (Tall vs. Tpass; Tall vs. Tfail; and Tpass vs. Tfail), as long as Dpass and Dfail were nonempty and so Tpass and Tfail had been estimated. The path-difference metric (PD) is defined as the Euclidean distance between pairs of taxa (Steel and Penny 1993; Mir and Russello 2010). In this study, because we are interested only in differences between topologies, we use the variant of the PD metric that ignores branch lengths. In order to compare path distances between trees with different number of taxa, we normalized PD (to obtain NPD) by the mean of a null distribution of PDs generated from 10K random pairs of trees with the same number of taxa (Bogdanowicz et al. 2012). Thus, an NPD of 0 indicates an identical pair of trees, an NPD of 1 indicates that a pair of trees is as similar as a pair of randomly selected trees with the same number of taxa; and an NPD >1 indicates a pair of trees that are less similar than a randomly selected pair of trees with the same number of taxa. Since path differences are always nonnegative, the NPD is also guaranteed to be nonnegative.
The QD metric is defined as the fraction of quartets (subsets of four taxa) that induce different subtrees between the two trees being compared. QD ranges between 0 and 1, where 0 means that two trees are identical and 1 means that they do not share any quartet subtrees. Compared with PD, QD has the advantage that its distribution is less sensitive to the underlying distribution of tree topologies (Steel and Penny 1993).
Tree Topology Tests
The NPD and the QD give us measures of the differences between pairs of trees, but they do not tell us whether the differences are phylogenetically significant in the three data sets (Dpass, Dall, and Dfail) derived from a given test. For example, trees that differ due to stochastic error associated with small data sets may be very different, but such differences may not be statistically significant. To assess the significance of the differences between Tpass, Tall, and Tfail, we used the weighted Shimodaira–Hasegawa (wSH) test (Shimodaira and Hasegawa 1999; Shimodaira 2002) implemented in IQ-TREE with 1,000 RELL replicates (Kishino et al. 1990). Given the alignment (Dpass), the wSH test computes a P value for each tree, where a small P value (<0.05) implies that the corresponding tree has a significantly worse likelihood than the best tree in the set of Tpass, Tall, and Tfail. We use Dpass for these tests because it is, by definition, the only data set that does not reject the underlying assumptions of the SH test. As such, we only compute sWH P values when Dpass is nonempty. Thus, we performed a wSH test for each of the three MaxSymTest variants: each of which asks whether Tall and/or Tfail can be rejected in favor of Tpass.
Correlation between Number of Substitutions and Model Violation
We hypothesized that partitions with more substitutions may be more likely to violate the SRH assumptions, since substitutions form the raw data for the matched-pairs tests of homogeneity. To assess this, we fitted a linear mixed-effects model for each of the three tests using the glmer function from the lme4 package in R (Bates et al. 2015). In this model, we treat each partition as a datapoint, the number of substitutions measured for that partition as a fixed effect, and the data set from which that partition was taken as a random effect. This allows us to estimate the extent to which the number of substitutions in a partition associates with whether a partition fails a given test of symmetry, after accounting for differences between the data sets. To calculate the R2 value, we use the r.squaredGLMM function from the MuMIn package in R (Barton 2009; Nakagawa and Schielzeth 2013).
Software Implementation
We implemented a new option –symtest in IQ-TREE to perform the three MaxSymTest matched-pairs tests of symmetry. In addition, the option –symtest-remove-bad allows users to remove from the final analysis partitions that fail the MaxSymTest. One can change the removal criterion to MaxSymTestmar or MaxSymTestint via the –symtest-type MAR|INT option. In addition, the cutoff P value can be changed using the –symtest-pval NUM option, where the default value is 0.05.
Reproducibility
The GitHub repository (https://github.com/roblanf/SRHtests) contains the raw data and Python and R scripts necessary to perform all analyses reported in this study.
Results
Violation of SRH Conditions Is Common across 35 Empirical Data Sets
Across all 3,572 partitions analyzed, 573 (16.0%) failed the MaxSymTest, 728 (20.4%) failed the MaxSymTestmar, and 312 (2.8%) failed the MaxSymTestint. In total, 840 (23.5%) of the partitions failed at least one test.
The proportion of partitions failing each test varied substantially among data sets (fig. 2), but on an average, 21.8% of the partitions in each data set failed the MaxSymTest, 27.5% failed the MaxSymTestmar, and 5.1% failed the MaxSymTestint.
The fraction of failing partitions also varied with the genome type (e.g., mitochondrial, chloroplast, or nuclear) and context (e.g., protein-coding, UCE, tRNA) from which the partition was sequenced (table 2) although we note that a substantial proportion of the partitions from almost every category failed at least one of the tests (table 2).
Table 2.
Type/Genome | Nuclear | Mitochondrial | Plastid | Virus |
---|---|---|---|---|
First codon positions | 20.2% | 27.6% | 33.3% | 25.0% |
Second codon positions | 21.0% | 7.4% | 0.0% | 25.0% |
Third codon positions | 76.6% | 44.8% | 0.0% | 75.0% |
Other (e.g., intron) | 27.8% | 100.0% | 0.0% | |
rRNA | 30.0% | 25.0% | ||
UCE | 22.5% | |||
tRNA | 0.0% |
There were no clear differences in the substitution models that were selected for the partitions that pass or fail the tests (see supplementary extended tables 1–3, Supplementary Material online). However, we note that the two most-frequently selected substitution models (for 35% of the partitions) were relatively simple: K80 (Kimura 1980) and HKY (Hasegawa et al. 1985).
Model Violation Has a Large Influence on Tree Topologies
Using both MaxSymTest and MaxSymTestmar, we compared each tree inferred from each data set (Tall) to the corresponding trees estimated from the failed (Tfail) and passed (Tpass) partitions. Disturbingly, for each of the two tree distance metrics that we considered (NPD and QD), we find that the tree inferred from the original data set tended to be more similar to the tree estimated from the failed partitions (table 3 and supplementary extended table 4, Supplementary Material online). Furthermore, the mean NPD distance between Tpass and Tfail across all 35 data sets for the MaxSymTest was 0.69, that is, the two trees are 69% as dissimilar as random pairs of trees. This suggests that violations of SRH assumptions drive large changes in tree topologies.
Table 3.
T fail | T pass | |
---|---|---|
MaxSymTest | ||
Tall | 14.3% (4.8%) | 4.8% (4.8%) |
Tpass | 80.9% (90.4%) | |
MaxSymTestmar | ||
Tall | 8.3% (0.0%) | 8.3% (4.2%) |
Tpass | 83.4% (95.8%) | |
MaxSymTestint | ||
Tall | 28.6% (28.6%) | 0.0% (0.0%) |
Tpass | 71.4% (71.4%) |
The results of the wSH tests (table 4) confirm that the differences between trees that we observe tend to be statistically significant. For example, when using the MaxSymTestmar, Tpass is a significantly better description of the Dpass data than Tall in ∼37% of the data sets, and better than Tfail in ∼89% of the data sets.
Table 4.
T all | T fail | |
---|---|---|
MaxSymTest | 25% | 79% |
MaxSymTestmar | 37% | 89% |
MaxSymTestint | 4% | 28% |
The Number of Substitutions Explains Less than One-Third of the Variance in Passing or Failing the Tests of Symmetry
The number of substitutions in a partition explained 27.5% of the variation in whether or not a partition passed or failed the MaxSymTest (supplementary extended fig. 7, Supplementary Material online). This proportion is very similar for MaxSymTestmar (24.4%) (supplementary extended fig. 8, Supplementary Material online), but is dramatically lower for the MaxSymTestint (1.8%) (supplementary extended fig. 9, Supplementary Material online). Thus, although the number of substitutions in a partition is a highly significant (P < 2e-16) predictor of passing or failing any of the tests, that it explains only about a quarter of the variation suggests that other factors, such as underlying differences in the extent to which partitions violate the SRH assumptions, are driving the remaining ∼75% of the variation.
Model Violation Due to Non-SRH Evolution Affects the Inferred Relationship between Even-Toed and Odd-Toed Ungulates in the Tree of Mammals
To examine the effects of model violation in more detail, we selected two data sets for more detailed consideration. Conflicting support for the placement of Xenacoelomorpha, the clade that contains Xenoturbella and Acoelomorpha, in the tree of life across different analyses has led to various hypotheses about the evolution of Bilateria (Cannon et al. 2016). In addition, the interordinal relationships in Laurasiatheria, especially the relationships between Fereuungulata (Perissodactyla, Cetartiodactyla, Carnivora, and Pholidota), in the tree of placental mammals is controversial (Cao et al. 1998; Zhou et al. 2012). It has been suggested that such inferences might be strongly affected by model violation and systematic error (Cao et al. 1998; Delsuc et al. 2005; Philippe et al. 2011; Tsagkogeorga et al. 2013). To assess whether data that pass or fail the MaxSymTestmar show different signals regarding the evolution of the Bilateria and the superorder Laurasiatheria, we examined in more detail the Tall, Tpass, and Tfail trees from recent studies that explored the tree of placental mammals (Lartillot and Delsuc 2012) and the tree of all animals (Cannon et al. 2016). The mammals’ data set comprises 78 mammalian taxa, including 73 placental mammals with 51 partitions representing the first, second, and third codon positions of the 17 genes (Lartillot and Delsuc 2012). The tree reconstructed from all of the partitions (Tall) and the tree reconstructed from the partitions that pass the MaxSymTest (Tpass, 29 partitions) both show Perissodactyla (odd-toed ungulates) as a sister group to Cetartiodactyla (even-toed ungulates) (fig. 3a and supplementary extended figs. 4 and 5, Supplementary Material online). Even so, the bootstrap support for this branch is not high: 73% for Tall and 34% for Tpass. On the other hand, the tree reconstructed from the data that fail the MaxSymTest (Tfail, 22 partitions) shows Perissodactyla as the sister group to the clade that contains Carnivora + Pholidota with 49% bootstrap support (fig. 3b and supplementary extended fig. 6, Supplementary Material online).
The animal data set comprises 76 metazoan taxa, 2 choanoflagellate outgroups, 212 genes, and 424 partitions representing first and second codon positions (Cannon et al. 2016). The tree reconstructed from all of the partitions (Tall) is identical to the trees reconstructed from the 381 partitions that pass the MaxSymTest (Tpass), the partitions that fail the MaxSymTest (Tpass, 43 partitions), and the tree shown in the original paper from both DNA and amino acid data (Cannon et al. 2016), which places Xenacoelomorpha as the sister group of Nephrozoa (Deuterostomia and Protostomia) with 100% bootstrap support (supplementary extended figs. 1–3, Supplementary Material online).
Discussion
In this article, we show that model violation is prevalent and has a strong impact on tree reconstruction in many phylogenetic data sets. This impact varies substantially between different data sets and different types of partitions. The trees inferred from different groups of partitions from the same data set often have topologies that are biologically and statistically significantly different.
Our results show great heterogeneity in the extent of model violation among different data sets and partitions. This is demonstrated by the varying proportion of partitions that failed the matched-pairs tests of homogeneity in each data set and in each genomic context (codon position, rRNA, tRNA, UCE, or other) and type of genome (nuclear, mitochondrial, plastid, and virus). Model violations are most frequently observed in the third codon positions for viral, mitochondrial and nuclear genomes, and intergenic spacers in plastid sequences. Yet, our results affirm that non-SRH evolution is far from constrained to these genomic regions. For example, in a data set of placental mammals, of the 22 partitions that failed the MaxSymTest, only 11 are third codon positions. The tree inferred from the partitions that show significant violation of the SRH conditions (Tfail) differs in its topology from the tree inferred from the partitions that do not show significant violation of the SRH conditions (Tpass) with respect to the interordinal relationships in Laurasiatheria (fig. 3). The tree inferred from partitions that violate the SRH conditions (Tfail) is consistent with the results from the original paper in that it places Perissodactyla as a sister group to Carnivora + Pholidota (Lartillot and Delsuc 2012). However, other studies using ML analysis show Perissodactyla to be a sister group to Cetartiodactyla (Graur et al. 1997; Murphy et al. 2001; Tsagkogeorga et al. 2013; Liu et al. 2017), which is also the relationship we find in this study with the tree inferred from partitions that do not show significant violation of the SRH assumptions.
Examining the results of the two other tests (MaxSymTestmar and MaxSymTestint) we noticed that all the partitions that failed the MaxSymTest also failed the MaxSymTestmar, suggesting that those partitions are violating the models mainly due to nonstationarity. Based on this observation, GC content may drive the differences between the trees inferred from all partitions and those inferred from partitions that failed neither MaxSymTest nor MaxSymTestmar. Trees with partitions that violate the models tend to group together clades with similar GC content (e.g., as in Betancur-r et al. 2013). However, it is hard to discern any clear evidence for this from examining the GC content of the clades (fig. 3). Yet, our results show that all the clades in the partitions that failed the MaxSymTest have on an average a higher GC content (fig. 3).
The results of our study also provide some insight into the likely cause of model violation in the data sets we examined. Figure 2 shows that violation of marginal symmetry (assessed with MaxSymTestmar) was much more common than violation of internal symmetry (assessed with MaxSymTestint). This suggests that nonstationarity, which is associated with marginal symmetry, is likely a more common cause of systematic bias than nonhomogeneity in the data sets that we examined (see also Jayaswal et al. 2005; Ababneh et al. 2006; Song et al. 2010). Yet, the difference between the proportion of partitions that failed the MaxSymTestmar and the proportion of partitions that failed the MaxSymTestint could also be due to the higher power of the MaxSymTestmar. Either way, this result hints that the development and application of nonstationary models (Yang 1994; Roberts and Yang 1995; Yap and Speed 2005) may be an important avenue toward reducing systematic bias in future analyses. Moreover, our results show a clear preference for simple substitution models with a single transition/transversion ratio over more complex models such as general time reversible. This suggests that developing nonstationary models with a single parameter for the transition/transversion ratio might be sufficient to reduce systematic bias in phylogenetic analysis.
One limitation of using the tests that we propose in this article is that their power will be limited if there are few differences between the sequences being examined. Indeed, our analyses show that in our representative sample of >3,500 partitions from published data sets, roughly ∼25% of the variance in whether a partition passes or fails a given test can be attributed to the number of observed differences between the sequences. Nevertheless, this implies that the remaining ∼75% of the variance in whether a partition passes or fails a test could be attributable to other processes, such as variation in the extent of model violation among partitions. This suggests that we should be cautiously optimistic: although a lack of power on small or slowly evolving partitions may induce some false negatives (i.e., failures to identify partitions that have evolved under non-SRH conditions), the tests we propose still have significant power to identify partitions that show the evidence of model violation. It is possible that removing such partitions from phylogenetic analyses may improve the accuracy of results by reducing the overall burden of model violation on the inference of the tree topology. We hope that our implementation of these tests in the user-friendly software IQ-TREE will allow empirical phylogeneticists to continue to explore whether this is the case.
Supplementary Material
Supplementary data are available at Genome Biology and Evolution online.
Supplementary Material
Acknowledgments
The authors would like to thank Lars Jermiin, David Bryant, Jeremy Brown, and one anonymous referee for providing thoughtful comments on this article. This work was supported by an Australian Research Council and Australian National University Future Scheme Grants to RML.
Literature Cited
- Ababneh F, Jermiin LS, Ma C, Robinson J.. 2006. Matched-pairs tests of homogeneity with applications to homologous nucleotide sequences. Bioinformatics 22(10):1225–1231. [DOI] [PubMed] [Google Scholar]
- Anderson FE, Bergman A, Cheng SH, Pankey MS, Valinassab T.. 2014. Lights out: the evolution of bacterial bioluminescence in Loliginidae. Hydrobiologia 725(1):189–203. In: Dryad Data Repository. doi:10.5061/dryad.93s3n. [Google Scholar]
- Barton K. 2009. MuMIn: multi-model inference, R package version 0.12. 0. Available from: http://r-forge. r-project. org/projects/mumin/, last accessed May 28, 2019.
- Bates D, Mächler M, Bolker B, Walker S.. 2015. Fitting linear mixed-effects models using lme4. J Stat Softw 67:48. [Google Scholar]
- Bazinet AL, Zwickl DJ, Cummings MP.. 2014. A gateway for phylogenetic analysis powered by grid computing featuring GARLI 2.0. Syst Biol. 63(5):812–818. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bergsten J, Nilsson AN, Ronquist F.. 2013. Bayesian tests of topology hypotheses with an example from diving beetles. Syst Biol. 62(5):660–673. In: Dryad Data Repository. doi:10.5061/dryad.s631d. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Betancur-r R, Li C, Munroe TA, Ballesteros JA, Ortí G.. 2013. Addressing gene tree discordance and non-stationarity to resolve a multi-locus phylogeny of the flatfishes (Teleostei: Pleuronectiformes). Syst Biol. 62(5):763–785. [DOI] [PubMed] [Google Scholar]
- Blanquart S, Lartillot N.. 2006. A Bayesian compound stochastic process for modeling nonstationary and nonhomogeneous sequence evolution. Mol Biol Evol. 23(11):2058–2071. [DOI] [PubMed] [Google Scholar]
- Bogdanowicz D, Giaro K, Wrobel B.. 2012. TreeCmp: comparison of trees in polynomial time. Evol Bioinformatics. 8:475–487. [Google Scholar]
- Bollback JP. 2002. Bayesian model adequacy and choice in phylogenetics. Mol Biol Evol. 19(7):1171–1180. [DOI] [PubMed] [Google Scholar]
- Bouckaert R, et al. 2014. BEAST 2: a software platform for Bayesian evolutionary analysis. PLoS Comput Biol. 10(4):e1003537.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bourlat SJ, et al. 2006. Deuterostome phylogeny reveals monophyletic chordates and the new phylum Xenoturbellida. Nature 444(7115):85.. [DOI] [PubMed] [Google Scholar]
- Boussau B, Gouy M.. 2006. Efficient likelihood computations with nonreversible models of evolution. Syst Biol. 55(5):756–768. [DOI] [PubMed] [Google Scholar]
- Bowker AH. 1948. A test for symmetry in contingency tables. J Am Stat Assoc. 43(244):572–574. [DOI] [PubMed] [Google Scholar]
- Brady A, Salzberg S.. 2011. PhymmBL expanded: confidence scores, custom databases, parallelization and more. Nat Methods. 8(5):367.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Broughton RE, Betancur RR, Li C, Arratia G, Orti G.. 2013. Multi-locus phylogenetic analysis reveals the pattern and tempo of bony fish evolution. PLoS Curr. 5. In: Dryad Data Repository. doi:10.1371/currents.tol.2ca8041495ffafd0c92756e75247483e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brown JM. 2014. Detection of implausible phylogenetic inferences using posterior predictive assessment of model fit. Syst Biol. 63(3):334–348. [DOI] [PubMed] [Google Scholar]
- Brown JM, ElDabaje R.. 2009. PuMA: Bayesian analysis of partitioned (and unpartitioned) model adequacy. Bioinformatics 25(4):537–538. [DOI] [PubMed] [Google Scholar]
- Brown JM, Thomson RC.. 2017. Bayes factors unmask highly variable information content, bias, and extreme influence in phylogenomic analyses. Syst Biol. 66(4):517–530. [DOI] [PubMed] [Google Scholar]
- Brown JM, Thomson RC.. 2018. Evaluating model performance in evolutionary biology. Annu Rev Ecol Evol Syst. 49:null. [Google Scholar]
- Brown RM, Siler CD, Das I, Min Y.. 2012. Testing the phylogenetic affinities of Southeast Asia’s rarest geckos: flap-legged geckos (Luperosaurus), flying geckos (Ptychozoon) and their relationship to the pan-Asian genus Gekko. Mol Phylogenet Evol. 63(3):915–921. In: Dryad Data Repository. doi:10.5061/dryad.7bn0fr99. [DOI] [PubMed] [Google Scholar]
- Cannon JT, et al. 2016. Xenacoelomorpha is the sister group to Nephrozoa. Nature 530(7588):89–93. In: Dryad Data Repository. doi:10.5061/dryad.493b7. [DOI] [PubMed] [Google Scholar]
- Cao Y, et al. 1998. Conflict among individual mitochondrial proteins in resolving the phylogeny of eutherian orders. J. Mol. Evol. 47:307–322. [DOI] [PubMed] [Google Scholar]
- Chernomor O, von Haeseler A, Minh BQ.. 2016. Terrace aware data structure for phylogenomic inference from supermatrices. Syst Biol. 65(6):997–1008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cognato AI, Vogler AP.. 2001. Exploring data interaction and nucleotide alignment in a multiple gene analysis of Ips (Coleoptera: Scolytinae). Syst Biol. 50(6):758–780. In: Dryad Data Repository. doi:10.5061/dryad.678. [DOI] [PubMed] [Google Scholar]
- Day JJ, et al. 2013. Continental diversification of an African catfish radiation (Mochokidae: Synodontis). Syst Biol. 62(3):351–365. In: Dryad Data Repository. doi:10.5061/dryad.b6225.2. [DOI] [PubMed] [Google Scholar]
- Delsuc F, Brinkmann H, Philippe H.. 2005. Phylogenomics and the reconstruction of the tree of life. Nat Rev Genet. 6(5):361.. [DOI] [PubMed] [Google Scholar]
- Devitt TJ, Cameron Devitt SE, Hollingsworth BD, McGuire JA, Moritz C.. 2013. Data from: Montane refugia predict population genetic structure in the Large-blotched Ensatina salamander In: Dryad Data Repository. doi:10.5061/dryad.k9g50. [DOI] [PubMed] [Google Scholar]
- Devitt TJ, Devitt SE, Hollingsworth BD, McGuire JA, Moritz C.. 2013. Montane refugia predict population genetic structure in the large-blotched Ensatina salamander. Mol Ecol. 22(6):1650–1665. In: Dryad Data Repository. doi:10.5061/dryad.k9g50. [DOI] [PubMed] [Google Scholar]
- Dornburg A, et al. 2012. Molecular phylogenetics of squirrelfishes and soldierfishes (Teleostei: Beryciformes: Holocentridae): reconciling more than 100 years of taxonomic confusion. Mol Phylogenet Evol. 65(2):727–738. In: Dryad Data Repository. doi:10.5061/dryad.3t19n. [DOI] [PubMed] [Google Scholar]
- Drummond AJ, Rambaut A.. 2007. BEAST: Bayesian evolutionary analysis by sampling trees. BMC Evol Biol. 7:214.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Duchene DA, Duchene S, Ho S.. 2017. New statistical criteria detect phylogenetic bias caused by compositional heterogeneity. Mol Biol Evol. 34(6):1529–1534. [DOI] [PubMed] [Google Scholar]
- Dunn M, Greenhill SJ, Levinson SC, Gray RD.. 2011. Evolved structure of language shows lineage-specific trends in word-order universals. Nature 473(7345):79.. [DOI] [PubMed] [Google Scholar]
- Dutheil J, Boussau B.. 2008. Non-homogeneous models of sequence evolution in the Bio++ suite of libraries and programs. BMC Evol Biol. 8:255.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eisen JA. 1998. Phylogenomics: improving functional predictions for uncharacterized genes by evolutionary analysis. Genome Res. 8(3):163–167. [DOI] [PubMed] [Google Scholar]
- Faircloth BC, Sorenson L, Santini F, Alfaro ME.. 2013. A phylogenomic perspective on the radiation of ray-finned fishes based upon targeted sequencing of ultraconserved elements (UCEs). PLoS One 8(6):e65923. In: Dryad Data Repository. doi:10.5061/dryad.j015n. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Farrell LE, Roman J, Sunquist ME.. 2000. Dietary separation of sympatric carnivores identified by molecular analysis of scats. Mol Ecol. 9(10):1583–1590. [DOI] [PubMed] [Google Scholar]
- Felsenstein J. 2004. Inferring phylogenies. Sunderland (MA): Sinauer Associates.
- Fong JJ, Brown JM, Fujita MK, Boussau B.. 2012. A phylogenomic approach to vertebrate phylogeny supports a turtle-archosaur affinity and a possible paraphyletic lissamphibia. PLoS One 7(11):e48990. In: Dryad Data Repository. doi:10.5061/dryad.25j6h. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Foster PG. 2004. Modeling compositional heterogeneity. Syst Biol. 53(3):485–495. [DOI] [PubMed] [Google Scholar]
- Foster PG, Hickey DA.. 1999. Compositional bias may affect both DNA-based and protein-based phylogenetic reconstructions. J Mol Evol. 48(3):284–290. [DOI] [PubMed] [Google Scholar]
- Gardner MJ, et al. 2002. Genome sequence of the human malaria parasite Plasmodium falciparum. Nature 419(6906):498.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goldman N. 1993. Statistical tests of models of DNA substitution. J Mol Evol. 36(2):182–198. [DOI] [PubMed] [Google Scholar]
- Goremykin V, Hellwig F.. 2005. Evidence for the most basal split in land plants dividing bryophyte and tracheophyte lineages. Plant Syst Evol. 254(1–2):93–103. [Google Scholar]
- Graur D, Gouy M, Duret L.. 1997. Evolutionary affinities of the order Perissodactyla and the phylogenetic status of the superordinal taxa Ungulata and Altungulata. Mol Phylogenet Evol. 7(2):195–200. [DOI] [PubMed] [Google Scholar]
- Gray RD, Drummond AJ, Greenhill SJ.. 2009. Language phylogenies reveal expansion pulses and pauses in Pacific settlement. Science 323(5913):479.. [DOI] [PubMed] [Google Scholar]
- Grenfell BT, et al. 2004. Unifying the epidemiological and evolutionary dynamics of pathogens. Science 303(5656):327.. [DOI] [PubMed] [Google Scholar]
- Groussin M, Boussau B, Gouy M.. 2013. A branch-heterogeneous model of protein evolution for efficient inference of ancestral sequences. Syst Biol. 62(4):523–538. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guindon S, et al. 2010. New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0. Syst Biol. 59(3):307–321. [DOI] [PubMed] [Google Scholar]
- Hasegawa M, Kishino H, Yano T-A.. 1985. Dating of the human-ape splitting by a molecular clock of mitochondrial DNA. J Mol Evol. 22(2):160–174. [DOI] [PubMed] [Google Scholar]
- Ho JW, et al. 2006. SeqVis: visualization of compositional heterogeneity in large alignments of nucleotides. Bioinformatics 22(17):2162–2163. [DOI] [PubMed] [Google Scholar]
- Ho SY, Jermiin L.. 2004. Tracing the decay of the historical signal in biological sequence data. Syst Biol. 53(4):623–637. [DOI] [PubMed] [Google Scholar]
- Hoang DT, Chernomor O, von Haeseler A, Minh BQ, Vinh LS.. 2018. UFBoot2: improving the ultrafast bootstrap approximation. Mol Biol Evol. 35(2):518–522. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Höhna S, et al. 2016. RevBayes: Bayesian phylogenetic inference using graphical models and an interactive model-specification language. Syst Biol. 65(4):726–736. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Horn JW, et al. 2014. Evolutionary bursts in Euphorbia (Euphorbiaceae) are linked with photosynthetic pathway. Evolution 68(12):3485–3504. In: Dryad Data Repository. doi:10.5061/dryad.sb1j1. [DOI] [PubMed] [Google Scholar]
- Hyman IT, Ho SY, Jermiin LS.. 2007. Molecular phylogeny of Australian Helicarionidae, Euconulidae and related groups (Gastropoda: Pulmonata: Stylommatophora) based on mitochondrial DNA. Mol Phylogenet Evol. 45(3):792–812. [DOI] [PubMed] [Google Scholar]
- Jayaswal V, Ababneh F, Jermiin LS, Robinson J.. 2011. Reducing model complexity of the general Markov model of evolution. Mol Biol Evol. 28(11):3045–3059. [DOI] [PubMed] [Google Scholar]
- Jayaswal V, Jermiin LS, Robinson J.. 2005. Estimation of Phylogeny Using a General Markov Model. Evol Bioinform 1:62–80. [PMC free article] [PubMed] [Google Scholar]
- Jayaswal V, Robinson J, Jermiin L.. 2007. Estimation of phylogeny and invariant sites under the general Markov model of nucleotide sequence evolution. Syst Biol. 56(2):155–162. [DOI] [PubMed] [Google Scholar]
- Jayaswal V, Wong TK, Robinson J, Poladian L, Jermiin LS.. 2014. Mixture models of nucleotide sequence evolution that account for heterogeneity in the substitution process across sites and across lineages. Syst Biol. 63(5):726–742. [DOI] [PubMed] [Google Scholar]
- Jermiin L, Ho SY, Ababneh F, Robinson J, Larkum AW.. 2004. The biasing effect of compositional heterogeneity on phylogenetic estimates may be underestimated. Syst Biol. 53(4):638–643. [DOI] [PubMed] [Google Scholar]
- Jermiin LS, Jayaswal V, Ababneh FM, Robinson J.. 2017. Identifying optimal models of evolution In: Keith JM, editor. Bioinformatics. Melbourne: Humana Press, New York, NY: p. 379–420. [DOI] [PubMed] [Google Scholar]
- Kalyaanamoorthy S, Minh BQ, Wong TKF, von Haeseler A, Jermiin LS.. 2017. ModelFinder: fast model selection for accurate phylogenetic estimates. Nat Methods. 14(6):587–589. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kawahara AY, Rubinoff D.. 2013. Convergent evolution of morphology and habitat use in the explosive Hawaiian fancy case caterpillar radiation. J Evol Biol. 26(8):1763–1773. In: Dryad Data Repository. doi:10.5061/dryad.gh895. [DOI] [PubMed] [Google Scholar]
- Kimura M. 1980. A simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide sequences. J Mol Evol. 16(2):111–120. [DOI] [PubMed] [Google Scholar]
- Kishino H, Miyata T, Hasegawa M.. 1990. Maximum likelihood inference of protein phylogeny and the origin of chloroplasts. J Mol Evol. 31(2):151–160. [Google Scholar]
- Knight R, et al. 2007. PyCogent: a toolkit for making sense from sequence. Genome Biol. 8(8):R171.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kumar S, Filipski AJ, Battistuzzi FU, Kosakovsky Pond SL, Tamura K.. 2012. Statistics and truth in phylogenomics. Mol Biol Evol. 29(2):457–472. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kumar S, Gadagkar SR.. 2001. Disparity index: a simple statistic to measure and test the homogeneity of substitution patterns between molecular sequences. Genetics 158(3):1321–1327. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lartillot N, Delsuc F.. 2012. Joint reconstruction of divergence times and life-history evolution in placental mammals using a phylogenetic covariance model. Evolution 66(6):1773–1787. In: Dryad Data Repository. doi:10.5061/dryad.tt28qk6f. [DOI] [PubMed] [Google Scholar]
- Lartillot N, Philippe H.. 2004. A Bayesian mixture model for across-site heterogeneities in the amino-acid replacement process. Mol Biol Evol. 21(6):1095–1109. [DOI] [PubMed] [Google Scholar]
- Liu L, et al. 2017. Genomic evidence reveals a radiation of placental mammals uninterrupted by the KPg boundary. Proc Natl Acad Sci U S A. 114(35):E7282–E7290. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martijn J, Vosseberg J, Guy L, Offre P, Ettema TJ.. 2018. Deep mitochondrial origin outside the sampled alphaproteobacteria. Nature 557(7703):101.. [DOI] [PubMed] [Google Scholar]
- Mäser P, et al. 2001. Phylogenetic relationships within cation transporter families of Arabidopsis. Plant Physiol. 126(4):1646. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McCormack JE, et al. 2013. A phylogeny of birds based on over 1,500 loci collected by target enrichment and high-throughput sequencing. PLoS One 8(1):e54848. In: Dryad Data Repository. doi:10.5061/dryad.sd080. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mir A, Russello F.. 2010. The mean value of the squared path-difference distance for rooted phylogenetic trees. J Math Anal Appl. 371(1):168–176. [Google Scholar]
- Moyle RG, et al. 2016. Tectonic collision and uplift of Wallacea triggered the global songbird radiation. Nat Commun. 7(1):12709. In: Dryad Data Repository. doi:10.5061/dryad.nf01p. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Murphy WJ, et al. 2001. Molecular phylogenetics and the origins of placental mammals. Nature 409(6820):614–618. [DOI] [PubMed] [Google Scholar]
- Murray EA, Carmichael AE, Heraty JM.. 2013. Ancient host shifts followed by host conservatism in a group of ant parasitoids. Proc Biol Sci. 280(1759):20130495. In: Dryad Data Repository. doi:10.5061/dryad.qn57t. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Murray S, Jørgensen MF, Ho SY, Patterson DJ, Jermiin LS.. 2005. Improving the analysis of dinoflagellate phylogeny based on rDNA. Protist 156(3):269–286. [DOI] [PubMed] [Google Scholar]
- Nabholz B, Künstner A, Wang R, Jarvis ED, Ellegren H.. 2011. Dynamic evolution of base composition: causes and consequences in avian phylogenomics. Mol Biol Evol. 28(8):2197–2210. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nakagawa S, Schielzeth H.. 2013. A general and simple method for obtaining R2 from generalized linear mixed‐effects models. Methods Ecol Evol. 4(2):133–142. [Google Scholar]
- Nesnidal MP, Helmkampf M, Bruchhaus I, Hausdorf B.. 2010. Compositional heterogeneity and phylogenomic inference of metazoan relationships. Mol Biol Evol. 27(9):2095–2104. [DOI] [PubMed] [Google Scholar]
- Nguyen LT, Schmidt HA, von Haeseler A, Minh BQ.. 2015. IQ-TREE: a fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol Biol Evol. 32(1):268–274. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oaks JR. 2011. A time-calibrated species tree of Crocodylia reveals a recent radiation of the true crocodiles. Evolution 65(11):3285–3297. In: Dryad Data Repository. doi:10.5061/dryad.5k9s0. [DOI] [PubMed] [Google Scholar]
- Paton T, Haddrath O, Baker AJ.. 2002. Complete mitochondrial DNA genome sequences show that modern birds are not descended from transitional shorebirds. Proc R Soc Lond B. 269(1493):839–846. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Philippe H, Delsuc F, Brinkmann H, Lartillot N.. 2005. Phylogenomics. Annu Rev Ecol Evol Syst. 36(1):541–562. [Google Scholar]
- Philippe H, et al. 2011. Acoelomorph flatworms are deuterostomes related to Xenoturbella. Nature 470(7333):255.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Phillips MJ, Delsuc F, Penny D.. 2004. Genome-scale phylogeny and the detection of systematic biases. Mol Biol Evol. 21(7):1455–1458. [DOI] [PubMed] [Google Scholar]
- Rightmyer MG, Griswold T, Brady SG.. 2013. Phylogeny and systematics of the bee genus Osmia (Hymenoptera: Megachilidae) with emphasis on North American Melanosmia: subgenera, synonymies and nesting biology revisited. Syst Entomol. 38(3):561–576. In: Dryad Data Repository. doi:10.5061/dryad.jd5ff. [Google Scholar]
- Roberts D, Yang Z.. 1995. On the use of nucleic acid sequences to infer early branchings in the tree of life. Mol Biol Evol. 12(3):451–458. [DOI] [PubMed] [Google Scholar]
- Ronquist F, et al. 2012. MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Syst Biol. 61(3):539–542. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rzhetsky A, Nei M.. 1995. Tests of applicability of several substitution models for DNA sequence data. Mol Biol Evol. 12(1):131–151. [DOI] [PubMed] [Google Scholar]
- Salipante SJ, Horwitz MS.. 2006. Phylogenetic fate mapping. Proc Natl Acad Sci U S A. 103(14):5448.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sand A, et al. 2014. tqDist: a library for computing the quartet and triplet distances between binary or general trees. Bioinformatics 30(14):2079–2080. [DOI] [PubMed] [Google Scholar]
- Sauquet H, et al. 2012. Testing the impact of calibration on molecular divergence times using a fossil-rich group: the case of Nothofagus (Fagales). Syst Biol. 61(2):289–313. In: Dryad Data Repository. doi:10.5061/dryad.qq106tm4. [DOI] [PubMed] [Google Scholar]
- Seago AE, Giorgi JA, Li J, Ślipiński A.. 2011. Phylogeny, classification and evolution of ladybird beetles (Coleoptera: Coccinellidae) based on simultaneous analysis of molecular and morphological data. Mol Phylogenet Evol. 60(1):137–151. In: Dryad Data Repository. doi:10.5061/dryad.dc1r2. [DOI] [PubMed] [Google Scholar]
- Sharanowski BJ, Dowling APG, Sharkey MJ.. 2011. Molecular phylogenetics of Braconidae (Hymenoptera: Ichneumonoidea), based on multiple nuclear genes, and implications for classification. Syst Entomol. 36(3):549–572. In: Dryad Data Repository. doi:10.5061/dryad.1688p. [Google Scholar]
- Sheffield NC, Song H, Cameron SL, Whiting MF.. 2009. Nonstationary evolution and compositional heterogeneity in beetle mitochondrial phylogenomics. Syst Biol. 58(4):381–394. [DOI] [PubMed] [Google Scholar]
- Shimodaira H. 2002. An approximately unbiased test of phylogenetic tree selection. Syst Biol. 51(3):492–508. [DOI] [PubMed] [Google Scholar]
- Shimodaira H, Hasegawa M.. 1999. Multiple comparisons of log-likelihoods with applications to phylogenetic inference. Mol Biol Evol. 16(8):1114–1116. [Google Scholar]
- Siler CD, Oliveros CH, Santanen A, Brown RM.. 2013. Multilocus phylogeny reveals unexpected diversification patterns in Asian wolf snakes (genus Lycodon). Zool Scr. 42(3):262–277. In: Dryad Data Repository. doi:10.5061/dryad.cp6gg. [Google Scholar]
- Stamatakis A. 2014. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30(9):1312–1313. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Steel MA, Penny D.. 1993. Distributions of tree comparison metrics – some new results. Syst Biol. 42:126–141. [Google Scholar]
- Stuart A. 1955. A test for homogeneity of the marginal distributions in a two-way classification. Biometrika 42(3–4):412–416. [Google Scholar]
- Sullivan J, Joyce P.. 2005. Model selection in phylogenetics. Annu Rev Ecol Evol Syst. 36(1):445–466. [Google Scholar]
- Sumner JG, Fernandez-Sanchez J, Jarvis PD.. 2012. Lie Markov models. J Theor Biol. 298:16–31. [DOI] [PubMed] [Google Scholar]
- Swofford DL. 2001. Paup*: phylogenetic analysis using parsimony (and other methods) 4.0. B5. Sunderland (MA): Sinauer Associates.
- Tarrío R, Rodríguez-Trelles F, Ayala FJ.. 2001. Shared nucleotide composition biases among species and their impact on phylogenetic reconstructions of the Drosophilidae. Mol Biol Evol. 18(8):1464–1473. [DOI] [PubMed] [Google Scholar]
- Tolley KA, Townsend TM, Vences M.. 2013. Large-scale phylogeny of chameleons suggests African origins and Eocene diversification. Proc R Soc B. 280(1759):20130184. In: Dryad Data Repository. doi:10.5061/dryad.11350. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tsagkogeorga G, Parker J, Stupka E, Cotton JA, Rossiter SJ.. 2013. Phylogenomic analyses elucidate the evolutionary relationships of bats. Curr Biol. 23(22):2262–2267. [DOI] [PubMed] [Google Scholar]
- Unmack PJ, Allen GR, Johnson JB.. 2013. Phylogeny and biogeography of rainbowfishes (Melanotaeniidae) from Australia and New Guinea. Mol Phylogenet Evol. 67(1):15–27. In: Dryad Data Repository. doi:10.5061/dryad.qq846. [DOI] [PubMed] [Google Scholar]
- Wainwright PC, et al. 2012. The evolution of pharyngognathy: a phylogenetic and functional appraisal of the pharyngeal jaw key innovation in labroid fishes and beyond. Syst Biol. 61(6):1001–1027. [DOI] [PubMed] [Google Scholar]
- Weiss G, von Haeseler A.. 2003. Testing substitution models within a phylogenetic tree. Mol Biol Evol. 20(4):572–578. [DOI] [PubMed] [Google Scholar]
- Wood HM, Matzke NJ, Gillespie RG, Griswold CE.. 2013. Treating fossils as terminal taxa in divergence time estimation reveals ancient vicariance patterns in the palpimanoid spiders. Syst Biol. 62(2):264–284. In: Dryad Data Repository. doi:10.5061/dryad.7231d.2. [DOI] [PubMed] [Google Scholar]
- Woodhams MD, Fernandez-Sanchez J, Sumner JG.. 2015. A new hierarchy of phylogenetic models consistent with heterogeneous substitution rates. Syst Biol. 64(4):638–650. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Worobey M, Han GZ, Rambaut A.. 2014. A synchronized global sweep of the internal genes of modern avian influenza virus. Nature 508(7495):254–257. In: Dryad Data Repository. doi:10.5061/dryad.m04j9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang Z. 1994. Estimating the pattern of nucleotide substitution. J Mol Evol. 39(1):105–111. [DOI] [PubMed] [Google Scholar]
- Yang Z, Rannala B.. 2012. Molecular phylogenetics: principles and practice. Nat Rev Genet. 13(5):303–314. [DOI] [PubMed] [Google Scholar]
- Yao H, et al. 2003. An accurate, sensitive, and scalable method to identify functional sites in protein structures. J Mol Biol. 326(1):255–261. [DOI] [PubMed] [Google Scholar]
- Yao Y-G, Bravi CM, Bandelt H-J.. 2004. A call for mtDNA data quality control in forensic science. Forensic Sci Int. 141(1):1–6. [DOI] [PubMed] [Google Scholar]
- Yap VB, Speed T.. 2005. Rooting a phylogenetic tree with nonreversible substitution models. BMC Evol Biol. 5:2.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou X, et al. 2012. Phylogenomic analysis resolves the interordinal relationships and rapid diversification of the laurasiatherian mammals. Syst Biol. 61(1):150–164. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zou L, Susko E, Field C, Roger AJ.. 2012. Fitting nonstationary general-time-reversible models to obtain edge-lengths and frequencies for the Barry–Hartigan model. Syst Biol. 61(6):927–940. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.