• Nenhum resultado encontrado

Measures of clade confidence do not correlate with accuracy of phylogenetic trees.

N/A
N/A
Protected

Academic year: 2017

Share "Measures of clade confidence do not correlate with accuracy of phylogenetic trees."

Copied!
9
0
0

Texto

(1)

Measures of Clade Confidence Do Not Correlate

with Accuracy of Phylogenetic Trees

Barry G. Hall1*, Stephen J. Salipante2

1Bellingham Research Institute, Bellingham, Washington, United States of America,2Department of Genome Sciences, University of Washington, Seattle, Washington, United States of America

Metrics of phylogenetic tree reliability, such as parametric bootstrap percentages or Bayesian posterior probabilities, represent internal measures of the topological reproducibility of a phylogenetic tree, while the recently introduced aLRT (approximate likelihood ratio test) assesses the likelihood that a branch exists on a maximum-likelihood tree. Although those values are often equated with phylogenetic tree accuracy, they do not necessarily estimate how well a reconstructed phylogeny represents cladistic relationships that actually exist in nature. The authors have therefore attempted to quantify how well bootstrap percentages, posterior probabilities, and aLRT measures reflect the probability that a deduced phylogenetic clade is present in a known phylogeny. The authors simulated the evolution of bacterial genes of varying lengths under biologically realistic conditions, and reconstructed those known phylogenies using both maximum likelihood and Bayesian methods. Then, they measured how frequently clades in the reconstructed trees exhibiting particular bootstrap percentages, aLRT values, or posterior probabilities were found in the true trees. The authors have observed that none of these values correlate with the probability that a given clade is present in the known phylogeny. The major conclusion is that none of the measures provide any information about the likelihood that an individual clade actually exists. It is also found that the mean of all clade support values on a tree closely reflects the average proportion of all clades that have been assigned correctly, and is thus a good representation of the overall accuracy of a phylogenetic tree.

Citation: Hall BG, Salipante SJ (2007) Measures of clade confidence do not correlate with accuracy of phylogenetic trees. PLoS Comput Biol 3(3): e51. doi:10.1371/journal.pcbi. 0030051

Introduction

Phylogenetic analysis, once the province of systematists and evolutionary biologists, has become a fundamental tool of computational biology and biological disciplines as diverse as biochemistry, epidemiology, and developmental biology. While systematists use phylogenetic analysis of molecular sequences to elucidate the historical relationships among species, other disciplines tend to focus more on the historical relationships of the sequences themselves. The results of phylogenetic analyses are typically presented as phylogenetic trees, diagrams that graphically illustrate those historical relationships. Phylogenetic trees are just estimates of those historical relationships, and it is therefore important to have some way to evaluate the quality and reliability of phyloge-netic reconstructions.

The most widely used method of estimating the reliability of trees is the nonparametric bootstrap [1]. The bootstrap method addresses the reliability of the tree topology (the branching order) by calculating the bootstrap percentage (BP) for each interior node, or clade, in a tree. In the bootstrap method, the sites in a set of aligned sequences are randomly sampled with replacement to create a pseudo-alignment, and that pseudo-alignment is used to produce an estimated ‘‘bootstrap tree.’’ Typically, 100–2,000 bootstrap trees are estimated, and the BP for a clade on the original phylogenetic tree is the percentage of the bootstrap trees that also include that clade. Thus, confidence in the groupings of taxa can be estimated. A drawback to the bootstrap method is that it can potentially be very time-consuming. For example, maximum likelihood is at present the most widely used statistical phylogenetic method, but because it is

computa-tionally intensive, performing a bootstrap analysis on maximum likelihood trees can require prohibitive amounts of time.

Recently, a new approach to estimating branch (or clade) support, the approximate likelihood ratio test (aLRT), has been introduced [2]. The aLRT is a fast and accurate method for assessing branch support for maximum likelihood trees. Under conventional LRT, the null hypothesis is that the branch has a length of zero (i.e., it does not exist), and the test statistic is 2(l1l0),wherel1is the likelihood of the most likely tree andl0 is the likelihood of the tree in which the branch does not exist. In aLRT, the test statistic is approximated by 2(l1l2), wherel2is the likelihood of the second most likely tree, an approximation that enormously decreases computa-tional time and results in a practical and slightly conservative test statistic. The significance of the aLRT test statistic is calculated from a mixedv2distribution, with half drawn from zero and half drawn from one degree of freedom. The aLRT approach is implemented in the beta version of PHYML 2.4.5 [3] (http://atgc.lirmm.fr/alrt). The most recent release of the

Editor:Wen-Hsiung Li, University of Chicago, United States of America

ReceivedNovember 6, 2006;AcceptedJanuary 31, 2007;PublishedMarch 16, 2007

Copyright:Ó2007 Hall and Salipante. This is an open-access article distributed

under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.

Abbreviations:aLRT, approximate likelihood ratio test; BP, bootstrap percentage; indels, insertions or deletions; PP, posterior probability; SH-like, Shimodaira– Hasegawa-like

* To whom correspondence should be addressed. E-mail: barryhall@zeninternet. com

(2)

beta version of PHYML also implements an alternative nonparametric Shimodaira–Hasegawa-like (SH-like) proce-dure that is typically more conservative than thev2approach, so PHYML now offers the option of assigning support as the smaller of the values calculated by the two methods.

In the last decade, a new method of estimating phyloge-netic trees, the Bayesian method, has gained increasing popularity [4–7]. The Bayesian method, as implemented by the program MrBayes [8,9], estimates the posterior proba-bilities (PPs) of clades by calculating, among the trees with the highest posterior probabilities, the fraction of the time that each clade appears as those trees are visited in proportion to their probabilities. The Bayesian method has the advantage that it calculates PPs during the process of estimating the consensus tree. It is therefore much faster to obtain PP estimates of clade reliability by the Bayesian method than to obtain BPs of clade reliability by maximum likelihood.

BP, aLRT, and PP are measures of clade support, but they are often presented as measures of the accuracy of the tree [5,10]. None, however, is a metric of accuracy. aLRT assesses the likelihood that a branch exists on a maximum likelihood tree. BP and PP are simply measures of repeatability; BP measures the repeatability with which a clade occurs among subsamples of the data used to create the original tree, and PP measures the repeatability with which a clade occurs among the set of nearly equally likely trees after the Bayesian process has converged on a set of trees with nearly identical likelihoods. Because of discrepancies between Bayesian posterior probabilities and bootstrapped maximum like-lihood percentages, there has been considerable controversy about BP versus PP as measures of clade reliability [11–15].

For real, empirical data, we cannot know the accuracy of a tree because we have no way of knowing the true branching order of the taxa or sequences that are being considered. Simulated datasets, in which the true tree is known, have been used to compare BP and PP with the accuracies of estimated trees. Several such studies have shown that BP underestimates clade reliability (i.e., clades in the estimated tree are more likely to exist in the true tree than is indicated by BP)

[10,11,13–16]. There have been conflicting reports about the relationship between PP and accuracy. In general, PP has been found to be less conservative than BP. Some studies [11,12] have concluded that PP is too liberal (i.e., over-estimates accuracy), while others [13,14] conclude that PP better reflects accuracy. Another study concluded that BP and PP can be taken as potential upper and lower estimates of accuracy, but that they are not interchangeable and cannot be directly compared [15]. Similarly, Anisimova and Gascuel reported that aLRT using the v2 approach is similar to posterior probabilities, and their unpublished data suggest that the SH-like approach is more conservative than the v2 approach (http://atgc.lirmm.fr/alrt).

The conclusions of the above studies are only as reliable as the extent to which the simulations mimic real evolutionary processes that generate the empirical data to which we actually apply phylogenetic methods. In all cases, the simulations incorporate specific evolutionary models, the most common being the K2P model [17], to guide the simulation process. The results will be no more realistic than the assumptions and biases of that model. Modeling evolution as a process of substitution confounds two distinct processes, mutation and selection, the outcome of which is the real substitution process [18]. The number of taxa used in the simulations reported in [10–16] ranges from four to 28. Many, and probably most, phylogenetic studies involve many more taxa. Typically, branch lengths are uniform, although some studies included a specific pattern of length variation. Importantly, the simulations only consider base substitutions, not insertions or deletions (indels). The resulting sequences thus need not be aligned. In reality, historical indels necessitate using multiple alignment programs to estimate the homologous characters within. The alignment process strongly affects the reliability of the resulting trees. For coding sequences, the accuracy of a tree is significantly increased by aligning the corresponding protein sequences and using that alignment to place the corresponding gaps into the DNA coding sequences [18]. However, when the average percentage identity of the amino acids is within the

‘‘twilight zone’’ of 20%–30%, only 80% of residues are

correctly aligned [19], and when identity is below 10%, less than 50% of residues are correctly aligned [20]. The failure to include indels in the simulation process therefore reduces considerably the confidence we can place in applying conclusions drawn from those simulations to real data.

The EvolveAGene simulation program [18,21] was designed to mimic the evolution of sequences in a more biologically realistic fashion. A real sequence is used for the root node of the tree, and a strictly bifurcating bilaterally symmetrical tree is evolved. Branch lengths are randomly varied from zero to a value chosen by the user. Mutation and selection are treated as separate processes. The mutation process is simulated by introducing random mutations, including base substitutions, insertions, and deletions, into the sequences according to the spontaneous mutation spectrum of Escherichia coli. (The mutational spectrum is the experimentally determined relative frequencies with which the various base substitutions and indels of different lengths occur, before selection or drift act on those mutations.) The selection process is simulated by (1) assuming that all frameshift and nonsense mutations are strongly deleterious and thus not accepting those mutations, and (2) accepting nonsynonymous base substitutions with a

PLoS Computational Biology | www.ploscompbiol.org 0474 March 2007 | Volume 3 | Issue 3 | e51

Author Summary

The construction of phylogenetic trees, which depict past relation-ships between groups of DNA or protein sequences, has valuable application in many fields of study, most commonly evolutionary and population biology. Before drawing conclusions from phyloge-netic trees, it is important to assess how accurate those reconstructions are. This is typically accomplished by examining measures of ‘‘clade credibility’’ (such as bootstrap or posterior probability values), which represent how reproducible relationships are within the tree based on the parameters of the phylogenetic analysis. However, such measures do not necessarily reflect how likely inferred relationships are to have actually occurred in nature. Therefore, using simulated data where relationships are known, we have determined how well several measures of clade credibility correlate with the likelihood that a deduced phylogenetic grouping actually exists in reality. Surprisingly, we found no such correlation, and that the inferred relationships were correctly assigned about as often in cases where clade credibility values were very low as where they were high. This finding suggests that current measures of phylogenetic tree reliability are not useful in predicting whether specific inferred relationships have actually occurred.

(3)

probability that corresponds to a user-specified nonsynon-ymous substitution per nonsynonnonsynon-ymous site to synonnonsynon-ymous substitution per synonymous site (dN/dS) ratio, which can be set to biologically realistic values. The EvolveAGene program has been used to compare accuracies of various phylogenetic methods [18] and to explore the accuracies with which parsimony and Bayesian methods can reconstruct ancestral protein sequences [21].

In this study we are not particularly interested in comparing PP, BP, and aLRT per se. Instead, we are interested in asking two questions: (1) for all measures of clade credibility, how well does the credibility of a clade reflect the probability that that clade really exists on the true tree; and (2) how well does the average clade support reflect the topological accuracy of the tree? Topological accuracy is defined as the fraction of clades on the estimated tree that actually exist on the true tree. In this study, both the true tree and the estimated trees are strictly bifurcating, so the number of interior clades is the same. Thus, the number of false positive errors (clades found on the estimated tree that do not exist on the true tree) is identical to the number of false negative errors (clades on the true tree that are not on the estimated tree). We simulate the evolution of several genes under biologically realistic conditions and find that none of the estimates of clade support correlates with topological accuracy; in other words, clade supports tell us nothing about the likelihood that an inferred clade actually exists. However, we find that the average clade support does correlate well with the topological accuracy of the tree.

Results/Discussion

Ten simulations were initiated from each of fiveE. coliK12 coding sequences to assess how well BPs from maximum likelihood trees, aLRT support by the v2 approach, aLRT support by the minimum of SH-like andv2approaches, and posterior probabilities from Bayesian trees corresponded to clade accuracies. The simulation conditions were chosen to generate datasets that were at the practical limits for reliable alignments. Indeed, the typical average Jukes–Cantor [22] distances among the sequences for those datasets was 1.396 .02 substitutions per site, well above the limit of 1.0, above which Nei and Kumar [23] state that neighbor-joining trees are unreliable.

We define‘‘true clades’’as clades in the estimated tree that exist in the true tree, and we define‘‘accuracy’’as the percent of the total clades in the estimated tree that are true clades. Tables 1–4 show, respectively, the results for Bayesian trees and for maximum likelihood trees by the bootstrap and the two aLRT approaches. In each case, rows are ranges of clade credibility values.

For all methods, accuracy increases as the lengths of the sequences increase. Mean BPs are conservative estimates of mean topological accuracy, and in keeping with [13] and [14] we find that average posterior probabilities are a better estimate of topological accuracy than are BPs. BPs under-estimate accuracy, particularly for trees based on the shorter sequences, more than do posterior probabilities or aLRT supports. We do not interpret this finding to mean that BPs should be the‘‘gold standard’’measure of reliability; indeed, we find the notion that the less-accurate estimate should be the gold standard to be slightly ludicrous.

Table 1. Bayesian Trees PP Ra nge nuoK (30 0 bp) rplF (530 bp) tau B (76 3 bp) add (999 bp) araB (1,698 bp) Num ber of Clade s Frac tion True Clades Num ber of Cla des Fra ction True Clade s Num ber of Cla des Fra ction True Cla des Numbe r of Clades Fra ction True Cla des Numbe r of Clades Fraction Tru e Clades 0 % – 1 0 % 0— 0— 0 — 0 — 0— 11%–20% 4 1.000 0 — 0 — 0 — 0 — 21%–30% 3 1.000 2 1.000 2 1.000 2 1.000 0 — 31%–40% 8 1.000 2 1.000 5 1.000 3 1.000 5 1.000 41%–50% 2 2 0.864 15 1.000 13 1.000 13 0.846 7 1.000 51%–60% 2 2 1.000 19 0.947 14 0.929 10 1.000 6 1.000 61%–70% 14 1 .000 9 1 .000 1 5 0 .867 14 0.929 5 0.800 71%–80% 2 6 0.846 14 0.929 13 0.923 17 0.941 7 1.000 81%–90% 2 5 0.960 23 1.000 15 0.867 18 1.000 11 1.000 91%–100% 486 0 .938 526 0 .943 533 0.962 533 0.961 5 69 0.982 0%–40% 15 1.000 4 1.000 7 1.000 5 1.000 5 1.000 81%–100% 511 0 .939 549 0 .945 548 0.959 551 0.962 5 80 0.982 Mean PP 6 SE 91.79% 6 0.699% 94.93% 6 0.54% 95.16% 6 0.548% 9 5.40% 6 0.522% 97.40% 6 0.411% Mean accuracy 93.7% 94.8% 95.7% 9 5.9% 98.1% Slope 6 SE a 0.09 6 0.08 0.07 6 0.05 0.15 6 0.07 0.007 6 0.09 0.02 6 0.15 p-Value b 0.29 0.19 0.09 0.94 0.91 aSlope of midp oint of cla de supp ort range versu s fraction of true clades. bThe p-value is the probab ility that the slope is no t diff erent from zero. doi:10.1371/ journal.pcb i.0030051 .t001

PLoS Computational Biology | www.ploscompbiol.org 0475 March 2007 | Volume 3 | Issue 3 | e51

(4)

Table 2.Maximum Likelihood Tree Bootstrap Clade Support

BP Range nuoK(300 bp) rplF(530 bp) tauB(763 bp) add(999 bp) araB(1,698 bp)

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

0%–10% 10 0.600 0 — 0 — 0 — 0 —

11%–20% 18 0.889 2 1.000 0 — 1 1.000 0 —

21%–30% 11 0.818 4 1.000 5 0.800 1 1.000 0 —

31%–40% 22 0.773 10 1.000 8 1.000 6 0.833 4 1.000

41%–50% 30 0.933 14 1.000 14 1.000 14 1.000 7 1.000

51%–60% 36 0.889 15 0.933 13 0.846 15 0.933 8 1.000

61%–70% 41 0.902 26 0.923 19 0.947 19 0.947 14 0.857

71%–80% 43 0.930 35 0.943 17 1.000 16 0.875 16 1.000

81%–90% 94 0.904 57 0.930 20 0.900 27 1.000 11 1.000

91%–100% 305 0.918 447 0.960 514 0.951 511 0.951 550 0.975

0%–40% 61 0.852 16 1.000 13 0.923 8 0.875 4 1.000

81%–100% 399 0.914 504 0.956 534 0.949 538 0.953 561 0.975

Mean bootstrap6SE 79.42%61.0% 90.50%60 .656% 93.41%60.611% 93.94%60.583% 96.5%60.45%

Mean accuracy 90.2 % 95.6% 94.9% 95.1% 97.4%

Slope6SEa 0.23

60.09 0.0960.03 0.0860.12 0.0360.08 0.0360.11

p-Valueb 0.03 0.02 0.64 0.75 0.82

a

Slope of midpoint of clade support range versus fraction of true clades.

b

Thep-value is the probability that the slope is not different from zero. doi:10.1371/journal.pcbi.0030051.t002

Table 3.Maximum Likelihood Tree aLRT Clade Support byv2Test of Significance

aLRT Clade Support Range

nuoK(300 bp) rplF(530 bp) tauB(763 bp) add(999 bp) araB(1,698 bp)

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

0%–10% 0 — 0 — 0 — 0 — 0 —

11%–20% 11 0.818 9 1.000 8 1.000 5 1.000 4 1.000

21%–30% 6 0.833 3 1.000 6 1.000 6 1.000 3 1.000

31%–40% 8 0.875 7 1.000 10 0.800 4 1.000 3 1.000

41%–50% 15 0.800 10 0.900 3 1.000 5 1.000 5 0.800

51%–60% 11 0.818 6 0.833 8 1.000 7 0.857 6 1.000

61%–70% 16 0.750 16 1.000 8 1.000 11 0.818 2 0.500

71%–80% 21 0.905 9 1.000 14 0.929 20 0.950 5 1.000

81%–90% 30 0.900 21 0.905 12 0.917 10 0.900 8 1.000

91%–100% 492 0.907 529 0.940 541 0.950 542 0.958 574 0.976

0%–40% 25 0.840 19 1.000 24 0.917 15 1.000 10 1.000

81%–100% 522 0.906 550 0.938 553 0.949 552 0.956 582 0.976

Mean aLRT6SE 91.91%60.75% 94.36%60.65% 94.72%60.66% 95.33%60.59% 97.38%60.49%

Mean accuracy 89.51% 94.09% 94.91% 95.41% 97.23%

Slope6SEa 0.09

60.07 0.0760.08 0.0360.09 0.1360.08 0.0760.23

p-Valueb 0.20 0.41 0.74 0.17 0.79

a

Slope of midpoint of clade support range versus fraction of true clades.

b

Thep-value is the probability that the slope is not different from zero. doi:10.1371/journal.pcbi.0030051.t003

PLoS

Computational

Biology

|

www.plos

compbiol.or

g

March

2007

|

Volume

3

|

Issue

3

|

e51

0476

Clade

Support

Unrelated

to

(5)

Table 4.Maximum Likelihood Trees aLRT Clade Support by Minimum of SH-Like andv2Test of Significance

aLRT Clade Support Range

nuoK(300 bp) rplF(530 bp) tauB(763 bp) add(999 bp) araB(1,698 bp)

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

0%–10% 10 0.800 8 1.000 9 0.889 6 1.000 4 1.000

11%–20% 11 0.818 6 1.000 8 1.000 6 1.000 2 1.000

21%–30% 11 0.727 5 1.000 7 1.000 7 1.000 5 1.000

31%–40% 5 0.800 6 1.000 10 1.000 5 0.800 5 1.000

41%–50% 12 0.917 12 0.917 6 0.833 8 0.875 4 0.500

51%–60% 12 0.833 5 1.000 4 1.000 14 0.857 6 1.000

61%–70% 15 0.733 19 1.000 2 1.000 8 1.000 3 1.000

71%–80% 41 0.951 22 1.000 22 0.909 18 0.944 7 1.000

81%–90% 92 0.891 61 0.934 46 0.913 32 0.938 26 1.000

91%–100% 401 0.908 466 0.951 496 0.950 506 0.929 548 0.974

0%–40% 37 0.838 25 1.000 34 1.000 24 0.958 16 1.000

81%–100% 493 0.904 527 0.949 542 0.946 538 0.929 574 0.975

Mean aLRT6SE 86.88%60.86% 91.02%60.75% 91.85%60.73% 92.46%60.75% 95.71%60.58%

Mean accuracy 89.51% 95.41% 94.59% 92.95% 97.38%

Slope6SEa 0.14

60.07 0.0560.03 0.0260.07 0.0460.08 0.0260.18

p-Valueb 0.10 0.18 0.77 0.58 0.93

a

Slope of midpoint of clade support range versus fraction of true clades.

b

Thep-value is the probability that the slope is not different from zero. doi:10.1371/journal.pcbi.0030051.t004

Table 5.nuoK(300 bp) Trees Based on 16 Random Sequences Sampled from 128 Evolved Sequences

aLRT Clade Support Range

Maximum Likelihood Bootstrap Maximum Likelihood aLRT (Minimum of SH-Like andv2)

Maximum Likelihood aLRT (v2) Bayesian PPs

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

0%–10% 0 — 5 0.800 1 1.000 0 —

11%–20% 2 0.500 2 1.000 3 1.000 0 —

21%–30% 2 1.000 3 1.000 3 0.667 1 1.000

31%–40% 2 1.000 4 0.250 4 0.750 3 0.667

41%–50% 8 0.625 7 0.857 6 0.667 6 0.833

51%–60% 13 0.769 3 0.667 4 0.750 8 0.875

61%–70% 12 0.750 7 0.857 7 0.857 4 0.750

71%–80% 12 0.917 8 1.000 3 1.000 12 0.917

81%–90% 16 0.812 28 0.893 12 0.833 16 0.938

91%–100% 63 0.841 63 0.841 87 0.816 80 0.850

0%–40% 6 0.833 14 0.786 11 0.818 4 0.750

81%–100% 79 0.835 91 0.857 99 0.818 96 0.864

Mean clade support6SE 80.37%61.90% 79.36%62.30% 85.22%62.14% 87.37%61.62%

Mean accuracy 81.54% 84.62% 81.54% 86.15%

Slope6SEa 0.1360.22 0.0860.26 0.0560.15 0.0560.18

p-Valueb 0.59 0.78 0.77 0.78

a

Slope of midpoint of clade support range versus fraction of true clades.

b

Thep-value is the probability that the slope is not different from zero.

Simulation conditions were as for Tables 1–4 except that the average branch length was 0.15 base substitutions per site. doi:10.1371/journal.pcbi.0030051.t005

PLoS

Computational

Biology

|

www.plos

compbiol.or

g

March

2007

|

Volume

3

|

Issue

3

|

e51

0477

Clade

Support

Unrelated

to

(6)

Our results, however, differ strikingly from those of [13] and [14] with respect to the correspondence between individual clade confidences and the accuracy of those clades. Alfaro et al. [13] found that for most topologies accuracy was higher than either BP or PP when clade confidences were greater than;40%, but lower than clade confidences when those measures were less than ;40%. Hillis and Bull [10] obtained similar results for BP, and Wilcox et al. [14] obtained similar results for PP, but a low crossover point at about 20% for BP. In contrast, we find no significant correlation between individual clade supports and the probability that a clade is correct, whichever method of clade support is used. We regressed the fraction of clades that actually exist within each decile against the midpoint for each decile of clade support; thus, a slope approaching one would be expected for a perfect correlation between those values, whereas a slope of zero would indicate no correlation. Only two out of the 28 plots have slopes that are significantly different from zero (genes nuoK and rplF for bootstrap support of maximum likelihood trees, withp¼0.03 and 0.02,

Figure 1.Typical Phylogenetic Trees

(A) True 64-taxon tree initiated withnuoKsequences. The arrow indicates a near trichotomy.

(B) Bayesian tree estimated from the same data as in (A). Numbers are posterior probabilities of clades whose posterior probabilities are,90%. Arrows indicate the clades that do not exist in the true tree.

doi:10.1371/journal.pcbi.0030051.g001

Figure 2.Topology of a Typical Unbalanced Tree of 16 Taxa

doi:10.1371/journal.pcbi.0030051.g002

PLoS Computational Biology | www.ploscompbiol.org 0478 March 2007 | Volume 3 | Issue 3 | e51

(7)

Table 7.Mean Clade Support versus Mean Accuracy

Mean Clade Supporta

Mean Accuracy of the Estimated Treeb

Maximum Likelihood Bootstrap Support

Maximum Likelihood aLRT Support (Minimum ofv2and SH-Like)

Maximum Likelihood aLRT Support (v2)

Bayesian Inference PPs

Slope6SE 0.6460.19 0.78860.11 1.2660.12 0.6760.03

p-Value 0.019 0.0018 0.0001 ,0.0001

Correlation coefficient 0.83 0.97 0.99 0.99

80% 87% 85% 76% 85%

85% 90% 89% 82% 89%

90% 93% 93% 88% 92%

95% 96% 97% 95% 96%

100% 100% 101% 101% 99%

a

Clade support in linear regression of estimated tree accuracy against clade support.

b

The regressed (best-fit) tree accuracy for the clade support value in the first column calculated from the intercept and the slope above.

Consider the 87% value in the second column: when, in a plot of accuracy of the estimated tree against bootstrap support, the bootstrap support was 80%, the linear regression best-fit value of estimated tree accuracy was 87%. doi:10.1371/journal.pcbi.0030051.t007

Table 6.nuoK(300 bp) Unbalanced Trees Based on 16 Sequences Sampled from 128 Evolved Sequences

aLRT Clade Support Range

Maximum Likelihood Bootstrap Maximum Likelihood aLRT (Minimum of SH-Like andv2)

Maximum Likelihood aLRT (v2) Bayesian PPs

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

Number of Clades

Fraction True Clades

0%–10% 0 — 0 — 0 — 0 —

11%–20% 5 0.600 1 0.000 0 — 0 —

21%–30% 2 1.000 2 1.000 1 1.000 4 1.000

31%–40% 3 1.000 6 1.000 5 0.800 6 0.500

41%–50% 4 1.000 5 1.000 5 1.000 5 1.000

51%–60% 8 1.000 4 1.000 5 1.000 2 1.000

61%–70% 5 1.000 6 1.000 6 1.000 3 1.000

71%–80% 9 1.000 7 0.857 2 1.000 7 1.000

81%–90% 7 0.714 14 0.929 9 0.889 7 1.000

91%–100% 87 0.920 85 0.882 97 0.887 96 0.906

0%–40% 10 0.800 9 0.889 6 0.833 10 0.700

81%–100% 94 0.904 99 0.888 106 0.887 103 0.912

Mean clade support6SE 85.15%62.05% 85.7%61.85% 89.63%61.72% 88.65%61.89%

Mean accuracy 91.54% 90.0% 90.0% 90.77%

Slope6SEa 0.07

60.21 0.5160.41 0.0461.03 0.2260.28

p-Valueb 0.74 0.25 0.76 0.47

a

Slope of midpoint of clade support range versus fraction of true clades.

b

P is the probability that the slope is not different from zero.

Simulation conditions were as for Tables 1–4 except that the average branch length was 0.15 base substitutions per site. doi:10.1371/journal.pcbi.0030051.t006

PLoS

Computational

Biology

|

www.plos

compbiol.or

g

March

2007

|

Volume

3

|

Issue

3

|

e51

0479

Clade

Support

Unrelated

to

(8)

respectively). One of these slopes is slightly positive, and the other is slightly negative; thus, both likely represent outliers. The absence of correlation between clade support and the likelihood that a clade exists means that, whatever the method, clade support values provide no information about, and have no predictive power as to, the likelihood that the clade exists. We attribute the differences between our results and those of others [10,13,14] to our use of a more biologically realistic simulation.

It is conceivable that this startling finding is the result of reconstructing relatively large (64 taxon) trees under some false assumption that is unknowingly incorporated into the simulation: perhaps with so many taxa, resulting in an enormous number of possible trees, any clade that has even mild support is likely to be a true clade. We think this possibility is unlikely, because the fraction of false clades that have 81%–100% support is roughly the same as the fraction of false clades that have low (0%–40%) support. Nevertheless, to test the idea that our findings are an artifact of considering large trees, we tested all four methods with 16-taxon datasets ofnuoK,replicated ten times each. (ThenuoKgene was chosen because it is the shortest gene and the gene for which all methods were the least accurate.) As before, the taxa were a random sample from a 128-taxon dataset. When the number of taxa was reduced from 64 to 16, the quality of the alignments was reduced so that the amino acid identity was ,20%, below the zone in which alignments are reliable. The average branch length was therefore reduced from 0.18 to 0.15 substitutions per site to produce alignments that exhibited an average of 24% amino acid identity, well within the‘‘twilight zone’’[19]. Table 5 shows that the results are essentially the same as in Tables 1–4: there is no significant correlation between clade support and the fraction of true clades. Both mean accuracy and mean clade support are generally lower than in Tables 1–4, but it remains the case that clade support provides no information about the likelihood that a clade actually exists.

Random sampling of taxa typically results in well-balanced trees (Figure 1), and it is conceivable that our findings apply only to trees with similar topology. To test that possibility, we nonrandomly sampled tennuoK datasets to generate highly pectinate, unbalanced 16-taxon trees (Figure 2). For the unbalanced trees, the topological accuracies were higher than for the 16-taxon balanced trees, and all methods of clade support again underestimated that accuracy. Again, there was no significant correlation between clade support and the likelihood that a clade existed on the true tree (Table 6).

We conclude that our results are general, and not simply attributable to large trees or to balanced trees.

This is one of those good news–bad news stories beloved by comedians. The bad news is that none of the methods of assessing clade support provides any reliable estimation that the clade has been correctly assigned. The good news is that, even with data that are near the practical limits for phylogenetic tree reconstruction, both maximum likelihood and the Bayesian method estimate topologies so well that even when clade support is very low there is a better than 80% chance that the clade is correctly assigned.

In addition, for each method, averaged over the 70 datasets, there is a significant correlation between the average branch support and the accuracy of the tree (Table 7), and for all methods except nonparametric bootstrap, the average clade

support value is a good, if slightly conservative, estimator of the overall fraction of clades that actually exist. It might be argued that having a good estimate of overall accuracy is not very useful, and that we are generally interested in identifying unreliable branches. Our results show that we simply cannot identify what particular branches are unreliable based on measures of clade support. Thus, with current methods of determining clade credibility we cannot have what we might generally want, and it is important to acknowledge the limitations of those metrics. On the other hand, methods of determining clade confidence do provide a good estimation of the overall reliability of a phylogenetic tree, and permit us to infer how many untrustworthy branches may be present. Just as we can make good predictions about the diffusion of a mass of molecules over time, but not about the motions of individual molecules in that mass, we can make good estimations of the overall topological accuracy of a tree, but not about the accuracy of individual branches.

Materials and Methods

Simulations.Simulations were performed by EvolveAGene [18,21]. Five coding sequences fromE. coliK12 were selected from theE. coli genome entirely on the basis of length, and used to initiate the simulations:nuoK,300 bp, encodes the NADH dehydrogenase subunit K; rplF, 530 bp, encodes 50S ribosomal protein L6;tauB, 763 bp, encodes a taurine transport ATP-binding protein; add, 999 bp, encodes adenosine deaminase; andaraB,1,698 bp, encodes ribuloki-nase. The genes are not functionally related to each other, and none exhibits detectable homology to another by pairwise BLAST comparisons. For the simulations in Tables 1–4, the average branch length was 0.18 substitutions per site, with lengths ranging from 0 to 0.36 substitutions per site; for Tables 5 and 6 the average branch length was 0.15 substitutions per site, ranging from 0 to 0.30 substitutions per site. The tree was evolved for seven‘‘generations’’

to give 128 terminal taxa; thus, the average length from the root to the tip was 1.26 substitutions per site. The probability of accepting an indel was 0.02, and the probability of accepting a nonsynonymous base substitution was 0.2. Ten independent simulations were carried out from each of the five root sequences.

When all of the terminal sequences descended from the root node were included in the dataset, both Bayesian and maximum likelihood trees included very few nodes with clade confidences ,80% (unpublished data). When trees were based on a random subsample of the sequences, both methods produced trees with more low-confidence clades. Datasets used to estimate trees were therefore based on a random sample of 16 or 64 of the 128 evolved sequences. Alignments. Sequences were aligned by ClustalW [24] as imple-mented by MEGA 3.1 [25]. Sequences were translated to their corresponding protein sequences by MEGA 3.1, aligned with a gap-opening penalty of 3.0 and a gap-extension penalty of 1.8. The average pairwise amino acid identities in the resulting alignment were typically 21%–22%, near the lower boundary of the‘‘twilight zone’’below which alignments are not sufficiently reliable to produce valid phylogenetic trees [19,20]. Triplet gaps corresponding to the gaps in the protein alignment were introduced back into the DNA sequences by MEGA 3.1. The resulting DNA sequence alignments were saved in the FASTA format and converted to the PHYLIP format (for input to PHYML) and to the Nexus format (for input to MrBayes) by a Perl script.

Estimation of phylogenetic trees. Trees were estimated by two methods: maximum likelihood as implemented by PHYML 2.4.4 [3], and the Bayesian method as implemented by MrBayes 3.1.2 [9]. In both cases, trees were estimated using the GTRþinvariantsþgamma model.

For maximum likelihood trees, clade confidences were estimated from 100 bootstrap replicates.

Bayesian trees were estimated from 600,000 generations, sampling every 100 generations, with a heating parameter of 0.15, in two parallel runs. The consensus trees were calculated using the allcompat option (strict consensus) from the final 4,501 trees of each run. Convergence, as judged by the diagnostic average standard deviation of the split frequencies between two parallel runs falling below 0.02, typically occurred before generation 120,000 (1,200 trees).

PLoS Computational Biology | www.ploscompbiol.org 0480 March 2007 | Volume 3 | Issue 3 | e51

(9)

A typical true tree, in this case initiated with thenuoKsequence, is shown in Figure 1A. Note that the true tree includes one near trichotomy, and that the distance from the root to the tips ranges from 0.52 substitutions per site for taxon ZZZZZZZ to 1.55 substitutions per site for taxon PPPPPPPPPP. The corresponding Bayesian estimated tree is shown in Figure 2. Posterior probabilities ,90% are indicated. This Bayesian tree is typical in that only 12 of the 61 interior nodes have posterior probabilities,90%, but only four of those low-PP clades (indicated by arrows) are not present in the true tree.

Calculation of topological accuracy. A Perl script, InferAcc, was used to compare the estimated trees with the true trees. Clades were sorted into bins as indicated in Tables 1–6, and the clade was scored as existing if it was present in the true tree. Mean accuracy is the

fraction of clades in the estimated tree that exist in the true tree averaged over the ten trees in the set.

Acknowledgments

We are grateful to reviewer number two for very helpful comments. Author contributions. BGH and SJS conceived and designed the experiments and analyzed the data. BGH performed the experiments and contributed reagents/materials/analysis tools. BGH and SJS wrote the paper.

Funding.SJS is supported by the Poncin Scholarship Fund. Competing interests.The authors have declared that no competing interests exist.

References

1. Felsenstein J (1985) Confidence limits on phylogenies: An approach using the bootstrap. Evolution 39: 783–791.

2. Anisimova M, Gascuel O (2006) Approximate likelihood-ratio test for branches: A fast, accurate, and powerful alternative. Syst Biol 55: 539–552. 3. Guindon S, Gascuel O (2003) A simple, fast, and accurate algorithm to estimate large phylogenies by maximum likelihood. Syst Biol 52: 696–704. 4. Hall BG (2004) Phylogenetic trees made easy: A how-to manual. Sunderland

(Massachusetts): Sinauer Associates 221 p.

5. Huelsenbeck JP, Larget B, Miller RE, Ronquist F (2002) Potential applications and pitfalls of Bayesian inference of phylogeny. Syst Biol 51: 673–688.

6. Mau B, Newton M, Larget B (1999) Bayesian phylogenetic inference via Markov chain Monte Carlo methods. Biometrics 55: 1–12.

7. Rannala B, Yang ZH (1996) Probability distribution of molecular evolu-tionary trees: A new method of phylogenetic inference. J Mol Evol 43: 304– 311.

8. Huelsenbeck JP, Ronquist F (2001) MrBayes: Bayesian inference of phylogeny. Bioinformatics 17: 754–755.

9. Ronquist F, Huelsenbeck JP (2003) MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 19: 1572–1574.

10. Hillis DM, Bull JJ (1993) An empirical test of bootstrapping as a method for assessing confidence in phylogenetic analysis. Syst Biol 42: 182–192. 11. Suzuki Y, Glazko GV, Nei M (2002) Overcredibility of molecular

phylogenies obtained by Bayesian phylogenetics. Proc Natl Acad Sci U S A 99: 16138–16143.

12. Taylor DJ, Piel WH (2004) An assessment of accuracy, error, and conflict with support values from genome-scale phylogenetic data. Mol Biol Evol 21: 1534–1537.

13. Alfaro ME, Zoller S, Lutzoni F (2003) Bayes or bootstrap? A simulation study comparing the performance of Bayesian Markov chain Monte Carlo sampling and bootstrapping in assessing phylogenetic confidence. Mol Biol Evol 20: 255–266.

14. Wilcox TP, Zwickl DJ, Heath TA, Hillis DM (2002) Phylogenetic relation-ships of the dwarf boas and a comparison of Bayesian and bootstrap measures of phylogenetic support. Mol Phylogenet Evol 25: 361–371. 15. Douady CJ, Delsuc F, Boucher Y, Doolittle WF, Douzery EJ (2003)

Comparison of Bayesian and maximum likelihood bootstrap measures of phylogenetic reliability. Mol Biol Evol 20: 248–254.

16. Zharkikh A, Li W-H (1992) Statistical properties of bootstrap estimation of phylogenetic variability from nucleotide sequences: I. Four taxa with a molecular clock. Mol Biol Evol 9: 9: 1119–1147.

17. Kimura M (1980) A simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide sequences. J Mol Evol 16: 111–120.

18. Hall BG (2005) Comparison of the accuracies of several phylogenetic methods using protein and DNA sequences. Mol Biol Evol 22: 792–802. 19. Doolittle RF (1981) Similar amino acid sequences: Chance or common

ancestry? Science 214: 149–159.

20. Thompson JD, Plewniak F, Poch O (1999) A comprehensive comparison of multiple sequence alignment programs. Nucleic Acids Res 27: 2682–2690. 21. Hall BG (2006) Simple and accurate estimation of ancestral protein

sequences. Proc Natl Acad Sci U S A 103: 5431–5436.

22. Jukes TH, Cantor CR (1969) Evolution of protein molecules. In: Munro HN, editor. Mammalian protein metabolism. New York: Academic Press. pp. 21– 32.

23. Nei M, Kumar S (2000) Molecular evolution and phylogenetics. New York: Oxford University Press. 333 p.

24. Thompson JD, Higgins DG, Gibson TJ (1994) CLUSTAL W: Improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Res 22: 4673–4680.

25. Kumar S, Tamura K, Nei M (2004) MEGA3: Integrated software for molecular evolutionary genetics analysis and sequence alignment. Brief Bioinform 5: 150–163.

PLoS Computational Biology | www.ploscompbiol.org 0481 March 2007 | Volume 3 | Issue 3 | e51

Referências

Documentos relacionados

The iterative methods: Jacobi, Gauss-Seidel and SOR methods were incorporated into the acceleration scheme (Chebyshev extrapolation, Residual smoothing, Accelerated

Dated or calibrated phylogenetic trees, in which branch lengths correspond to evolutionary divergence times, are important requirements for computing measures of

Based on data from research stations and from industry, the Australian beef genetic improvement scheme, BREEDPLAN, started providing estimated breeding values (EBVs) for RFI

didático e resolva as ​listas de exercícios (disponíveis no ​Classroom​) referentes às obras de Carlos Drummond de Andrade, João Guimarães Rosa, Machado de Assis,

Therefore, multiple genes should be targeted for an aDNA analysis; (3) exploratory analysis of available ex- tant taxa can provide a priori clues as to which gene regions to

social assistance. The protection of jobs within some enterprises, cooperatives, forms of economical associations, constitute an efficient social policy, totally different from

Abstract: As in ancient architecture of Greece and Rome there was an interconnection between picturesque and monumental forms of arts, in antique period in the architecture

O objectivo deste trabalho foi estabelecer a relação entre o tipo de enzima e a extensão da reacção de hidrólise e as características dos hidrolisados proteicos obtidos a partir de