ArticlePDF Available

The shape of mammalian phylogeny: Patterns, processes and scales

The Royal Society
Philosophical Transactions B
Authors:

Abstract and Figures

Mammalian phylogeny is far too asymmetric for all contemporaneous lineages to have had equal chances of diversifying. We consider this asymmetry or imbalance from four perspectives. First, we infer a minimal set of 'regime changes'-points at which net diversification rate has changed-identifying 15 significant radiations and 12 clades that may be 'downshifts'. We next show that mammalian phylogeny is similar in shape to a large set of published phylogenies of other vertebrate, arthropod and plant groups, suggesting that many clades may diversify under a largely shared set of 'rules'. Third, we simulate six simple macroevolutionary models, showing that those where speciation slows down as geographical or niche space is filled, produce more realistic phylogenies than do models involving key innovations. Lastly, an analysis of the spatial scaling of imbalance shows that the phylogeny of species within an assemblage, ecoregion or larger area always tends to be more unbalanced than expected from the phylogeny of species at the next more inclusive spatial scale. We conclude with a verbal model of mammalian macroevolution, which emphasizes the importance to diversification of accessing new regions of geographical or niche space.
Content may be subject to copyright.
Research
The shape of mammalian phylogeny:
patterns, processes and scales
Andy Purvis1,2,*, Susanne A. Fritz3, Jesu
´s Rodrı
´guez4,
Paul H. Harvey5and Richard Grenyer2,6
1
Department of Life Sciences, Imperial College London, Silwood Park, Ascot SL5 7PY, UK
2
NERC Centre for Population Biology, Imperial College London, Silwood Park,
Ascot SL5 7PY, UK
3
Department of Biology, Centre for Macroecology, Evolution and Climate, University of Copenhagen,
2100 Copenhagen, Denmark
4
Centro Nacional de Investigacio
´n de la Evolucio
´n Humana (CENIEH), 09002 Burgos, Spain
5
Department of Zoology, University of Oxford, South Parks Road, Oxford OX1 3PS, UK
6
School of Geography and the Environment, University of Oxford, South Parks Road,
Oxford OX1 3QY, UK
Mammalian phylogeny is far too asymmetric for all contemporaneous lineages to have had equal
chances of diversifying. We consider this asymmetry or imbalance from four perspectives. First,
we infer a minimal set of ‘regime changes’—points at which net diversification rate has changed—
identifying 15 significant radiations and 12 clades that may be ‘downshifts’. We next show that
mammalian phylogeny is similar in shape to a large set of published phylogenies of other vertebrate,
arthropod and plant groups, suggesting that many clades may diversify under a largely shared set of
‘rules’. Third, we simulate six simple macroevolutionary models, showing that those where speciation
slows down as geographical or niche space is filled, produce more realistic phylogenies than do models
involving key innovations. Lastly, an analysis of the spatial scaling of imbalance shows that the
phylogeny of species within an assemblage, ecoregion or larger area always tends to be more unba-
lanced than expected from the phylogeny of species at the next more inclusive spatial scale. We
conclude with a verbal model of mammalian macroevolution, which emphasizes the importance to
diversification of accessing new regions of geographical or niche space.
Keywords: macroevolution; phylogenetic imbalance; evolutionary radiations; simulation;
stochastic model; community phylogenetics
1. INTRODUCTION
As our ability to reconstruct phylogenies has devel-
oped, it has become clear that the Tree of Life’s
structure is much less regular than most early depic-
tions: proliferation has been much more rapid in
some branches and at some times than others
(reviewed by [13]). In mammals, for instance, a
recent taxonomy listed 2277 species of rodent and
1166 species of bat, but only one aardvark, two
flying lemurs and three elephants [4]—even though
all these clades are tens of millions of years old.
There is a long history of both deterministic and sto-
chastic explanations for such diversity differences,
both in general (e.g. [57]) and in mammals (e.g.
[6,8,9]).
Phylogenies of present-day species provide two
valuable—though still incomplete—lines of evidence
that can be analysed to gain more precise insight into
macroevolution [13]. First, analysing the temporal
spacing of nodes can reveal changes in diversification
rates over time. Such an analysis of mammalian phylo-
geny found evidence for a pulse of early radiation
about 100 85 Ma, followed by low rates until about
10 Myr after the cretaceous paleogene (or K– Pg)
boundary, when diversification again sped up [10].
However, analyses of nodes ages depend on the ade-
quacy of the methods used to estimate dates—still a
contentious area (e.g. [11,12])—and are unlikely to
estimate extinction history accurately [1315].
In this paper, we therefore focus on the second line
of evidence—the phylogeny’s asymmetry or imbal-
ance. If two sister clades have diversified to very
different extents, their species have probably had
different underlying probabilities of diversifying [16].
Measures of imbalance seek to capture the difference
in species-richness between sister clades [1], and are
commonly used to test whether contemporaneous
species could all have had the same chances of diversi-
fication, a null model known as the equal-rates Markov
(ERM) model. We begin by analysing the overall
imbalance of mammalian phylogeny before using a
*Author for correspondence (a.purvis@imperial.ac.uk).
One contribution of 12 to a Theme Issue ‘Global biodiversity
of mammals’.
Phil. Trans. R. Soc. B (2011) 366, 2462–2477
doi:10.1098/rstb.2011.0025
2462 This journal is q2011 The Royal Society
range of empirical and simulation approaches to try to
understand mammalian macroevolution.
2. THE SHAPE OF MAMMALIAN PHYLOGENY
We used a modification [17] of Bininda-Emonds et al.’s
[10] phylogeny because it adopts a more recent taxon-
omy [4]. We excluded four domesticated species,
generated by artificial rather than natural selection,
from all of our analyses. The resulting tree has 925
bifurcations with enough present-day descendants to
be informative about imbalance. Importantly, it con-
tains the great majority (5016/5416) of the recognized
species: the imbalance of highly incomplete phylogenies
will be an unbiased estimate of the true imbalance
if sampling is random, but strict randomness may
be unlikely [18]. Wilkinson et al.[19] noted that the
method used to construct the supertree is biased with
respect to tree shape, tending to favour more unba-
lanced topologies. We therefore also analysed two
near-complete, well-resolved, species-level ‘superma-
trix’ Bayesian phylogenies of orders—Carnivora [20]
and Cetartiodactyla [21]: comparing the imbalance of
these trees with the relevant portions of the super-
tree provides a preliminary assessment of whether the
supertree’s imbalance could be artifactual.
We characterized imbalance in two ways. The first is
a purely phenomenological measure of pattern. We
used Fusco & Cronk’s [22] approach to compute an
imbalance score, I, for each of the 925 informative
nodes, according to
I¼Bm
Mm;
where Bis the number of species in the more diverse
sister clade, mis the smallest value Bcould take and
Mthe largest possible value for B. These Ivalues
range from zero (as symmetric as possible) to one (com-
pletely asymmetric). We applied weights as in Purvis
et al.[23] such that, under ERM, the weighted mean I
(hereafter, I
w
) has an expectation of 0.5 for any node
size (i.e. total number of extant descendant species).
The empirical I
w
for the whole set of nodes was com-
pared with the null expectation using a weighted
t-test. We likewise tested marsupials, all placentals and
four placental superorders separately. On rejecting
ERM, we then constructed imbalance signatures to pro-
vide a richer description of imbalance; we grouped
nodes into 10 bins of equal width on the log(node
size) axis between the clade’s minimum and maximum
node size, computed I
w
within each bin and regressed I
w
against mean log(node size) among bins. Although the
nodal Ivalues have a strongly non-normal distribution,
use of within-bin I
w
gives regressions a much more nor-
mally distributed error term. In these regressions, node
size might be acting as a proxy for node age—whatever
processes are causing imbalance can operate for longer
in older clades. We tested this possibility using weighted
multiple regression of nodal Ion log(node size) and
log(clade age). The R package caper (C. D. L. Orme
2011, unpublished data, http://caper.r-forge.r-project.
org/) was used to compute nodal imbalance scores.
We also used a process-based measure of imbalance,
namely the estimate of
b
from Aldous’s [24]
b
-split
model of clade growth. The range of
b
is from – 2 to
þ1; trees where
b
,0 are more unbalanced than
average ERM trees, and those with
b
.0aremore
balanced. Two trees that grew under the same
b
-split
model should yield very similar
b
estimates. Aldous
[24] was motivated by mathematical simplicity rather
than biological considerations, and biological meaning
of particular
b
values remains opaque [25]. Nonethe-
less, a
b
of approximately – 1 has been found to
provide a good fit to a wide range of (incomplete) pub-
lished phylogenies [25], perhaps implying a common
macroevolutionary process. At the very least,
b
reflects
tree shape pattern differently from I
w
.Toestimate
b
,we
modified the function maxlik.betasplit in the R package
apTreeshape [26] in two ways. First, we removed the
requirement for a fully resolved phylogeny. (We used
simulations, not shown, to check that estimates of
b
were not systematically biased by randomly collapsing
branches in ERM phylogenies.) Second, to avoid
underflow errors, we amended one of the internal func-
tions to work with the log(beta) distribution rather than
the beta distribution. We then optimized
b
for the
whole tree and for the same subtrees and binned sets
of nodes as identified above.
Table 1. Estimates of I
w
and
b
for portions of mammalian phylogeny. n¼number of informative nodes. Significance of
departure from the equal-rates Markov expectation is assessed by weighted t-test for I
w
and from 1000 parametric bootstrap
replicates for
b
. 95% CI, confidence interval for
b
from bootstrapping.
clade nI
w
+s.e.
b
95% CI
Mammalia 925 0.631 +0.013*** 20.857*** 20.974 to 20.710
Marsupialia 110 0.650 +0.038*** 20.904*** 21.215 to 20.392
Eutheria 812 0.627 +0.014*** 20.847*** 20.985 to 20.684
Afrotheria 15 0.657 +0.090 20.878 21.671 to 6.113
Xenarthra 7 0.420 +0.160 7.264 22.000 to .500
Laurasiatheria 398 0.603 +0.020*** 20.765*** 20.955 to 20.498
Euarchontoglires 389 0.653 +0.020*** 20.938*** 21.110 to 20.731
Carnivora: supertree 108 0.590+0.040 20.752*** 21.104 to 20.137
Carnivora: cyt b tree 140 0.672 +0.033 21.102*** 21.276 to 20.855
Cetartiodactyla: supertree 69 0.616 +0.032 20.770 21.216 to 0.014
Cetartiodactyla: cyt b tree 148 0.662 +0.032 21.002*** 21.186 to 20.732
*p,0.05; **p,0.01, ***p,0.001.
The shape of mammalian phylogeny A. Purvis et al. 2463
Phil. Trans. R. Soc. B (2011)
Both I
w
and
b
indicate that chances of diversifica-
tion have varied among contemporaneous species
(table 1; first row). The values of I
w
and
b
are very
similar for marsupials and placentals separately, and
for three of the four placental superorders (table 1),
the exception being Xenarthra, which had only seven
informative nodes. The cyt b phylogenies were slightly
more unbalanced than the corresponding portions of
the supertree (table 1), suggesting that the supertree’s
imbalance is not artifactual.
The imbalance signature in figure 1ashows that I
w
increases with node size across the whole phylogeny.
The regressions of I
w
on log(node size) did not differ
significantly among placental superorders (omitting
Xenarthra on grounds of sample size; F
2,21
¼1.67, p.
0.2), but placental and marsupial imbalance signatures
do differ significantly (F
1,31
¼4.83, p¼0.04). However,
a clearer distinction is that I
w
increases more with
log(node size) in the two more diverse placental super-
orders (Laurasiatheria and Euarchontoglires) than
in Afrotheria or Marsupialia (F
1,31
¼6.25, p¼0.02;
figure 1c). ANCOVA showed no differences between
the imbalance signatures of cyt b tree and supertree
for either order (Carnivora: F
2,244
¼1.30, p.0.2;
Cetartiodactyla: F
2,213
¼0.61, p.0.5).
Multiple regression suggests that node age does not
confound the relationship seen in figure 1a: nodal I
does not depend on log(node age) (t
922
¼21.085,
p.0.2) but increases weakly with log(node size)
(slope ¼0.024, t
923
¼2.418, p¼0.02). Although the
two predictors are correlated, variance inflation factors
were less than 2, indicating no strong collinearity. Con-
trary to expectation,
b
shows a weak tendency ( p¼
0.06) to become less negative—closer to ERM—for
larger nodes (figure 1b).
The finding that mammalian phylogeny is too unba-
lanced for ERM is unsurprising: analyses of large
individual trees (e.g. [27,28]) and collections of smal-
ler trees (e.g. [18,29 31]) have repeatedly shown that
complete or nearly complete empirical phylogenies are
markedly too asymmetric for ERM. The estimate of
b
is significantly less negative than the value of 21
suggested by Blum & Francois [25] for their smaller,
incomplete trees, but very close to the estimate derived
from their largest fully bifurcating trees (20.89).
The dependence of I
w
on node size is in line with
some previous studies. Purvis & Agapow [31], analysing
a set of 61 phylogenies across a wide range of taxa, found
that nodes with a given number of higher taxa (e.g.
genera) descended from them were more unbalanced
on average than nodes having the same number of
species descended from them. Holman [32] showed
that I
w
increased roughly linearly with log(node size)
across the same set of phylogenies, and a similar pattern
was also reported for a collection of highly incomplete
trees [33]. Freckleton et al.[34] showed I
w
to rise with
log(node size) in carnivora but to decrease in New
World primates. An increase in I
w
with node size indi-
cates the presence of small basal clades; however,
clades are unlikely to persist for long at low numbers
under diversity independent dynamics. The frequency
of small, old clades (especially in Laurasiatheria and
Euarchontoglires) therefore suggests either that their
dynamics are diversity-dependent or that they are now
declining deterministically (having previously grown
deterministically). The surprising finding that—unlike
I
w
b
becomes closer to ERM expectations at deeper
nodes suggests that it is best not to view the value of
b
as a process parameter; it is not constant. A possible
mechanism is if many clades having a greatly reduced
diversification rate die out before they become old,
making the tree self-pruning.
The overall shape of mammalian phylogeny is com-
plex, reflecting a mixture of diversification rates. The
0
0.2
0.4
0.6
0.8
1.0
(a)
(b)
(c)
within-bin weighted mean I
−1.0
−0.5
0.0
0.5
1.0
within-bin b
234567
0
0.2
0.4
0.6
0.8
1.0
lo
g
(node size)
within-bin weighted mean I
Figure 1. (a)I
w
signature for the whole mammalian phylogeny.
Points are within-bin I
w
; vertical lines indicate +1s.e. (Note
the large s.e. of the second point from the right, associated
with a small sample size.) Horizontal dashed line, ERM expec-
tation. Solid line is weighted regression, with sample sizes as the
weights: slope ¼0.025, t
8
¼2.573, p¼0.03. (b)
b
signature
for the whole tree. Points are within-bin estimates of
b
; vertical
lines are confidence intervals under a
x
2
approximation. (The
second point from the right is again unusual but uncertain.)
Horizontal dashed line, ERM expectation. Solid line is
weighted regression, with the inverse of the confidence interval
width as weights: slope ¼0.088, t
8
¼2.238, p¼0.06. (c)I
w
signature for each of four superorders. Points are within-bin
mean I
w
. Triangles, Marsupialia; diamonds, Afrotheria;
squares, Laurasiatheria; circles, Euarchontoglires. Dotted line
is regression through Marsupialia and Afrotheria
(slope ¼20.067, t
14
¼21.768, p¼0.1); solid line is regres-
sion through Laurasiatheria and Euarchontoglires (slope ¼
0.032, t
17
¼2.685, p¼0.02); regressions weighted as in (a).
2464 A. Purvis et al. The shape of mammalian phylogeny
Phil. Trans. R. Soc. B (2011)
next section is to identify clades that stand out as
unusual compared with the tree as a whole.
3. WINNERS AND LOSERS: PINPOINTING SHIFTS
IN DIVERSIFICATION RATE
Here, we distinguish between two kinds of deviation
from the tree-wide imbalance signature. Individual
nodes with unusual imbalance can suggest which
intrinsic and extrinsic phenomena may have regulated
diversification [16,22]. On the other hand, consistent
variation in imbalance—i.e. a whole subtree with
imbalance differing from the parental phylogeny—is
analogous to non-stationarity in models of spatial or
time-series data: it suggests that different processes
have been operating in different parts of the phylogeny.
To locate individual changes in diversification rate,
we used the D
1
statistic [35,36] as implemented in
apTreeshape [26]. D
1
considers two pieces of evidence
in deciding whether diversification rate has increased
along a branch: the imbalance of the ancestral node
and the imbalance of the descendant. Nodal imbal-
ance is expressed as the ratio between the likelihood
of the observed split under a single-rate ERM process,
and the likelihood of the same split if the larger daughter
clade had diversified with a 100-fold higher rate; larger
likelihood ratios therefore indicate higher imbalance. D
1
is simply the difference between the likelihood ratios of
the nodes at the start and the end of a branch; signifi-
cance is assessed by simulating a single-rate ERM
process. An upshifted descendant clade will tend to
make the parental clade appear to be upshifted too (a
phenomenon known as ‘trickle down’; [37]). In such
cases, both likelihood ratios will be large but the D
1
between them relatively small. Consequently (following
the contingency table in [36]), here we ascribe an
increase in diversification rate only to branches where
a significant D
1
is observed between an unbalanced
ancestral node and a more balanced descendant. Of
1335 bifurcating nodes in the supertree, 15 provided
such evidence of an increase in diversification rate.
These nodes are presented in table 2, and depicted
with an upward-facing black triangle on figure 2.
Moore et al.[37] argue that D
1
can only detect
increases in diversification rate, because clades under-
going a decrease are unlikely to survive to be observed.
We suggest that D
1
can also be used to pinpoint some
decreases in diversification rate (‘downshifts’), albeit
with caveats. For small clades, the likelihood under a
single-rate ERM process will be similar to that under
a heterogenous rate, as there is little evidence to differ-
entiate between the two: small clades have inherently
low likelihood ratios. As a result, a significant D
1
can
be found along a branch between an unbalanced
ancestral node and a smaller descendant. If imbalance
at the ancestral node cannot be ascribed to an increase
in rate on the branch leading to its larger daughter, and
if a significant D
1
is observed on the branch leading to
its smaller daughter, we contend this to suggest a
decrease in net diversification rate in the smaller
Table 2. Rate-shifted clades. MSW taxon ¼name of clade in Wilson & Reeder [4]. n¼number of species in clade, n
sister
¼
number of species in sister clade, n
ancs
¼number of species in sister group to ancestral node.
MSW taxon (3rd edn) age (Myr) nn
sister
n
ancs
node label
(a) Clades arising from significant ( p0.05) upshifts in diversification rate (100-fold shift in
l
sister
) as measured by D
1
[37]
‘Boreoeutheria’ 96.2 4614 29 69 8
Dipodomyinae þPerognathinae þsome Geomyidae (Rodentia) 35.9 96 2 1456 443
Sciurinae þXerinae þCalloscurinae (Rodentia; Sciuridae) 40.2 275 1 1 585
Sciurinae (Rodentia; Sciuridae) 18.2 80 4 7 681
within Sylvivagus (Lagomorpha; Leporidae) 8.3 10 1 2 809
Simiiformes 52.4 265 7 79 885
Carnivora 64.1 282 8 325 1330
Hipposideridae þRhinolophidae (Chiroptera) 47.5 147 16 5 1702
Artibeus (Chiroptera; Phyllostomidae) 8.6 17 1 26 1845
Molossus and allies (Chiroptera; Molossidae) 22.8 58 4 10 1898
Subgenus Sorex (Soricomorpha; Soricidae) 24 47 3 3 2162
Phalangeriformes þMacropodiformes (Diprotodontia) 45.8 122 4 1 2285
Macropus (Diprotodontia; Macropodidae) 10.1 13 1 2 2357
Dasyuridae (Dasyuromorphia) 25.4 63 1 2 2391
Didelphinae (Didelphimorphia; Didelphidae) 51.6 79 5 216 2464
(b) Clades tentatively identified as resulting from downshifts in diversification rate
Monotremata 63.6 4 5012 n/a 2
Anomaluromorpha (Rodentia) 54.9 9 1447 98 15
Castoridae (Rodentia) 12.1 2 96 1456 442
Sciurotamias (Rodentia; Sciuridae) 5 2 120 6 674
Dermoptera 12.8 2 351 20 1115
Giraffidae þAntilocaprinae (Cetartiodactyla) 20.2 3 189 8 1251
Camelidae (Cetartiodactyla) 21.8 3 306 16 1315
Harpionycteris (Chiroptera; Pteropodidae) 3.8 2 85 12 1617
Noctilionidae (Chiroptera) 3.5 2 158 1 1763
Desmodontinae (Chiroptera; Phyllostomidae) 26.8 3 147 8 1772
Solenodeontidae (Soricomorpha) 40.5 2 390 1669 2216
‘Afrotheria’ 89.5 69 4643 300 2236
The shape of mammalian phylogeny A. Purvis et al. 2465
Phil. Trans. R. Soc. B (2011)
daughter clade. Such downshifted clades are depicted
with downward triangles in figure 2 and table 2 gives
details. We caution that our evidence is only sugges-
tive, and note that our approach does not permit the
identification of single species as downshifts.
D
1
localizes rate shifts to an individual branch, but its
performance is yet to be evaluated in the context of gra-
dual changes across the phylogeny. These would
probably best be detected using the temporal spacing
of nodes, but the supertree’s incomplete resolution
and dating complicates such an analysis. Instead, we
used local variation in Aldous’
b
to paint a wider picture
of variation in imbalance across the supertree, and to
look for subtrees having consistent but unusual imbal-
ance signatures. Our motivation is that any whole-tree
imbalance statistic, while useful, is a measure of the
central tendency of imbalance for a set of nodes and
can therefore average away interesting biological
patterns. We therefore calculated the maximum-likeli-
hood estimate of
b
for every subtree in the phylogeny.
Several subtrees yield
b
values whose confidence
intervals do not overlap with those of the whole-tree
estimate. Primates are significantly more balanced
and sciurids significantly more unbalanced, as shown
by the coloured circles in figure 2 (red subtrees are
unbalanced, dark green are balanced). Figure 2 also
shows that variation in imbalance is not randomly
located around the phylogeny: many other clades have
imbalance signatures that differ consistently, although
not significantly, from the whole-tree imbalance (pink
and mint green subtrees have a
b
outside the 95% CI
for the whole-tree estimate). Whole-tree imbalance
measures do seem to obscure interesting variation,
although in mammals the effect size may be slight.
What mechanisms might underlie this variation
in imbalance across mammalian phylogeny? Compar-
ing the topologies of Primates and Sciuridae—the
clades with the most extreme balance and imbalance,
respectively—suggests that the answer could lie in
the interaction of historical biogeography and biotic
responses to global change. The differing geographical
nature of diversification in response to global change
has, we suspect, played a large part in generating the
observed differences in imbalance.
Crown-group Primates are old for their habitat. We
infer Simiiformes to be a significant radiation, at about
52.4 Ma, a timing that closely matches the first ap-
pearance of crown-group Primates (Euprimates) in
the fossil records of Europe, Asia and North America
at the onset of the Eocene [38,39]. However, by the
Mid-Eocene, Primates as a whole were beginning to
decline even as crown-group Simiiformes were diversi-
fying. Simiiformes radiated in a world that was cooling
and drying throughout the Eocene [40], as the area of
habitat similar to modern tropical forests was declining
[41] and possibly filling up with competitors such as
arboreal rodents [42]. New World primates, in par-
ticular, show an extremely balanced topology.
Repeated variation in the availability of tropical forest
has been suggested to govern diversification regimes
Figure 2. Diversification rate shifts localized on the supertree. Major clades are summarized with pictograms (clockwise from top:
Rodentia, ‘Marsupials’, ‘Afrotheria’, ‘Eulipotyphyla’, Chiroptera, Carnivora, Cetartiodactyla, Primates, Lagomorpha, Sciuridae).
Filled circles indicate subtrees with betas different from that of the whole tree (see text for details). More unbalanced: red
(non-overlapping CI) and pink (maximum-likelihood estimate of
b
for subtree outside whole-tree CI); more balanced: green
(non-overlapping CI) and light green (ML estimate of
b
for subtree outside whole-tree CI). Triangles indicate nodes with
significant (p0.05) upshifts as indicated by D
1
(upwards pointing) and putative downshifts (downwards pointing).
2466 A. Purvis et al. The shape of mammalian phylogeny
Phil. Trans. R. Soc. B (2011)
in South American vertebrates [43] and specifically
the topological balance of Primates (the ‘simultaneous
across-lineage vicariance’ of Heard & Cox [44]). We
suggest that the balanced phylogeny of Primates is a
consequence of such vicariant speciation within a
limited geographical region, throughout much of the
history of crown-group Primates (and speculate
that Xenarthra—also an unusually balanced clade
(table 1)—may have had a similar history). Towards
the end of the Eocene, sciurids experienced the flip
side of the same coin, as xeric habitats, savannah and
temperate forest all increased rapidly in area leading,
we speculate, to a diffusion of lineages through newly
available space and a resulting unbalanced topology.
Many of the diversification rate upshifts in table 2 can
be argued to correspond to sudden access to novel geo-
graphical or habitat space, although palaeontological
data are patchy, limiting the scope for meta-analysis.
Kisel et al.[45] present evidence that, in mammals,
colonization of new geographical areas usually elevates
diversification.
Downshifts in diversification rate are even harder
than upshifts to infer or date reliably, so interpretation
of the clades in table 2 is necessarily more speculative.
Limited geographical space, geographical isolation or
disproportionate extinction can all be expected to
lead to a depauperate extant clade [46,47]. In some
cases, extinction does appear to be a contributory
factor, either in deep time—Afrotherians, and perhaps
monotremes—or more recently—camelids, giraffids
(e.g. [48]) and solenodons [49]. Most of the down-
shifted clades are moderately widespread and some
(e.g. Castoridae and Camelidae) are very wide-ran-
ging. From an ecological perspective, many seem to
be highly unusual mammals: vampire and fishing
bats, high-altitude xeric artiodactyls, lodge-building
riparian rodents, toothless monotremes and non-
chiropteran gliders (including one highly balanced
clade in the otherwise universally unbalanced Sciuri-
dae). Unusualness is a statistical concept, of course,
and species-rich groups are bound to appear typical,
but specialization to a narrow adaptive zone may be
the best predictor of species-poverty. We speculate
that such specializations may reduce the likelihood of
radiation—because the niches are not broad enough
to support multiple species in sympatry—but at the
same time confer a strong incumbency advantage
[50] that reduces the per-species background extinc-
tion rate if niches are stable over time [51]. Such a
mechanism would lead to species in relictual groups
tending to be highly specialized, which is certainly
true for some of the species-poor basal clades (e.g.
Monotremata, Xenarthra).
If biogeographic history has been important in
shaping mammalian diversity [52], then it should be con-
sidered alongside ecological or life-history differences
when trying to understand clade-richness patterns [53].
It may also explain why ecological or life-history differ-
ences alone have so far explained little of the variance
in species-richness amongmammalian clades inphyloge-
netic comparative analyses. A test of whether dispersal
ability is associated with richness in mammals, as it is
in birds [54], would be of great interest.
4. IS MAMMALIAN PHYLOGENY UNUSUAL?
How common is the shape of mammalian phylogeny?
Previous studies suggest that tree shape may be con-
sistent (and consistently too unbalanced for ERM)
among groups as different as plants, insects and ver-
tebrates [29,30,55]. However, these studies had low
power: they all reduced each phylogeny’s shape to a
single number, and the largest compilation had 120
trees. More recent suggestions of a general shape for
phylogenies [25,56] have been based on incomplete
trees. To provide a more stringent test, we compare
the mammalian imbalance signature to that of a com-
pilation from the literature of 243 non-overlapping and
non-mammalian near-complete phylogenies. The
compilation includes all the non-mammalian trees
analysed previously by Purvis & Agapow [31] and
Holman [32], supplemented by some more recent
phylogenies [57,58]. All trees selected met the criteria
used in Purvis & Agapow [31], excepting that Bayesian
0.2
0.4
0.6
0.8
1.0 (a)(b)(c)
(d)(e)(f)
within-bin Iw
24681012
−2.0
−1.0
0
0.5
1.0
log (node size)
within-bin b
24681012
log (node size)
24681012
log (node size)
Figure 3. I
w
and
b
signatures for (a,d) plants, (b,e) arthropods and (c,f) non-mammalian vertebrates. Top row as in figure 1a.
Bottom row as in figure 1b. One point in flies well above the plotting region (
b
¼3.54), and is also an outlier in c.
The shape of mammalian phylogeny A. Purvis et al. 2467
Phil. Trans. R. Soc. B (2011)
trees (summarized as the maximum clade credibility
tree from the posterior distribution) were also used.
Outgroups were removed prior to analysis.
The phylogenies provided 2496 informative nodes,
representing many groups from three major clades—
arthropods (73 trees, 582 informative nodes); plants
(66 trees, 812 nodes); and non-mammalian vertebrates
(102 trees, 1102 nodes). Figure 3 shows how I
w
and
b
vary with log(node size) in these three groups. The I
w
regression lines for plants and arthropods are not
statistically distinct (ANCOVA: F
2,16
¼0.511, p.0.6;
pooled intercept ¼0.622 +0.0167 s.e., pooled slope ¼
0.0159 +0.0034 s.e.); neither are the mammalian and
non-mammalian vertebrate lines (ANCOVA: F
2,16
¼
1.355, p.0.2; pooled intercept ¼0.532 +0.0217
s.e.; pooled slope ¼0.0306 +0.0077 s.e.). Vertebrates
(mammals þnon-mammals) and non-vertebrates
(plants þarthropods) do not reject a pooled slope
(ANCOVA: F
1,36
¼3.402, p¼0.07; slope ¼0.0194 +
0.0035) but their intercepts differ significantly (non-ver-
tebrates 0.607 +0.0179 s.e.; vertebrates 0.561 +
0.0145 s.e.; F
1,37
¼10.09, p¼0.003). This overall
model explains 61.5 per cent of the variance in I
w
among bins with three parameters, providing a succinct
phenomenological description of imbalance.
The relationship between
b
and log(node size) is
harder to model because of
b
’s asymmetric error dis-
tribution and apparent nonlinearity in plants (figure
3d) and arthropods (figure 3e). A least-squares regres-
sion, weighted by the inverse of the confidence interval
width, has a significantly positive slope within plants
(slope ¼0.036 +0.014 s.e.; t
8
¼2.62, p¼0.03) and
arthropods (slope ¼0.057 +0.021 s.e., t
8
¼2.73,
p¼0.03; these regressions are not distinct: F
2,16
¼
0.713, p.0.5) but not in non-mammalian vertebrates
(t
8
¼0.106, p.0.9). Like I
w
,
b
is consistent with a
model having a pooled slope (0.094 +0.0424 s.e.)
but an intercept that is further from ERM expectation
for the non-vertebrates (21.466 +0.304 s.e.) than for
the pooled vertebrates (21.040 +0.241 s.e.).
The supertree’s shape is not particularly unusual
compared with the other vertebrate phylogenies in our
dataset. However, vertebrate imbalance patterns differ
from those seen in plants and arthropods. This differ-
ence was not detected in the previous comparisons of
phylogeny shape [25,29,55,56], but emerges from
using more informative summaries of shape and from
analysing near-complete phylogenies. The difference
parallels one in the temporal spacing of nodes reported
by McPeek [59], who found that molecular phylogenies
of chordates showed more evidence of slowdown than
did those of arthropods or an angiosperm group (we
are unable to repeat McPeek’s test as our non-mamma-
lian phylogenies mostly lack timescales). The parallel
may simply be coincidence, but both a slowdown and
relatively balanced nodes near the tips of the tree are
to be expected if competition with closely related species
inhibits speciation; perhaps this situation is more
common in vertebrates than in arthropods or plants.
Whether competition is or is not part of why these
groups show different imbalance signatures, it cannot
explain why phylogenies are unbalanced relative to
ERM expectations. Why is imbalance so widespread?
Many models of clade growth have been proposed that,
when simulated, generate trees that are more unbalanced
than ERM (reviewed by Mooers et al.[3]). In the next
partofthepaper,wesimulateseveralofthesemodels
and analyse the phylogenies they produce. If only some
models produce trees whose imbalance patterns resemble
those in figure 3, we may be closer to understanding the
sorts of processes that have structured diversity.
5. WHAT SIMPLE PROCESSES CAN PRODUCE
REALISTIC TREE SHAPES?
We simulated the growth of 100 phylogenies to 1000
‘species’ each under six stochastic macroevolutionary
models, detailed in table 3 (this is obviously not an
exhaustive set: e.g. [3]), and constructed imbalance
signatures using the pooled data from each model.
Table 3. Details of the six macroevolutionary models that were simulated. X
0
¼ancestral value of trait X;X
i
¼value of trait
Xin species i;
l
i
¼instantaneous speciation rate in species i;
l
0
,
l
1
¼instantaneous speciation rate for species in the 0 state
or 1 state, respectively; r
ab
¼per-lineage rate of transition of Xfrom state ato state b;t
i
¼time since species iwas last
involved in a speciation event. Models 15 were simulated using MeSA (www.agapow.net/software/mesa), model 6 with
PhyloGen (tree.bio.ed.ac.uk/software/phylogen/).
model details
1. punctuationally evolving key
trait [60,61]
X
0
¼100; X
i
changes (in both daughters) only at speciation events; changes are drawn
from a normal distribution,
m
¼0,
s
¼50. If X
i
becomes negative it is set to
0.
l
i
¼0.001 þX
i
/100 000
2. gradually evolving key trait
[60,61]
X
0
¼100; Xchanges continuously by Brownian motion with
m
¼0,
s
¼5 per time
unit; X
i
and hence
l
i
were assessed every 0.1 time units and at every speciation event.
Negative X
i
were truncated to 0.
l
i
¼0.0001 þX
i
/10 000
3. binary key trait X
0
¼0,
l
0
¼1,
l
1
¼10. r
01
¼r
10
¼0.05. States and rates were assessed every time unit
and at every speciation event
4. fast-evolving binary key trait X
0
¼0,
l
0
¼1,
l
1
¼10. r
01
¼r
10
¼10. States and rates were assessed every time unit
and at every speciation event
5. patency [61]
l
i
¼max(515t
i
, 0.6); ages and rates were assessed every 0.001 time units and at every
speciation event
6. spatial model Initial
l
¼1. Ancestral species placed on an infinite square grid (i.e. each cell is
adjacent to four others). Species occupy only one cell; they are selected at random to
speciate but can do so only if they are adjacent to at least one empty cell
2468 A. Purvis et al. The shape of mammalian phylogeny
Phil. Trans. R. Soc. B (2011)
The first three models have speciation probabilities
depending on a slowly evolving key trait, a class of
models first simulated by Heard [60]; in the fourth,
the binary trait that determines speciation rate evolves
so rapidly that it changes along most of the phylogeny’s
branches. The fifth model (first simulated by [61], see
also [62]) has speciation rates depending not on
species’ traits but on their ages, declining smoothly
from a high value straight after speciation to a back-
ground rate; we term this model patency to contrast
it with Losos & Adler’s [63] latency model in which
newly speciated lineages are unable to speciate again
for a while. Although it is implausible that species
age per se determines speciation rate, age could be
proxy for a macroecological trait that does itself
change as species age, like global population or
geographical range size [64,65]. The last process we
simulate is a spatial model in which species occupy
squares in a grid and can speciate only if they are
next to empty squares. In this model, species are
selected at random to speciate and, if they are
able to do so, the new species is placed in a rando-
mly chosen adjoining empty cell. The axes can be
viewed as latitude and longitude or dimensions of
niche space.
Figure 4 shows the I
w
and
b
signatures from each of
the six models. The models in which speciation rates
depend on slowly evolving traits produce qualitatively
different signatures (figure 4a,c) from those seen in
our empirical phylogenies (figure 3). These models
generated pronounced imbalance only deep within
the phylogeny: near the tips, nodes are consistent
with the ERM (because there is little or no interspecies
variation in the key trait on which clade selection can
act: [60,66]). The remaining three processes produce
more realistic signatures, with significant imbalance
even among small nodes (figure 4b,d).
The poor fit of models in which speciation rates
evolve only slowly suggests that (i) the relationship
between diversification rate and phenotype is more
complex than that simulated, (ii) different traits
affect diversification in different clades, or (iii) most
variation in diversification rate near the tips of the phy-
logeny does not depend on slowly evolving traits. A
previous simulation study [60] concluded that
models with slowly evolving traits could produce
imbalance comparable to that seen in empirical phylo-
genies, but that study measured the shape of each
simulated or empirical tree using an index that is
most sensitive to imbalance at the root.
The patency model can produce trees similar to
the empirical data for non-vertebrates (dotted line in
figure 4b,d). A slightly more complex spatial model
might produce quantitatively as well as qualitati-
vely realistic imbalance signatures: unrealistic features
of our simulation include that the initial ancestor
could diversify equally in every direction (leading to
balanced nodes at the root of the tree), and that
space was divided into contiguous equal-sized units.
Patency’s relatively good fit adds to growing evidence
that many clades may be at or near equilibrium diver-
sity. The North American Cenozoic mammalian fossil
record shows negative diversity-dependence [67] at the
species level, as does Phanerozoic marine macroinver-
tebrate genus-level diversity [68]. Complete molecular
phylogenies with timescales typically indicate that
0.4
0.5
0.6
0.7
0.8
0.9 (a)(b)
(c)(d)
within−bin Iw
234567
−2.0
−1.5
−1.0
−0.5
0.0
0.5
lo
g
(node size)
within−bin b
234567
lo
g
(node size)
Figure 4. I
w
and
b
signatures for six simulation models, which are described fully in the text. (a,c) Instantaneous speciation rate
depends linearly on the values of an evolving trait (squares, binary trait; circles, gradually evolving trait; diamonds, punctuational
trait). (b,d) Modelsin which close relatives often have very different speciation rates. Squares, rates depend on a fast-evolving binary
trait; circles, rates decline exponentially with time since speciation; diamonds, spatial model; dashed line, ERM expectation. Best-
fit lines for vertebrate data (solid lines) and non-vertebrate data (dotted lines) are provided for comparison.
The shape of mammalian phylogeny A. Purvis et al. 2469
Phil. Trans. R. Soc. B (2011)
speciation rates have declined as clades have grown
[58,59]. More generally, numbers of species in non-
nested clades usually do not depend strongly on their
ages (e.g. [15,27,46,69]). Mammalian lineages that
disperse successfully to a different biogeographic
realm tend to be more diverse than their sister clades
in their ancestral regions [45]. Dispersal ability is the
strongest known predictor of clade richness among
bird families [54], but good broad-scale data on dis-
persal ability are not, so far as we are aware, available
for mammals (or indeed for many non-avian groups).
Taken together, our results may explain why studies
finding significant correlates of diversity (reviewed by
Coyne & Orr [70]) have nearly all compared old,
rather than young, lineages. They may also help to
explain the dearth of significant correlates, and why
the explanatory power has typically been low. We
suggest that study of lineages’ opportunity for diversi-
fication, as well as their intrinsic attributes, is likely to
prove fruitful. The next section therefore considers the
spatial context of mammalian faunas alongside the
imbalance patterns they show.
6. THE GEOGRAPHICAL SCALE
OF PHYLOGENETIC IMBALANCE
Phylogenies arise through ecology and evolution over
long time periods and, crucially, in a geographical con-
text. Despite the importance of geography, only one
study has so far analysed phylogeny shape geographically.
Heard & Cox [44], in a study of primates in African and
South American nature reserves, showed how imbalance
can be partitioned into a hierarchy of spatial scales. The
shape of the phylogenyof a local assemblage can be com-
pared with that of the regional source pool, and the latter
compared with global imbalance. Such a decomposition
can shed light on the spatial scale at which imbalance is
generated, and perhaps on the processes that produce
it. The approach has strong conceptual links with com-
munity phylogenetics [71,72]. However, we are trying
not only to see how the local community is assembled
from a source pool, but also whether local processes
scale up to shape species diversity at larger scales.
We considered four spatial scales: global, ‘biorealm’ (a
convenient shorthand for unique combinations of World
Wide Fund for Nature (WWF) biome and biogeographic
realm: definitions from Olson et al.[73]), WWF ecore-
gion [73] and local assemblage (from local species
checklists). We computed
b
for the phylogeny of mam-
mals within each unit (e.g. each biorealm) at each level;
we chose
b
over I
w
because it varies less with node size
(figure 1). We also compared the imbalance of each
unit with a biogeographic null expectation [44]—the
value expected if the unit’s species were drawn at
random from the species within the next spatial scale
up. For example, an ecoregion’s
b
would be compared
with the distribution of
b
values obtained from 1000 ran-
domizations, each of which samples the ecoregion’s
number of species from the biorealm that contains it.
As a summary statistic, we subtract from each observed
b
the mean of the corresponding null distribution; we
term this difference
b
dev
. If ecoregions’ species are
random subsets of those in the corresponding biorealms,
the
b
dev
will be centred around 0.
Cooper et al.[72], in a community phylogenetic
analysis, found that mammalian assemblages tended
to contain fewer closely related species than predicted
under the biogeographic null model, consistent with
competitive exclusion among close relatives. Given
that mammalian phylogeny overall is unbalanced, an
overdispersed sample—in which species-rich groups
will tend to be under-represented—will tend to be
less unbalanced. We therefore predict that the
b
dev
for assemblages within ecoregions will tend to be posi-
tive (i.e. assemblages more balanced than ecoregions),
though we note that Cooper et al.[72] restricted their
analyses to species within the same family. At large
scales, in situ diversification presumably becomes an
important process [74], leading to diversifying clades
being over-represented relative to the biogeographic
null. However, the same pattern could be caused by
habitat filtering—a top-down rather than bottom-up
process, whereby some clades are under-represented
because ecological conditions do not suit them. For
both processes, we expect
b
dev
at the large scale to
be negative, as spatial units should be more unba-
lanced than the biogeographic null. Competitive
exclusion might also be more likely in species-rich,
energy-poor or ecologically uniform systems.
Species lists for biorealms and ecoregions were gen-
erated by overlaying maps of the units (downloaded
from www.worldwildlife.org/science/ecoregions/item
1267.html, accessed August 2006) with distribution
maps from the Global Mammal Assessment [75]; his-
torical, uncertain or introduced species ranges were
excluded, and only extant non-marine species present
in the phylogeny were considered. The final dataset
contained 4846 mammalian species within 62 bior-
ealms and 796 ecoregions. Assemblage data came
from a literature compilation of 242 species checklists
[76,77] containing a total of 1911 species and repre-
senting 140 ecoregions; the lists were for non-volant
mammals only, so Chiroptera (bats) were pruned
from the trees prior to randomizations of assemblages
within ecoregions. Because estimates of
b
from small
datasets seemed highly variable, we excluded all
samples with fewer than 20 species, leaving us with
59 biorealms, 728 ecoregions and 233 assemblages.
Figure 5 presents maps of
b
and
b
dev
at each spatial
scale. As expected, we find that strong imbalance at
the largest spatial scale (biorealms) largely reflects
the global imbalance of mammalian phylogeny. All
59 biorealms yielded negative estimates of
b
,37of
them significantly negative (figure 5a), whereas the
mean
b
dev
for biorealms (figure 5b) was not signifi-
cantly different from 0 (mean ¼0.02; Wilcoxon
signed-rank test, V¼943, p.0.05). Eight biorealms
had phylogenies significantly more unbalanced than
the biogeographic null while four biorealm phylogenies
were significantly more balanced.
Imbalance was again ubiquitous at the intermediate
scale of ecoregions within biorealms: the estimate of
b
was positive in only 19 of the 728 ecoregions (none
significantly), and was significantly negative in 112
(figure 5c). Furthermore, the mean
b
dev
showed sig-
nificant imbalance when compared with the biorealm
source pools (mean ¼20.07; Wilcoxon signed-rank
test, V¼101 487, p,0.001). The biogeographic
2470 A. Purvis et al. The shape of mammalian phylogeny
Phil. Trans. R. Soc. B (2011)
null model was rejected in 82 of the 720 ecoregions
where it could be tested (the ecoregion and biorealm
had the same set of species in the other eight eco-
regions); of these, 56 ecoregion phylogenies were
more balanced than expected and 26 more unbalanced
(figure 5d).
Moving to the level of assemblages within ecoregions,
b
was negative in 209 of 233 assemblages but signifi-
cantly so in only two (figure 5e); this lack of
significance is unsurprising given that the assemblage
phylogenies are seldom large. However, contrary to
our expectations,
b
dev
is significantly negative on average
(mean ¼–0.12; Wilcoxon signed-rank test, V¼8503,
p,0.001), indicating that assemblage phylogenies
tend to be unbalanced compared with their ecoregion
source pools. This finding suggests that habitat filtering
(or, less likely at this small scale, in situ speciation) dom-
inates over competitive exclusion, though community
phylogenetic approaches (e.g. [71,74]) would provide
a more direct test. Twelve assemblages rejected the bio-
geographic null model, out of 215 that could be tested;
seven were more balanced than expected and five more
unbalanced (figure 5f).
To investigate environmental influences on imbalance
signatures, we used multiple regression to predict
b
dev
from different variables that reflected current envi-
ronmental conditions or system characteristics. The
environmental variables for our spatial units were calcu-
lated using an ArcInfo macro; only assemblages with
a digitized polygon were included in this analysis
(215 assemblages; [76]). Variables and their sources
were: mean annual actual evapotranspiration (AET, in
millimetres; [78]; downloaded from http://www.grid.
unep.ch); mean annual temperature (in 8C; [79],
downloaded from http://www.worldclim.org/); mean
elevation and range in elevation (in metres; [80], down-
loaded from http://edc.usgs.gov); and ecosystem count
as a coarse measure of habitat heterogeneity ([81],
downloaded from http://edcsns17.cr.usgs.gov/glcc). Pre-
cipitation was included at first but had to be excluded
because of variance inflation (it was highly collinear
with AET). All models also included the area (log-trans-
formed) and number of species within the focal unit.
Because in situ diversification is expected to produce
concentrations of newly formed species, we additionally
included as a predictor the median terminal branch
length for the species in each spatialunit on the complete
mammalian phylogeny as a proxy for the recency of the
species; however, because terminal polytomies in the
supertree typically reflect lack of knowledge, we adjusted
the branch lengths following Mooers et al.[3].
Following Lichstein et al.[82] and Dormann et al.
[83], we tested for the presence of spatial autocorrela-
tion in the regression residuals from our initial,
ordinary least-squares (OLS) model with Moran’s
Icorrelograms, and accounted for the spatial autocor-
relation with spatial autoregression (SAR, simultaneous
autoregressive model). We used an SAR
error
model,
which models the autoregressive process in the error
term and has been recommended as a reliable spatial
method [84]. Moran’s Icorrelograms and spatial
models were generated with the R packages spdep [85]
and ncf [86]. We tested standardized Ivalues for signifi-
cance with a one-tailed randomization test for positive
autocorrelation (999 permutations; [82]). In the SAR
models, we defined neighbours as data points closer to
each other than a model-specific maximum distance,
which was chosen by optimizing the model AIC value
(Akaike’s information criterion) following Cooper &
Purvis [87]. We used a row-standardized coding
–1.61
(a)(b)
–2.18
–1.25 –0.75 –0.25 0.25 0.75 4.52 –0.60 –0.20 0.20 0.60 1.00 4.26
bb dev
(c)(d)
(e)(f)
Figure 5. Maps of (a,c,e)
b
and (b,d,f)
b
dev
for (a,b) biorealms, (c,d) ecoregions and (e,f) assemblages. See text for explanation.
The shape of mammalian phylogeny A. Purvis et al. 2471
Phil. Trans. R. Soc. B (2011)
scheme for the spatial weights matrix [84], calculated r
2
values using Nagelkerke’s formula [88] and assessed the
contribution of each variable to OLS and SAR model
fits with likelihood-ratio tests for nested models [82].
Although
b
dev
was significantly negative on average
at each spatial scale (meaning that the phylogeny of the
species present in the spatial unit tended to be more
unbalanced than the biogeographic null), the signifi-
cant environmental predictors of
b
dev
depended on
scale (table 4), perhaps reflecting the differing relative
importance of top-down and bottom-up controls on
imbalance (though we caution that no model has
high explanatory power, and that the precision of
environmental variables is often scale-dependent).
At the largest scale, biorealm phylogenies were more
unbalanced relative to the global phylogeny when the
biorealm contained more different ecosystems and, in
OLS models, was at high altitude. The altitude term
distinguishes the montane biome from the remainder.
Habitat filtering (a top-down process) and in situ diver-
sification (bottom-up) are both likely to be important
at this broad scale. Indeed, the latter may follow
from the former—if few clades are able to persist at
high altitudes, those that can are likely to radiate into
a wider range of niches than they might occupy in
less harsh environments.
Moving to ecoregions within biorealms, the map of
b
dev
(figure 5d) suggests that the most negative values
are in deserts, the Northern high latitudes, and the
Himalayas; the Amazonian basin fauna, by contrast,
has a relatively balanced phylogeny (positive
b
dev
).
The regression models find stronger imbalance to be
associated with small ecoregions containing few
species and, in OLS models, having low AET and
more different ecosystems. Habitat filtering is likely
to be more important at this scale than at the broader
scale, and in situ diversification less so.
In situ diversification is unlikely at the scale of
assemblages within ecoregions, whose patterns pre-
sumably reflect a mixture of habitat filtering and
competitive exclusion. The generally negative
b
dev
values suggest that the former dominates, especially
in warm, topographically heterogeneous areas, but
the signs of the environmental predictors of
b
dev
are
consistent with a role for competitive exclusion in
assemblages that are species-rich, in cold, uniform
places.
Although interpretation of our results remains ten-
tative, the spatial decomposition of imbalance in
mammalian phylogeny reveals many patterns that
cannot be attributed to the biogeographic null. This
shows the possible usefulness of Heard & Cox’s [44]
groundbreaking approach for understanding how
bottom-up and top-down processes might interact to
control clade diversity and hence imbalance. Their rela-
tive lack of non-null results may reflect their choice of
the only large mammalian clade—Primates—that does
not have an unbalanced phylogeny at a global level.
7. A DESCRIPTIVE MODEL OF MAMMALIAN
MACROEVOLUTION
What do our analyses suggest about the processes that
have shaped mammalian phylogeny? The I
w
and
b
signatures of the supertree highlight the persistence
of small basal clades, and also show that closely related
species often have very different chances of diversifica-
tion. Our simulations indicate that the latter is more
easily achieved if speciation rates decrease with time
since speciation or, more plausibly, with the occu-
pancy of adjacent areas or niches, than if they are
determined by slowly evolving traits. The occupancy
model is also compatible with our analysis of the
spatial scale of imbalance. It suggests that diversifica-
tion will show negative diversity-dependence, a
pattern seen in many molecular phylogenies [58,59]
and the mammalian fossil record [67]. Equilibrium
diversity seems likely to depend on both the geographi-
cal area available to the clade and the amount of
energy available to support populations of its species
in the face of competition from other lineages.
Four main ways have been proposed by which a
clade or subclade’s diversity might increase determi-
nistically beyond this initial equilibrium (e.g. [6,9,53,
59,8992]):
Spreading in space. A species that colonizes new
geographical space, by dispersal to a disjunct
region or through an expansion of suitable habitat,
may diversify taxonomically without markedly
expanding its ecological niche [53,91,93]. The
equilibrium diversity of such a clade should
depend on the size of the new area and the
number of competitors present.
Spreading in niche space. An adaptive breakthrough
to new niche space (a new adaptive zone; [9,93])
may permit a lineage to diversify more rapidly
than its sister clade [53,89]; again, equilibrium
diversity presumably depends on the size and occu-
pancy of the adaptive zone [89,94].
Improvement. An adaptation that makes a species
markedly better at occupying its niche (through,
for example, increased metabolic efficiency or
disease resistance) may enable the species to out-
compete other species both within and at the
margin of its niche [9,89,93]. How far the resulting
clade is expected to spread through niche space is
governed by how much of a competitive edge the
improvement gives. Clades with such key inno-
vations are always expected to be more diverse
than their sister clades without them (though they
may competitively wipe out their sister clades; [89]).
Narrower species. The first three mechanisms raise
diversity by increasing the resource base; an
alternative is to slice the resource base more
thinly [89,95]. Sexual selection by female choice,
for example, makes rapid speciation more likely,
and is repeatedly associated with high diversity in
sister-clade comparisons within birds [69,96]. Eco-
logical specialization [97], behavioural and societal
complexity, philopatry, changes to karyotypic
architecture, or a ‘cellular’ population structure
might also be relevant to mammals.
If each of these four classes of event occurs at random
with respect to phylogeny (a reasonable null starting
point) then, by the law of averages, they are more
likely to happen within species-rich clades than in
2472 A. Purvis et al. The shape of mammalian phylogeny
Phil. Trans. R. Soc. B (2011)
Table 4. Non-spatial multiple regression (OLS) and spatial autoregression (SAR) model results for
b
dev
at each of three spatial levels. Apart from characterizing each model, we also show
each variable’s slope, significance and likelihood-ratio test result (LR).
biorealms ecoregions assemblages
OLS SAR OLS SAR OLS SAR
AIC 235.3 234.1 320.6 82.5 335.6 314.5
d.f. 50 48 694 692 171 169
r
2
28.7 29.6 17.6 41.4 15.1 25.4
neighbourhood distance n.a. 1970.79 n.a. 626.9 n.a. 700.6
autocorrelation parameter n.a. 0.173 n.a. 0.614 n.a. 0.4
slope LR slope LR slope LR slope LR slope LR slope LR
AET ,0.001 ,0.1 ,0.001 ,0.1 ,0.001** 222.9 ,0.001 2.7 ,0.001 5.1 ,0.001 0.3
temperature 20.001 1.9 .20.001 1.5 20.004*4.1 ,0.001 ,0.1 20.019*6.8 20.016 3.2
mean elevation .20.001*5.2 .20.001 2.8 ,0.001 ,0.1 ,0.001 0.2 ,0.001 2.1 ,0.001 1.6
elevation range ,0.001 0.6 ,0.001 0.1 ,0.001 ,0.1 .20.001 0.2 .20.001*4.7 .20.001 3.0
ecosystem count 20.008** 10.2 20.007** 5.3 20.006*** 13.3 20.002 1.5 20.005 0.2 ,0.001 ,0.1
area 20.016 0.8 20.013 0.5 0.020*4.6 0.028** 10.2 20.010 0.1 20.003 ,0.1
species number ,0.001 0.7 ,0.001 0.3 0.002*** 47.0 0.001** 7.4 0.015*** 12.7 0.014*** 12.6
median terminal branch length 20.197 1.0 20.173 0.7 0.091 1.1 0.117 1.4 20.411 3.3 20.442 3.3
Significance levels: *** p,0.001, ** p,0.01, *p,0.05.
The shape of mammalian phylogeny A. Purvis et al. 2473
Phil. Trans. R. Soc. B (2011)
species-poor ones; any initial pattern of imbalance in
the phylogeny is therefore accentuated. Adaptive
improvements will, likewise by the law of averages,
tend to arise in species adapted to wide rather than
narrow adaptive zones and living in large rather than
small regions. Individuals within these species will
tend to outcompete those in related species in the
same region, causing competitive extinctions as
Darwin [98] envisaged; and there may be successive
waves of improvement. For an innovation to spread
further, however, the clade bearing it will have to
either colonize new areas or break through to new
niches, both of which may have long waiting times
(i.e. low probabilities per unit time). ‘Unimproved’
lineages can therefore persist as relicts in isolated
regions from which competitors are absent (e.g. Sole-
nodon), or in isolated niches (e.g. myrmecophagy in
Pholidota and Tubulidentata, or haematophagy in
Desmodontinae) where their incumbency advantage
permits them to outcompete more ‘advanced’ but
less well-adapted species. If these relictual lineages
have diversity-dependent dynamics, they may persist
indefinitely even at low diversity; this provides a plaus-
ible mechanism for the long persistence of small basal
clades suggested by the I
w
signatures.
Major environmental perturbations would be
expected to move the system away from this quasi-
equilibrial state, changing the breadths of the adaptive
zones. Some clades would then decline or go extinct
whereas others, whose adaptive zones had become
broader, would be able to diversify.
Under this model, phylogenies are unbalanced
because regions and niches vary in the diversity they
can support, because new radiations will probably orig-
inate in already-diverse clades, and because relictual
lineages are hard to extirpate. Different clades radiate
in different places, however, so the global phylogeny
will tend to be less unbalanced than expected from the
phylogeny of species found within a biorealm or ecore-
gion. Apparently downshifted clades arise as survivors
of otherwise supplanted groups or as occupants of
narrow adaptive zones: island endemics may be the
ultimate evolutionary dead-ends [99]. Much of the vari-
ation in richness among clades emerges as being
independent of trait variation, reflecting historical and
geographical contingencies.
The model also makes predictions about the tempo-
ral pattern of nodes in phylogeny. A complex overall
pattern is expected: although the overall dynamics
are equilibrial, some subclades in the early phase of
their radiation will be diversifying exponentially,
others approaching an equilibrium diversity, and
others declining deterministically. Furthermore, radi-
ations based on niche broadening might have a
different temporal pattern from those based on finer
subdivision of the niche [59]. One prediction, how-
ever, is that lineages occupying biomes that have
recently expanded in area and energy are expected
to show rapid recent radiation; recent findings of
high speciation rates in North temperate regions for
birds [100], mammals [28] and an angiosperm clade
(carnations; [101]) are consistent with this prediction.
Our verbal model is an attempt to develop a simple
set of ecologically and evolutionarily plausible
processes that could together have generated the
shape of the mammalian phylogeny. Process-based
simulation models are increasingly used in macroevolu-
tion (e.g. [59,102]), permitting tentative identification
of parameters to which diversity patterns might be
most sensitive. Although phylogenies without fossils
are inevitably missing much important information
about clade dynamics, large complete phylogenies are
nonetheless a rich source of patterns against which
model outputs can be compared, especially when com-
bined with environmental or trait data. We anticipate
that multi-scale models, incorporating local processes
such as adaptation and competition as well as global
processes such as global change and tectonic move-
ment, will have much to offer in the quest for global
biodiversity models, and for our understanding of why
so much of the Tree of Life is so asymmetric.
We are grateful to Ingi Agnarsson for sending the cytochrome
b phylogenies of carnivores and cetartiodactyls; to Ally
Phillimore for providing his bird phylogenies; to Andrew
Rambaut for programming the spatial simulation model; to
Paul Agapow for programming the other simulation models;
to Rodolphe Bernard, Natalie Cooper, David Orme and
Gavin Thomas for advice and assistance; to Rob Beck,
Lynsey McInnes and two referees for their extremely helpful
comments; and to Kate Jones for her patience. This work
was supported by the NERC, the Leverhulme Trust, the
Danish National Research Foundation, MICINN project
CGL2009-12703-C03-01, the JCyL GR 249-200 and an
Imperial College Research Excellence Award.
REFERENCES
1 Mooers, A. Ø. & Heard, S. B. 1997 Evolutionary pro-
cess from phylogenetic tree shape. Q. Rev. Biol. 72,
3154. (doi:10.1086/419657)
2 Purvis, A. 1996 Using interspecific phylogenies to test
macroevolutionary hypotheses. In New uses for new phylo-
genies (eds P. H. Harvey, A. J. Leigh Brown, J. Maynard
Smith & S. Nee), pp. 153– 168. Oxford, UK: Oxford
University Press.
3 Mooers, A. Ø, Harmon, L. J., Blum, M. G. B., Wong,
D. H.J. & Heard, S. B. 2007 Some models of phylogenetic
tree shape. In Reconstructing evolution: new mathematical
and computational advances (eds O. Gasceul & M. Steel),
pp. 149–170. Oxford, UK: Oxford University Press.
4 Wilson, D. E. & Reeder, D. M. 2005 Mammal species of
the world: a taxonomic and geographic reference, 3rd edn.
Baltimore, MD: The Johns Hopkins University Press.
5 Yule, G. U. 1924 A mathematical theory of evolution,
based on the conclusions of Dr. J. C. Willis F.R.S.
Phil. Trans. R. Soc. Lond. B 213, 2187. (doi:10.1098/
rstb.1925.0002)
6 Simpson, G. G. 1953 The major features of evolution.
New York, NY: Columbia University Press.
7 Stanley, S. M. 1979 Macroevolution. San Francisco, CA:
W H Freeman.
8 Anderson, S. 1974 Patterns of faunal evolution. Q. Rev.
Biol. 49, 115. (doi:10.1086/408171)
9 Van Valen, L. 1971 Adaptive zones and the orders of
mammals. Evolution 25, 420428. (doi:10.2307/
2406935)
10 Bininda-Emonds, O. R. P. et al. 2007 The delayed rise
of present-day mammals. Nature 446, 507– 512.
(doi:10.1038/nature05634)
11 Schwartz, R. S. & Mueller, R. L. 2010 Branch length
estimation and divergence dating: estimates of error in
2474 A. Purvis et al. The shape of mammalian phylogeny
Phil. Trans. R. Soc. B (2011)
Bayesian and maximum likelihood frameworks. BMC
Evol. Biol. 10,5.(doi:10.1186/1471-2148-10-5)
12 Pulque
´rio, M. J. F. & Nichols, R. A. 2007 Dates from the
molecular clock: how wrong can we be? Tre n d s E c o l . Ev o l .
22, 180 184. (doi:10.1016/j.tree.2006.11.013)
13 Purvis, A. 2008 Phylogenetic approaches to the study of
extinction. Annu. Rev. Ecol. Evol. Syst. 39, 301319.
(doi:10.1146/annurev.ecolsys.063008.102010)
14 Rabosky, D. L. 2010 Extinction rates should not be esti-
mated from molecular phylogenies. Evolution 64, 1816–
1824. (doi:10.1111/j.1558-5646.2009.00926.x)
15 Ricklefs, R. E. 2007 Estimating diversification rates
from phylogenetic information. Trends Ecol. Evol. 22,
601610. (doi:10.1016/j.tree.2007.06.013)
16 Slowinski, J. B. & Guyer, C. 1989 Testing null models
in questions of evolutionary success. Syst. Zool. 38,
189191. (doi:10.2307/2992389)
17 Fritz, S. A., Bininda-Emonds, O. R. P. & Purvis, A.
2009 Geographical variation in predictors of mamma-
lian extinction risk: big is bad, but only in the tropics.
Ecol. Lett. 12, 538 549. (doi:10.1111/J.1461-0248.
2009.01307.X)
18 Mooers, A. Ø. 1995 Tree balance and tree complete-
ness. Evolution 49, 379– 384. (doi:10.2307/2410349)
19 Wilkinson, M. et al. 2005 The shape of supertrees to
come: tree shape related properties of fourteen super-
tree methods. Syst. Biol. 54, 419– 431. (doi:10.1080/
10635150590949832)
20 Agnarsson, I., Kuntner, M. & May-Collado, L. J.
2010 Dogs, cats, and kin: a molecular species-level phy-
logeny of Carnivora. Mol. Phylogenet. Evol. 54, 726
745. (doi:10.1016/j.ympev.2009.10.033)
21 Agnarsson, I. & May-Collado, L. J. 2008 The phylo-
geny of Cetartiodactyla: the importance of dense
taxon sampling, missing data, and the remarkable
promise of cytochrome b to provide reliable species-
level phylogenies. Mol. Phylogenet. Evol. 48, 964985.
(doi:10.1016/j.ympev.2008.05.046)
22 Fusco, G. & Cronk, Q. C. B. 1995 A new method for
evaluating the shape of large phylogenies. J. Theoret.
Biol. 175, 235243. (doi:10.1006/jtbi.1995.0136)
23 Purvis, A., Katzourakis, A. & Agapow, P. M. 2002 Eval-
uating phylogenetic tree shape: two modifications to
Fusco and Cronk’s method. J. Theoret. Biol. 214,99
103. (doi:10.1006/jtbi.2001.2443)
24 Aldous, D. J. 2001 Stochastic models and descriptive
statistics for phylogenetic trees, from Yule to today.
Stat. Sci. 16, 23 34. (doi:10.1214/ss/998929474)
25 Blum, M. G. B. & Francois, O. 2006 Which random pro-
cesses describe the tree of life? A large-scale study of
phylogenetic tree imbalance. Syst. Biol. 55, 685 691.
(doi:10.1080/10635150600889625)
26 Bortolussi, N., Durand, E., Blum, M. & Francois, O.
2006 apTreeshape: statistical analysis of phylogenetic
tree shape. Bioinformatics 22, 363 364. (doi:10.1093/
bioformatics/bti798)
27 Davies,T.J.,Barraclough,T.G.,Chase,M.W.,Soltis,P.S.,
Soltis, D. E. & Savolainen, V. 2004 Darwin’s abominable
mystery: insights from a supertree of the angiosperms.
Proc. Natl Acad. Sci. USA 101, 1904– 1909. (doi:10.
1073/Pnas.0308127100)
28 Davies, T. J. et al. 2008 Phylogenetic trees and the future
of mammalian biodiversity. Proc. Natl Acad. Sci. USA
105, 11 556 –11 563. (doi:10.1073/pnas.0801917105)
29 Guyer, C. & Slowinski, J. B. 1991 Comparison of
observed phylogenetic topologies with null expectations
among three monophyletic lineages. Evolution 45,
340350. (doi:10.2307/2409668)
30 Stam, E. 2002 Does imbalance in phylogenies reflect
only bias? Evolution 56, 1292– 1295.
31 Purvis, A. & Agapow, P. M. 2002 Phylogeny imbalance:
taxonomic level matters. Syst. Biol. 51, 844– 854.
(doi:10.1080/10635150290102546)
32 Holman, E. W. 2005 Nodes in phylogenetic trees: the
relation between imbalance and number of descendant
species. Syst. Biol. 54, 895– 899. (doi:10.1080/1063
5150500354696)
33 Heath, T. A., Zwickl, D. J., Kim, J. & Hillis, D. M. 2008
Taxon sampling affects inferences of macroevolutionary
processes from phylogenetic trees. Syst. Biol. 57,
160166. (doi:10.1080/10635150701884640)
34 Freckleton, R. P., Pagel, M. & Harvey, P. H. 2003
Comparative methods for adaptive radiations. In Macro-
ecology: concepts and consequences (eds T. M. Blackburn
& K. J. Gaston), pp. 391407. Oxford, UK: Blackwell.
35 Chan, K. M. A. & Moore, B. R. 2002 Whole-tree
methods for detecting differential diversification rates.
Syst. Biol. 51, 855– 865. (doi:10.1080/10635150
290102555)
36 Chan, K. M. A. & Moore, B. R. 2005 SymmeTREE:
whole-tree analysis of differential diversification rates.
Bioinformatics 21, 1709 1710. (doi:10.1093/bioinfor-
matics/bti175)
37 Moore, B. R., Chan, K. M. A. & Donoghue, M. J.
2004 Detecting diversification rate variation in super-
trees. In Phylogenetic supertrees: combining information to
reveal the Tree of Life (ed. O. R. P. Bininda-Emonds),
pp. 487– 533. Dordrecht, The Netherlands: Kluwer.
38 Gingerich, P. D. 2006 Environment and evolution
through the Paleocene– Eocene thermal maximum.
Trends Ecol. Evol. 21, 246– 253. (doi:10.1016/j.tree.
2006.03.006)
39 Beard, K. C. 2008 The oldest North American primate
and mammalian biogeography during the Paleocene
Eocene thermal maximum. Proc. Natl Acad. Sci. USA
105, 38153818. (doi:10.1073/pnas.0710180105)
40 Zachos, J., Pagani, M., Sloan, L., Thomas, E. &
Billups, K. 2001 Trends, rhythms, and aberrations
in global climate 65 Ma to present. Science 292,
686693. (doi:10.1126/science.1059412)
41 Woodburne, M. O., Gunnell, G. F. & Stucky, R. K.
2009 Climate directly influences Eocene mammal
faunal dynamics in North America. Proc. Natl Acad.
Sci. USA 106, 13 399 –13 403. (doi:10.1073/Pnas.
0906802106)
42 Maas, M. C., Krause, D. W. & Strait, S. G. 1988 The
decline and extinction of Plesiadapiformes (Mammalia,
?Primates) in North-America: displacement or replace-
ment? Paleobiology 14, 410– 431.
43 Haffer, J. 1997 Alternative models of vertebrate specia-
tion in Amazonia: an overview. Biodivers. Conserv. 6,
451476. (doi:10.1023/A:1018320925954)
44 Heard, S. B. & Cox, G. H. 2007 The shapes of phylo-
genetic trees of clades, faunas, and local assemblages:
exploring spatial pattern in differential diversification.
Am. Nat. 169, E107E118. (doi:10.1086/512690)
45 Kisel, Y., McInnes, L., Toomey, N. H. & Orme,
C. D. L. 2011 How diversification rates and diversity
limits combine to create large-scale speciesarea
relationships. Phil. Trans. R. Soc. B 366, 25142525.
(doi:10.1098/rstb.2011.0022)
46 Ricklefs, R. E. 2003 Global diversification rates of pas-
serine birds. Proc. R. Soc. Lond. B 270, 2285 2292.
(doi:10.1098/rspb.2003.2489)
47 Heard, S. B. & Mooers, A. Ø. 2002 Signatures of
random and selective mass extinctions in phylogenetic
tree balance. Syst. Biol. 51, 889– 897. (doi:10.1080/
10635150290102591)
48 Hurlbert, R. 2001 The fossil vertebrates of Florida.
Gainesville, FL: University Press of Florida.
The shape of mammalian phylogeny A. Purvis et al. 2475
Phil. Trans. R. Soc. B (2011)
49 Turvey, S. T., Oliver, J. R., Storde, Y. M. N. & Rye, P.
2007 Late Holocene extinction of Puerto Rican native
land mammals. Biol. Lett. 3, 193196. (doi:10.1098/
Rsbl.2006.0585)
50 McKinney, M. L. 1999 Biodiversity dynamics: niche
preemption and saturation in diversity equilibria. In Biod-
viersity dynamics (eds M. L. McKinney & J. A. Drake),
pp. 1– 16. New York, NY: Columbia University Press.
51 Colles, A., Liow, L. H. & Prinzing, A. 2009 Are special-
ists at risk under environmental change? Neoecological,
paleoecological and phylogenetic approaches. Ecol. Lett.
12, 849863. (doi:10.1111/j.1461-0248.2009.01336.x)
52 Springer, M. S., Meredith, R. W., Janecke, J. E. &
Murphy, W. J. 2011 The historical biogeography of
Mammalia. Phil. Trans. R. Soc. B 366, 2478 2502.
(doi:10.1098/rstb.2011.0023)
53 Moore, B. R. & Donoghue, M. J. 2007 Correlates of
diversification in the plant clade Dipsacales: geographic
movement and evolutionary innovations. Am. Nat. 170,
S28S55. (doi:10.1086/519460)
54 Phillimore, A. B., Freckleton, R. P., Orme, C. D. L. &
Owens, I. P. F. 2006 Ecology predicts large-scale pat-
terns of phylogenetic diversification in birds. Am. Nat.
168, 220229. (doi:10.1086/505763)
55 Guyer, C. & Slowinski, J. B. 1993 Adaptive radiation
and the topology of large phylogenies. Evolution 47,
253263. (doi:10.2307/2410133)
56 Herrada, E. A., Tessone, C. J., Klemm, K., Eguiluz, V. M.,
Hernandez-Garcia, E. & Duarte, C. M. 2008 Universal
scaling in the branching of the Tree of Life. PLoS ONE
3,e2757.(doi:10.1371/journal.pone.0002757)
57 Thomas, G. H., Wills, M. A. & Szekely, T. 2004
A supertree approach to shorebird phylogeny.
BMC Evol. Biol. 4, 28. (doi:10.1186/1471-2148-4-28)
58 Phillimore, A. B. & Price, T. D. 2008 Density-
dependent cladogenesis in birds. PLoS Biol. 6,
483489. (doi:10.1371/journal.pbio.0060071)
59 McPeek, M. 2008 The ecological dynamics of clade
diversification and community assembly. Am. Nat.
172, e270e284. (doi:10.1086/an.0)
60 Heard, S. B. 1996 Patterns in phylogenetic tree balance
with variable and evolving speciation rates. Evolution 50,
21412148. (doi:10.2307/2410685)
61 Agapow, P. M. & Purvis, A. 2002 Power of eight tree
shape statistics to detect nonrandom diversification: a
comparison by simulation of two models of cladogen-
esis. Syst. Biol. 51, 866– 872. (doi:10.1080/10635150
290102564)
62 Steel, M. & McKenzie, A. 2001 Properties of phyloge-
netic trees generated by Yule-type speciation models.
Math. Biosci. 170, 91112. (doi:10.1016/S0025-
5564(00)00061-4)
63 Losos, J. B. & Adler, F. R. 1995 Stumped by trees? A
generalized null model for patterns of organismal diver-
sity. Am. Nat. 145, 329342. (doi:10.1086/285743)
64 Hubbell, S. P. 2001 The unified neutral theory of biodiver-
sity and biogeography. Princeton, NJ: Princeton
University Press.
65 Vrba, E. S. & DeGusta, D. 2004 Do species populations
really start small? New perspectives from the Late
Neogene fossil record of African mammals. Phil.Trans.R.
Soc. Lond. B 359, 285 –293. (doi:10.1098/rstb.2003.1397)
66 Kirkpatrick, M. & Slatkin, M. 1993 Searching for evol-
utionary patterns in the shape of a phylogenetic tree.
Evolution 47, 1171– 1181. (doi:10.2307/2409983)
67 Alroy, J. 2009 Speciation and extinction in the fossil
record of North American mammals. In Speciation
and patterns of diversity (eds R. Butlin, J. Bridle &
D. Schluter), pp. 301323. Cambridge, MA:
Cambridge University Press.
68 Alroy, J. 2008 Dynamics of origination and extinction in
the marine fossil record. Proc. Natl Acad. Sci. USA 105,
11 53611 542. (doi:10.1073/pnas.0802597105)
69 Owens, I. P. F., Bennett, P. M. & Harvey, P. H. 1999
Species richness among birds: body size, life history,
sexual selection or ecology? Proc. R. Soc. Lond. B 266,
933939. (doi:10.1098/rspb.1999.0726)
70 Coyne, J. A. & Orr, H. A. 2004 Speciation. Sunderland,
MA: Sinauer.
71 Webb, C. O., Ackerley, D. D., McPeek, M. A. & Dono-
ghue, M. J. 2002 Phylogenies and community ecology.
Annu. Rev. Ecol. Syst. 33, 475– 505. (doi:10.1146/
annurev.ecolsys.33.010802.150448)
72 Cooper, N., Rodriguez, J. & Purvis, A. 2008 A common
tendency for phylogenetic overdispersion in mammalian
assemblages. Proc. R. Soc. B 275, 2031 2037. (doi:10.
1098/rspb.2008.0420)
73 Olson, D. M. et al. 2001 Terrestrial ecoregions of the world:
a new map of life of Earth. Bioscience 51, 933 –938. (doi:10.
1641/0006-3568(2001)051[0933:TEO TWA]2.0.CO;2)
74 Cardillo, M. 2011 Phylogenetic structure of mammal
assemblages at large geographic scales: linking phylogenetic
community ecology with macroecology. Phil.Trans.R.Soc.
B366, 25452553. (doi:10.1098/rstb.2011.0021)
75 IUCN. Red list of threatened species. International Union for
Conservation of Nature (IUCN) (database on Internet).
See http://www.iucnredlist.org/mammals (last accessed:
November 2008).
76 Hortal, J., Rodrı
´guez, J., Nieto-Dı
´az, M. & Lobo, J. M.
2008 Regional and environmental effects on the species
richness of mammal assemblages. J. Biogeogr. 35, 1202–
1214. (doi:10.1111/j.1365-2699.2007.01850.x)
77 Rodrı
´guez, J. 1999 Use of cenograms in mammalian
palaeoecology. A critical review. Lethaia 32, 331 347.
(doi:10.1111/j.1502-3931.1999.tb00551.x)
78 GRID-Europe dataset GNV183. 1994 United Nations
Environment Programme (database on Internet). See
http://www.grid.unep.ch (last accessed: February 2007).
79 Hijmans, R. J., Cameron, S. E., Parra, J. L., Jones, P. G. &
Jarvis, A. 2005 Very high resolution interpolated
climate surfaces for global land areas. Int. J. Climatol. 25,
1965– 1978. (doi:10.1002/joc.1276)
80 GTOPO30 global digital elevation model. 1996 U.S.
Geological Survey Centre for Earth Resources Obser-
vation and Science (USGS EROS) (database on
Internet). See http://edc.usgs.gov (last accessed: March
2007).
81 Global Ecosystems V. 1. 1994 U.S. Geological Survey
Centre for Earth Resources Observation and Science
(USGS EROS) (database on Internet). See http://
edcsns17.cr.usgs.gov/glcc (last accessed: May 2008).
82 Lichstein, J. W., Simons, T. R., Shriner, S. A. & Franzreb,
K. E. 2002 Spatial autocorrelation and autoregressive
models in ecology. Ecol. Monogr. 72, 445– 463. (doi:10.
1890/0012-9615(2002)072[0445:SAAAMI]2.0.CO;2)
83 Dormann, C. F. et al. 2007 Methods to account for
spatial autocorrelation in the analysis of species distri-
butional data: a review. Ecography 30, 609– 628.
(doi:10.1111/j.2007.0906-7590.05171.x)
84 Kissling, W. D. & Carl, G. 2008 Spatial autocorrelation
and the selection of simultaneous autoregressive
models. Glob. Ecol. Biogeogr. 17, 59– 71. (doi:10.1111/
j.1466-8238.2007.00334.x)
85 Bivand, R. 2007 spdep. Spatial dependence: weighting
schemes, statistics and models. R package version 0.4– 7.
With contributions by L. Anselin, O. Berke, A. Bernat,
M. Carvalho, Y. Chun, C. Dormann, S. Dray,
R. Halbersma, N. Lewin-Koh, H. Ono, P. Peres-Neto,
M. Tiefelsdorf & D. Yu. See http://cran.r-project.org/
package¼spdep.
2476 A. Purvis et al. The shape of mammalian phylogeny
Phil. Trans. R. Soc. B (2011)
86 Bjørnstad, O. N. 2006 ncf. Spatial nonparametric covari-
ance functions. R package version 1.0-9. See http://cran.r-
project.org/web/packages/ncf.
87 Cooper, N. & Purvis, A. 2010 Body size evolution in
mammals: complexity in tempo and mode. Am. Nat.
175, 727738. (doi:10.1086/652466)
88 Nagelkerke, N. J. D. 1991 A note on the general defi-
nition of the coefficient of determination. Biometrika
78, 691 692. (doi:10.1093/biomet/78.3.691)
89 Heard, S. B. & Hauser, D. L. 1995 Key evolutionary
innovations and their ecological mechanisms. Hist.
Biol. 10, 151173. (doi:10.1080/10292389509380518)
90 Erwin, D. H. 1992 A preliminary classification of evol-
utionary radiations. Hist. Biol. 6, 133 147. (doi:10.
1080/10292389209380423)
91 Rosenzweig, M. L. 1995 Species diversity in space and
time. Cambridge, MA: Cambridge University Press.
92 Schluter, D. 2000 The ecology of adaptive radiation.
Oxford, UK: Oxford University Press.
93 Simpson, G. G. 1944 Tempo and mode in evolution.
New York, NY: Columbia University Press.
94 de Queiroz, A. 2002 Contingent predictability in evol-
ution: key traits and diversification. Syst. Biol. 56,
917929. (doi:10.1080/10635150290102627)
95 Purvis, A. & Hector, A. 2000 Getting the measure
of biodiversity. Nature 405, 212219. (doi:10.1038/
35012221)
96 Barraclough, T. G., Harvey, P. H. & Nee, S. 1995
Sexual selection and taxonomic diversity in passerine
birds. Proc. R. Soc. Lond. B 259, 211 215. (doi:10.
1098/rspb.1995.0031)
97 Vrba, E. S. 1984 Evolutionary pattern and process in
the sister-group Alcelaphini-Aepycerotini (Mammalia:
Bovidae). In Living fossils (eds N. Eldredge & S. M.
Stanley), pp. 62– 79. New York, NY: Springer.
98 Darwin, C. 1859 On the origin of species by means of natu-
ral selection London, UK: John Murray.
99 Nee, S. 2005 Phylogenetic futures after the latest mass
extinction. In Phylogeny and conservation (eds A. Purvis,
J. L. Gittleman & T. Brooks), pp. 387399. Cambridge,
MA: Cambridge University Press.
100 Weir, J. T. & Schluter, D. 2007 The latitudinal gradient
in recent speciation and extinction rates of birds and
mammals. Science 315, 15741576. (doi:10.1126/
science.1135590)
101 Valente, L. M., Savolainen, V. & Vargas, P. 2010 Unpar-
alleled rates of species diversification in Europe.
Proc. R. Soc. B 277, 1489 1496. (doi:10.1098/rspb.
2009.2163)
102 Rangel, T. F. L. V. B., Diniz-Filho, J. A. F. & Colwell,
R. K. 2007 Species richness and evolutionary niche
dynamics: a spatial pattern-oriented simulation
experiment. Am. Nat. 170, 602 616. (doi:10.1086/
521315)
The shape of mammalian phylogeny A. Purvis et al. 2477
Phil. Trans. R. Soc. B (2011)
... Unbiased estimation of ''tip rates'' of species-specific speciation, l 0 , was previously demonstrated to require all extant taxa to be sampled or otherwise modeled (e.g., using the tip DR statistic 29,30 ). However, in mammals, it has only recently become possible to estimate tip rates robustly, thanks to new species-level timetrees that model uncertainty in topology and node ages. 2 Speciation and extinction rates through time have not yet been characterized across these ''backbone-and-patch'' mammal trees, 2 nor have they been used to evaluate deep-time questions relative to previous supertree-based inferences (e.g., Bininda-Emonds et al. 31 and Purvis et al. 32 ) or fossil mammal occurrences. 8 Herein, we apply this two-part framework to investigate a key deep-time question in the radiation of mammals: did early mammals exhibit a burst of lineage diversification coincident with, well before, or well after the Cretaceous-Paleogene (K-Pg) boundary $66 million years ago (Ma)? ...
... 2 The 24 branch-specific shifts in net diversification we recover in the backbone-and-patch timetrees (Figures 1, 3B, and 4A) compared to 27 shifts detected in the supertree (15 up-shifts, 12 down-shifts 32 ) and 9 up-shifts in the supermatrix timetree. 76 To their credit, Purvis et al. 32 analyzed only 1,335 bifurcating nodes in the supertree, which avoided some rate artifacts of polytomies. However, both studies returned overconfident estimates by treating the consensus phylogeny as known without error. ...
... However, both studies returned overconfident estimates by treating the consensus phylogeny as known without error. 77 If we only compare up-shifts, given the likely erasure of extant lineages as net diversification slows down, 17,78,79 we find three lineages shared by our study, Purvis et al., 32 and Yu et al. 76 : ...
Article
Full-text available
[revised from pre-print posted on Current Biology's SSRN server: https://papers.ssrn.com/sol3/papers.cfm?abstract_id=3761886] Reconstructing the tempo at which biodiversity arose is a fundamental goal of evolutionary biologists, yet the relative merits of evolutionary-rate estimates are debated based on whether they are derived from the fossil record or time-calibrated phylogenies (timetrees) of living species. Extinct lineages unsampled in timetrees are known to ‘pull’ speciation rates downward, but the temporal scale at which this bias matters is unclear. To investigate this problem, we compare mammalian diversification-rate signatures in a credible set of molecular timetrees (N=5,911 species, ca. 70% from DNA) to those in fossil genus durations (N=5,320). We use fossil extinction rates to correct or ‘push’ the timetree-based (pulled) speciation-rate estimates, finding a major pulse of speciation ca. 66-56 million years ago (Ma) between the Cretaceous-Paleogene (K-Pg) boundary and the Paleocene-Eocene Thermal Maximum (PETM). However, three-quarters of the K-Pg-to-PETM originating taxa did not leave modern descendants, indicating that this rate signature is realistically not detectable from extant lineages alone. For groups without substantial fossil records, thankfully all is not lost. Pushed and pulled speciation rates converge starting ca. 10 Ma, and are equal at the present-day when recent evolutionary processes can be estimated without bias using species-specific ‘tip’ rates of speciation. Clade-wide moments of tip rates also enable enriched inference, as the skewness of tip rates is shown to approximate a clade’s extent of past diversification-rate shifts. Molecular timetrees need fossil-correction to address deep-time questions, but they are sufficient for shallower time questions where extinctions are fewer.
... Next-generation sequencing has afforded unprecedented opportunities to generate pathogen genome sequences in a highly scalable manner, and theoretical tools have been developed to interrogate these data, largely through reconstructed phylogenetic trees. There has been considerable interest over the years in comparing the shapes of phylogenetic trees in order to understand evolutionary processes [1][2][3][4][5][6][7][8]. A tree's shape specifies its connectivity structure. ...
... The lengths of its branches typically reflect either the time or genetic distance between branching events. Following the observation that reconstructed evolutionary trees are more asymmetric than random models predict [9], there have been efforts to summarise tree asymmetry in trees reconstructed from data and relate it to predicted asymmetry in evolutionary and ecological models [4][5][6][7][10][11][12][13][14][15]. There is also interest in establishing whether taxa from two phylogenies might correspond to each other, for example in the context of parasites and hosts or fossils of different origins [16], and in comparing simulated trees with trees from data in epidemiology, for example using Approximate Bayesian Computation [17][18][19]. ...
Article
Full-text available
The shape of phylogenetic trees can be used to gain evolutionary insights. A tree’s shape specifies the connectivity of a tree, while its branch lengths reflect either the time or genetic distance between branching events; well-known measures of tree shape include the Colless and Sackin imbalance, which describe the asymmetry of a tree. In other contexts, network science has become an important paradigm for describing structural features of networks and using them to understand complex systems, ranging from protein interactions to social systems. Network science is thus a potential source of many novel ways to characterize tree shape, as trees are also networks. Here, we tailor tools from network science, including diameter, average path length, and betweenness, closeness, and eigenvector centrality, to summarize phylogenetic tree shapes. We thereby propose tree shape summaries that are complementary to both asymmetry and the frequencies of small configurations. These new statistics can be computed in linear time and scale well to describe the shapes of large trees. We apply these statistics, alongside some conventional tree statistics, to phylogenetic trees from three very different viruses (HIV, dengue fever and measles), from the same virus in different epidemiological scenarios (influenza A and HIV) and from simulation models known to produce trees with different shapes. Using mutual information and supervised learning algorithms, we find that the statistics adapted from network science perform as well as or better than conventional statistics. We describe their distributions and prove some basic results about their extreme values in a tree. We conclude that network science-based tree shape summaries are a promising addition to the toolkit of tree shape features. All our shape summaries, as well as functions to select the most discriminating ones for two sets of trees, are freely available as an R package at http://github.com/Leonardini/treeCentrality .
... Unbiased estimation of 'tip rates' of species-specific speciation, λ 0 , was previously demonstrated to require all extant taxa to be sampled or otherwise modeled (e.g., using the tip DR statistic 29,30 ). However, in mammals, it has only recently become possible to estimate tip rates robustly, thanks to new species-level timetrees that model uncertainty in topology and node ages. 2 Speciation and extinction rates through time have not yet been characterized across these 'backbone-and-patch' mammal trees, 2 nor have they been used to evaluate deep-time questions relative to previous supertree-based inferences (e.g., 31,32 ) or fossil mammal occurrences. 8 Herein, we apply this two-part framework to investigate a key deep-time question in the radiation of mammals: Did early mammals exhibit a burst of lineage diversification coincident with, well before, or well after the Cretaceous-Paleogene (K-Pg) boundary ca. ...
... 2 The 24 branch-specific shifts in net diversification we recover in the backbone-and-patch timetrees (Figures 1, 3B, 4A) compare to 27 shifts detected in the supertree (15 up-shifts, 12 down-shifts 32 ) and 9 up-shifts in the supermatrix timetree. 76 To their credit, Purvis et al. 32 analyzed only 1,335 bifurcating nodes in the supertree, which avoided some rate artifacts of polytomies. However, both studies returned overconfident estimates by treating the consensus phylogeny as known without error. ...
Article
Full-text available
Reconstructing the tempo at which biodiversity arose is a fundamental goal of evolutionary biologists, yet the relative merits of evolutionary-rate estimates are debated based on whether they are derived from the fossil record or time-calibrated phylogenies (timetrees) of living species. Extinct lineages unsampled in timetrees are known to “pull” speciation rates downward, but the temporal scale at which this bias matters is unclear. To investigate this problem, we compare mammalian diversification-rate signatures in a credible set of molecular timetrees (n = 5,911 species, ∼70% from DNA) to those in fossil genus durations (n = 5,320). We use fossil extinction rates to correct or “push” the timetree-based (pulled) speciation-rate estimates, finding a surge of speciation during the Paleocene (∼66–56 million years ago, Ma) between the Cretaceous-Paleogene (K-Pg) boundary and the Paleocene-Eocene Thermal Maximum (PETM). However, about two-thirds of the K-Pg-to-PETM originating taxa did not leave modern descendants, indicating that this rate signature is likely undetectable from extant lineages alone. For groups without substantial fossil records, thankfully all is not lost. Pushed and pulled speciation rates converge starting ∼10 Ma and are equal at the present day when recent evolutionary processes can be estimated without bias using species-specific “tip” rates of speciation. Clade-wide moments of tip rates also enable enriched inference, as the skewness of tip rates is shown to approximate a clade’s extent of past diversification-rate shifts. Molecular timetrees need fossil-correction to address deep-time questions, but they are sufficient for shallower time questions where extinctions are fewer.
... This pattern likely indicates a decreased rate of diversification, not of morphological change (bradytely). Such lineages can be called 'ancient singletons' and are relatively common among various taxa, including mammals (Purvis et al., 2011). Their existence is connected to a more general phenomenon: an overall imbalance of a phylogenetic tree that cannot be explained by existing simple (birth-death or pure birth) models and thus requires involvement of some biological process (Pigot et al., 2010). ...
Article
Gerbillinae is the second largest subfamily (after Murinae) in the Muridae family comprising ~16 genera and more than a hundred species. Phylogenetic relationships among the main taxa of gerbils are still not fully understood. In particular, a major issue is the phylogenetic position of the monotypic genus Ammodillus (endemic to the Horn of Africa), which according to morphological data may be one of the earliest offshoots from the gerbilline stem; however, its precise affiliations have been unknown due to the lack of genetic data. Here, by means of a multilocus dataset including one mitochondrial and five nuclear markers, we for the first time elucidated phylogenetic placement of Ammodillus and provided the most complete tribal phylogeny of Gerbillinae to date. Phylogenetic reconstructions robustly supported Ammodillus as a sister taxon to all other living gerbillines and suggested that the separation of the ammodile lineage dates back to the Middle/Late Miocene boundary (~11.7 Mya). The results are consiste nt with subdivision of the subfamily into four tribes: monotypic Ammodillini, Gerbillini (including Taterillus), Desmodilliscini (including Pachyuromys) and Gerbillurini. In the light of the new data, we discuss possible scenarios of Gerbillinae origin, highlight Ammodillus as a relatively ancient—albeit morphologically advanced—lineage that has never gained diversity and propose the term ‘ancient singleton’ for a taxon with a persistently low diversification rate.
... S ince Darwin, evolutionary biologists have sought to understand the processes that underlie large-scale diversifications, wherein large assemblages of closely related lineages evolve from a common ancestor [1][2][3][4] . Although several factors may mediate diversification, including geography, ecological opportunity, and key evolutionary innovations (KEI) 5 , the role of the ecological niche, particularly its climatic axes, remains poorly understood [6][7][8][9] . ...
Article
Full-text available
Large diversifications of species are known to occur unevenly across space and evolutionary lineages, but the relative importance of their driving mechanisms, such as climate, ecological opportunity and key evolutionary innovations (KEI), remains poorly understood. Here, we explore the remarkable diversification of rhacophorid frogs, which represent six percent of global amphibian diversity, utilize four distinct reproductive modes, and span a climatically variable area across mainland Asia, associated continental islands, and Africa. Using a complete species-level phylogeny, we find near-constant diversification rates but a highly uneven distribution of species richness. Montane regions on islands and some mainland regions have higher phylogenetic diversity and unique assemblages of taxa; we identify these as cool-wet refugia. Starting from a centre of origin, rhacophorids reached these distant refugia by adapting to new climatic conditions (‘niche evolution’-dominant), especially following the origin of KEIs such as terrestrial reproduction (in the Late Eocene) or by dispersal during periods of favourable climate (‘niche conservatism’-dominant). By examining climate, geographical and phylogenetic data, the diversification and evolution of rhacophorid frogs is explored
... Both the shape of a phylogenetic tree and its branch lengths carry evolutionary, epidemiological, or demographic information. For example, the shape of phylogenetic trees can reflect macroevolutionary patterns (Kirkpatrick and Slatkin 1993;Aldous 1996;Purvis et al. 2011), and the presence of selection can affect the shapes of trees (Maia et al. 2004;Dayarian and Shraiman 2014). Tree shapes and branch lengths of a phylogenetic tree relate to, for instance, the evolutionary dynamics of pathogens or transmission dynamics of epidemics (Grenfell et al. 2004;Poon et al. 2013;Volz et al. 2013;Plazzotta and Colijn 2016). ...
Article
Phylogenetic trees are a central tool in many areas of life science and medicine. They demonstrate evolutionary patterns among species, genes, and patterns of ancestry among sets of individuals. The tree shapes and branch lengths of phylogenetic trees encode evolutionary and epidemiological information. To extract information from tree shapes and branch lengths, representation and comparison methods for phylogenetic trees are needed. Representing and comparing tree shapes and branch lengths of phylogenetic trees are challenging, for a tree shape is unlabelled and can be displayed in numerous different forms, and branch lengths of a tree shape are specific to edges whose positions vary with respect to the displayed forms of the tree shape. In this paper, we introduce representation and comparison methods for rooted unlabelled phylogenetic trees based on a tree lattice that serves as a coordinate system for rooted binary trees with branch lengths and a graph polynomial that fully characterizes tree shapes. We show that the introduced tree representations and metrics provide distance-based likelihood-free methods for tree clustering, parameter estimation and model selection, and apply the methods to analyze phylogenies reconstructed from virus sequences.
... wherein large assemblages of closely related lineages evolve from a common ancestor, have 42 inspired evolutionary analyses [1][2][3][4] . Although several factors may mediate diversification, 43 including geography, ecological opportunity, and key innovations 5 , the role of the ecological 44 niche, particularly its climatic axes, remains poorly understood 6-9 . ...
Preprint
Full-text available
24 Although large diversifications of species occur unevenly across space and evolutionary 25 lineages, the relative importance of their driving mechanisms, such as climate, ecological 26 opportunity and key innovations, remains poorly understood. Here, we explore the 27 remarkable diversification of rhacophorid frogs, which represent six percent of global 28 amphibian diversity, utilize four distinct reproductive modes, and span a climatically variable 29 area across mainland Asia, associated continental islands, and Africa. Using a complete 30 species-level phylogeny, we find near-constant diversification rates but a highly uneven 31 distribution of species richness. Montane regions on islands and some mainland regions have 32 higher phylogenetic diversity and unique assemblages of taxa; we identify these as cool-wet 33 refugia. Starting from a centre of origin, rhacophorids reached these distant refugia by 34 adapting to new climatic conditions ('niche evolution'-dominant), especially following the 35 origin of key innovations such as terrestrial reproduction (in the Late Eocene) or by dispersal 36
Article
Full-text available
A central focus of evolutionary biology is to understand species diversity by studying how they arrived at their current geographic distributions. The biogeography of the Old World tree frogs in the family Rhacophoridae has been extensively studied suggesting an early Paleogene origin in Asia (out of Asia hypothesis) with alternative hypotheses in play. However, these alternative hypotheses especially considering adjacency of biogeographical regions and plate tectonics have not been tested empirically. Here using a comprehensive time calibrated phylogeny and constrained dispersal multipliers we studied the biogeographical history and diversification of Rhacophoridae, distributed in five biogeographical regions. Five hypotheses suggesting different centers of origin, and additional hypotheses considering adjacency and plate tectonics were tested to delineate the biogeographical history of Rhacophoridae. In addition, various diversification models that accounted for factors such as lineage isolation time, diversity-dependence, paleotemperatures, speciation and extinction rates were also used to test patterns of diversification. Results confirmed an East/ Southeast Asian center of origin (out of Asia), with dispersal likely mediated by plate tectonics and adjacency of biogeographical regions, which could be linked to periodic sea level fluctuations and climate changes. The best-fitting diversification models explained diversification through lineage isolation time and paleotemperature regimes, while diversity-dependent models had low support. Speciation was linearly dependent on time and paleotemperatures, while extinction rates were exponentially dependent on time and linearly dependent on paleotemperature. Our findings demonstrate that variable extinction rates contribute towards maintaining a constant diversification rate for rhacophorids. We discuss that episodic major extinction events on the Indian Plate may have played a major role in shaping the early evolution of Rhacophoridae thus favoring an Out of Asia hypothesis in the empirical models. However, current biogeographic models may not be sufficient to explain the origin of Rhacophoridae, as multiple factors are likely at play. Ellepola G and Meegaskumbura M (2023) Diversification and biogeography of Rhacophoridae-a model testing approach.
Article
Full-text available
Did the end-Cretaceous mass extinction event, by eliminating non-avian dinosaurs and most of the existing fauna, trigger the evolutionary radiation of present-day mammals? Here we construct, date and analyse a species-level phylogeny of nearly all extant Mammalia to bring a new perspective to this question. Our analyses of how extant lineages accumulated through time show that net per-lineage diversification rates barely changed across the Cretaceous/Tertiary boundary. Instead, these rates spiked significantly with the origins of the currently recognized placental superorders and orders approximately 93 million years ago, before falling and remaining low until accelerating again throughout the Eocene and Oligocene epochs. Our results show that the phylogenetic 'fuses' leading to the explosion of extant placental orders are not only very much longer than suspected previously, but also challenge the hypothesis that the end-Cretaceous mass extinction event had a major, direct influence on the diversification of today's mammals. Molecular data and the fossil record can give conflicting views of the evolutionary past. For instance, empirical palaeontological evidence by itself tends to favour the 'explosive model' of diversification for extant placental mammals 1 , in which the orders with living representatives both originated and rapidly diversified soon after the Cretaceous/Tertiary (K/T) mass extinction event that eliminated non-avian dinosaurs and many other, mostly marine 2 , taxa 65.5 million years (Myr) ago 1,3,4. By contrast, molecular data consistently push most origins of the same orders back into the Late Cretaceous period 5-9 , leading to alternative scenarios in which placental line-ages persist at low diversity for some period of time after their initial origins ('phylogenetic fuses'; see ref. 10) before undergoing evolutionary explosions 1,11. Principal among these scenarios is the 'long-fuse model' 1 , which postulates an extended lag between the Cretaceous origins of the orders and the first split among their living representatives (crown groups) immediately after the K/T boundary 8. Some older molecular studies advocate a 'short-fuse model' of diversification 1 , where even the basal crown-group divergences within some of the larger placental orders occur well within the Cretaceous period 5-7. A partial molecular phylogeny emphasizing divergences among placental orders suggested that over 20 lineages with extant descendants (henceforth, 'extant lineages') survived the K/T boundary 8. However, the total number of extant lineages that pre-date the extinction event and whether or not they radiated immediately after it remain unknown. The fossil record alone does not provide direct answers to these questions. It does reveal a strong pulse of diversification in stem eutherians immediately after the K/T boundary 4,12 , but few of the known Palaeocene taxa can be placed securely within the crown groups of extant orders comprising Placentalia 4. The latter only rise to prominence in fossils known from the Early Eocene epoch onwards (,50 Myr ago) after a major faunal reorganization 4,13,14. The geographical patchiness of the record complicates interpretations of this near-absence of Palaeocene crown-group fossils 14-16 : were these clades radiating throughout the Palaeocene epoch in parts of the world where the fossil record is less well known; had they not yet originated; or did they have very long fuses, remaining at low diversity until the major turnover at the start of the Eocene epoch? The pattern of diversification rates through time, to which little attention has been paid so far, might hold the key to answering these questions. If the Cretaceous fauna inhibited mammalian diversification , as is commonly assumed 1 , and all mammalian lineages were able to radiate after their extinction, then there should be a significant increase in the net per-lineage rate of extant mammalian diversification , r (the difference between the per-lineage speciation and extinction rates), immediately after the K/T mass extinction. This hypothesis, along with the explosive, long-and short-fuse models, can be tested using densely sampled phylogenies of extant species, which contain information about the history of their diversification rates 17-20. Using modern supertree algorithms 21,22 , we construct the first virtually complete species-level phylogeny of extant mammals from over 2,500 partial estimates, and estimate divergence times (with confidence intervals) throughout it using a 66-gene alignment in conjunction with 30 cladistically robust fossil calibration points. Our analyses of the supertree indicate that the principal splits underlying the diversification of the extant lineages occurred (1) from 100-85 Myr ago with the origins of the extant orders, and (2) in or after the Early Eocene (agreeing with the upturn in their diversity known from the fossil record 4,13,14), but not immediately after the K/T boundary, where diversification rates are unchanged. Our findings-that more extant placental lineages survived the K/T boundary than previously recognized and that fewer arose immediately after it than previously suspected-extend the phylogenetic fuses of many extant orders and indicate that the end-Cretaceous mass extinction event had, at best, a minor role in driving the diversification of the present-day mam-malian lineages. A supertree with divergence times for extant mammals The supertree contains 4,510 of the 4,554 extant species recorded in ref. 23, making it 99.0% complete at the species level (Fig. 1; see also
Chapter
As progress in cell developmental biology carries on at a breakneck speed, new techniques constantly arise to plug the gaps left by traditional strategies. Cellular Interactions in Development provides detailed discussion and protocols of some of these new techniques, which allow the manipulation of developing organisms such as Drosophila or plants, when and where cells interact with each other to influence their development. The book looks at the really exciting innovations of the identification and functional test of molecules which control these cellular behaviours. The book also describes a number of new ways of hunting for these important proteins involved in cellular communication. A fully comprehensive manual which will prove indispensable to researchers in the fields of cell, developmental, and molecular biology.
Article
Phylogenetic trees represent the evolutionary histories of lineages and so bear the impression of the evolutionary forces that gave rise to those lineages. Advances in molecular and computational techniques continually increase the number and size of our phylogenetic estimates. In the 1990s, both we [41] and Purvis [52] surveyed the two main aspects of phylogenetic tree pattern: variation in realized diversification rate among contemporaneous lineages, and changes in realized diversification rates through time. The techniques highlighted in these reviews have been used very successfully (see, e.g. [4, 10, 11, 62, 63]).
Chapter
Phylogeny is a potentially powerful tool for conserving biodiversity. This book explores how it can be used to tackle questions of great practical importance and urgency for conservation. Using case studies from many different taxa and regions of the world, the volume evaluates how useful phylogeny is in understanding the processes that have generated today's diversity and the processes that now threaten it. The novelty of many of the applications, the increasing ease with which phylogenies can be generated, the urgency with which conservation decisions have to be made and the need to make decisions that are as good as possible together make this volume a timely and important synthesis which will be of great value to researchers, practitioners and policy-makers alike.