scispace - formally typeset
Open AccessJournal ArticleDOI

Genomic architecture and introgression shape a butterfly radiation

Reads0
Chats0
TLDR
Tests to distinguish incomplete lineage sorting from introgression indicate that gene flow has obscured several ancient phylogenetic relationships in this group over large swathes of the genome, and a hitherto unknown inversion that traps a color pattern switch locus is identified.
Abstract
We used 20 de novo genome assemblies to probe the speciation history and architecture of gene flow in rapidly radiating Heliconius butterflies. Our tests to distinguish incomplete lineage sorting from introgression indicate that gene flow has obscured several ancient phylogenetic relationships in this group over large swathes of the genome. Introgressed loci are underrepresented in low-recombination and gene-rich regions, consistent with the purging of foreign alleles more tightly linked to incompatibility loci. Here, we identify a hitherto unknown inversion that traps a color pattern switch locus. We infer that this inversion was transferred between lineages by introgression and is convergent with a similar rearrangement in another part of the genus. These multiple de novo genome sequences enable improved understanding of the importance of introgression and selective processes in adaptive radiation.

read more

Content maybe subject to copyright    Report

BUTTERFLY GENOMICS
Genomic architecture and introgression shape a
butterfly radiation
Nathaniel B. Edelman
1
*, Paul B. Frandsen
2,3
, Michael Miyagi
1
, Bernardo Clavijo
4
, John Davey
5,20
,
Rebecca B. Dikow
3
, Gonzalo García-Accinelli
4
, Steven M. Van Belleghem
6
, Nick Patterson
7,8
,
Daniel E. Neafsey
8,9
, Richard Challis
10
, Sujai Kumar
11
, Gilson R. P. Moreira
12
, Camilo Salazar
13
,
Mathieu Chouteau
14
, Brian A. Counterman
15
, Riccardo Papa
6,16
, Mark Blaxter
10
, Robert D. Reed
17
,
Kanchon K. Dasmahapatra
5
, Marcus Kronforst
18
, Mathieu Joron
19
, Chris D. Jiggins
20
,
W. Owen McMillan
21
, Federica Di Palma
4
, Andrew J. Blumberg
22
, John Wakeley
1
,
David Jaffe
8,23
, James Mallet
1
*
We used 20 de novo genome assemblies to probe the speciation history and architecture of gene
flow in rapidly radiating Heliconius butterflies. Our tests to distinguish incomplete lineage sorting
from introgression indicate that gene flow has obscured several ancient phylogenetic relationships
in this group over large swathes of the genome. Introgressed loci are underrepresented in
low-recombination and gene-rich regions, consistent with the purging of foreign alleles more tightly
linked to incompatibility loci. Here, we identify a hitherto unknown inversion that traps a color
pattern switch locus. We infer that this inversion was transferred between lineages by introgression
and is convergent with a similar rearrangement in another part of the genus. These multiple de novo
genome sequences enable improved understanding of the importance of introgression and selective
processes in adaptive radiation.
A
daptive radiations play a fundamental
role in generating biodiversity. Initiated
by key innovations and ecological op-
portunity, radiation is fueled by niche
competition that promotes rapid diver-
sification of species (1). Reticulate evolution
may enhance radiation by introducing genetic
variation, enabling rapidly emerging popula-
tions to take advantage of new ecological op-
portunities (2, 3). Diverging from its sister genus
Eueides ~1 2 million years (My) ago, Heliconius
radiated in a burst of speciation in the last ~5 My
(4). Introgressio n is well known in Heliconius,
with widespread reticulate evolution across the
genus (5), although this has been disputed (6).
Nonetheless, how introgression varies across
the genome is known only in one pair of sister
lineages (7, 8). Here, we use multiple de novo
whole-genome assemblies to improve the reso-
lution of introgression, incomplete lineage sort-
ing (ILS), and genome architecture in deeper
branches of the Heliconius phylogeny.
Phylogenetic analysis
Wegenerated20denovogenomeassemblies
for species in both major Heliconius subclades
and three additional genera of Heliconiini. We
then aligned the 16 highest-quality Heliconiini
assemblies to two Heliconius reference genomes
and seven other Lepidoptera genomes, result-
ing in an alignment of 25 taxa (9). De novo
assembly provides superior sequence infor-
mation for low-complexity regions, allows for
discovery of structural rearrangements, and
improves alignment of evolutionarily distant
clades (10). Other studies in Heliconius have
shown a high level of phylogenetic discordance,
arguably a result of rampant introgression
(4, 5). We attempte d to reconstruct a bifurcat-
ing species tree by estimating relationships
using protein-coding genes, conserved coding
regions, and conserved noncoding regions. We
generated phylogenies with coalescent-based
and concatenation approaches using both the
full Lepidoptera alignment and a restricted,
Heliconiini-only subalignment. These topolo-
gies were largely congruent among analytical
approaches, but weakly supported nodes were
resolved inconsistently. These approaches
therefore failed to resolve the phylogeny of
Helicon ius as a simple bifurcating tree (Fig. 1A
and fig. S20).
To determine whether hybridization was a
causeofthespeciestreeuncertainty,wecal-
culated Pattersons D statistics (11) for every
triplet of the 13 Heliconius species using a
member of the sister genus, Eueides tales,as
the outgroup. In 201 of 286 triplets, we observed
values significantly different from zero based
on block-jackknifing, demonstrating strong
evidence for introgression (fig. S53). However ,
these tests alone yield little quantitative infor-
mation about admixture. We therefore used
phyloNet (12) to infer reticulate phylogenetic
networks of these species based on random
samplesof10010-kbwindowsacrossthealign-
ment. For each sample, we coestimated all
100 regional gene trees and the overall species
network in parallel (12). To improve alignments,
we analyzed the melpomene-silvaniform group
with respect to the Helic onius melpomene
Hmel2.5 assembly (13)andtheerato-sara group
with respect to the H. erato demophoon v1 as-
sembly (9, 14). Most species exhibited an ad-
mixture event at some point in their history
using this method; we confirmed extensive re-
ticulation among silvaniform species and dis-
covered major gene-flow events in the erato-sara
clade. On the basis of these results, we pro-
posethereticulatephylogeniesshowninFig.
1, B to C.
Correlation of local ancestry with
genome architecture
We next analyzed the distribution of tree topol-
ogies across the genome, again treating each
major clade separately and using its respective
reference genome. The melpomene-silvaniform
group lacked topological consensus, unsurpris-
ingly because introgression, especially of key
mimicry loci, is well known in this clade (15).
The most common tree topology was found
in only 4.3% of windows, with an additional
14 topologies appearing in 1.0 to 3.4% of win-
dows (fig. S19 to S21). By contrast, we here focus
on the erato-sara group, in which two to-
pologies dominate (Fig. 2). One (Fig. 2B, Tree 2)
matched our bifurcating consensus topology
(Fig. 1A) and a re cently published tree (4),
whereas the other (Tree 1) differs in that it
places H. hecalesia and H. telesiphe as sisters.
Regions with local topologies discordant
from the species tree may have arisen through
introgression or ILS. To make within-topology
RESEARCH
Edelman et al., Science 366, 594599 (2019) 1 November 2019 1of6
1
Department of Organismic and Evolutionary Biology, Harvard University, Cambridge, MA 02138, USA.
2
Department of Plant and Wildlife Sciences, Brigham Young University, Provo, UT 84602,
USA.
3
Data Science Lab, Office of the Chief Information Officer, Smithsonian Institution, Washington, DC 20560, USA.
4
Earlham Institute, Norwich Research Park, Norwich NR4 7UZ, UK.
5
Bioscience Technology Facility, Department of Biology, University of York, York YO10 5DD, UK.
6
Department of Biology, University of Puerto Rico, Río Piedras Campus, San Juan, PR 00931-
3360, Puerto Rico.
7
Department of Human Evolutionary Biology, Harvard University, Cambridge, MA 02138, USA.
8
Broad Institute of MIT and Harva rd, Cambridge, MA, 02142 USA.
9
Harvard TH
Chan School of Public Health, Boston, MA 02115, USA.
10
Wellcome Sanger Institute, Wellcome Genome Campus, Cambridge CB10 1SA, UK.
11
Institute of Evolutionary Biology, University of
Edinburgh, Edinburgh EH9 3JT, UK.
12
Departamento de Zoologia, Universidade Federal do Rio Grande do Sul, Porto Alegre, 91501-970 Brasil.
13
Biology Program, Faculty of Natural Sciences and
Mathematics, Universidad del Rosario, Carrera 24, No. 63C-69, Bogotá D.C. 111221, Colombia.
14
Laboratoire Ecologie, Evolution, Interactions des Systèmes Amazoniens (LEEISA), USR 3456,
Université De Guyane, CNRS Guyane, 275 Route de Montabo, 97334 Cayenne, French Guiana.
15
Department of Biological Sciences, Mississippi State University, Starkville, MS 39762, USA.
16
Molecular Sciences and Research Center, University of Puerto Rico, San Juan, PR 00931-3360, Puerto Rico.
17
Department of Ecology and Evolutionary Biology, Cornell University, Ithaca, NY
14853, USA.
18
Department of Ecology and Evolution, University of Chicago, Chicago, IL 60637, USA.
19
CEFE, CNRS, Université de Montpellier, Univ ersité Paul Valéry Montpellier 3, EPHE, IRD,
34090 Montpellier, France.
20
Department of Zoology, University of Cambridge, Cambridge CB2 3EJ, UK.
21
Smithsonian Tropical Research Institute, Apartado 0843-03092 Panamá, Panama.
22
Department of Mathematics, University of Texas, Austin, TX 78712, USA.
23
10x Genomics, Pleasanton, CA 94566, USA.
*Corresponding author. Email: nedelman@g.harvard.edu (N.B.E.); jmallet@oeb.harvard.edu (J.M.)
on November 23, 2020 http://science.sciencemag.org/Downloaded from

locus-by-locus inferences, we developed a statis-
tical test to distinguish between ILS and intro-
gression based on the distribution of internal
branch lengths among windows for a given
three-taxon subtree, conditional on its topol-
ogy. We call this method quantifying intro-
gression via branch lengths (QuIBL). In the
absence of introgression, we expect internal
branch lengths of triplet topologies discordant
with the species tree (due to ILS) to be expo-
nentially distributed. However, if introgression
has occurred, then their distribution should
have that same exponential component but also
include an additional component with a non-
zero mode corresponding to the time between
the introgression event and the most recent
common ancestor of all three species (9). Like
other tree-based methods, QuIBL is potentially
sensitive to the assumption that each tree is
inferred from loci with limited internal recom-
bination (fig. S75). We therefore chose small
(5-kb) windows to reduce the probability of
intralocus recombination breakpoints.
For every triplet in the erato-sara clade, we
calculated the likelihood that the distribution
of internal branch lengths is consistent with
introgression or with ILS only. We formally
distinguished between these two models using
a Bayesian information criterion (BIC) test with
a strict cutoff of DBIC > 10. Consistent with our
results from D statistics, we found that 13 of
20 triplets have evidence for introgression
(table S13). For example, using QuIBL on the
triplet H. eratoH. hecalesiaH. telesiphe,we
infer that 76% of discordant loci, or 38% of
all loci genome-wide, are introgressed. Aver-
aging over all triplets, we infer that 71% (67%
with BIC filtering) of loci with discordant
gene trees have a history of introgression, or
20% (19 % with BIC filter ing) of all triple t
loci, indicating a broad signal of introgression
throughout the clade [Eq. 7.7, table S13; see
(9) for additional discussion].
In hybrid populations, individuals have ge-
nomic regions that originate from different
speciesandmaybeincompatiblewiththe
recipient genome or with their environ-
ment ( 16). Linked selection causes harm-
less or even beneficial introgressed loci to
be removed along with these deleterious loci
if they are tightly linked; this effect depends
on the strength of selection and the local re-
combination rate (17, 18). We therefore expect
introgressed loci to be enriched in regions
where selection is likely to be weak, such as
gene deserts or regions of high recombina-
tion, where harmless introgressed loci more
readily recombine away from linked incom-
patibility loci.
In Heliconius, even distant species such as
H. erat o and H. melpomene have the same
number of broadly collinear chromosomes (13),
facilitating direct comparisons among species.
Furthermor e, each chromosome in Heliconius
has approximately one crossover per meiosis
i
n males (there is no crossing over in female
Heliconius)(14 , 19). Chromosomes vary in
length, and chromosome size is inversely
proportional to recombination rate per base
pair (8, 13). We found a strong correlation
between the fraction of windows in each
chromosome that show a given topology and
physical chromosome length (Fig. 3A). Such
relationships exist for all eight trees in Fig. 2B
(9), but we focus here on the two most common
trees: Tree 1 has a strongly negative correlation
with chromosome size (r
2
= 0.883, t=11.7, 18 df,
p < 0.0001), whereas Tree 2 (concordant with
our inferred species tree) has a positive cor-
relation (r
2
=0.726,t =6.9,18df,p < 0.0001). Re-
sults from QuIBL indicate that 94% of windows
that recover a Tree 1 triplet topology are con-
sistent with introgression (fig . S70 and table
S13). The Z (sex) chromosome 21 is strongly
Edelman et al., Science 366, 594599 (2019) 1 November 2019 2of6
Fig. 1. Phylogeny and phylogenetic networks of Heliconius do not support
a bifurcating tree. (A) All nodes resolved in a majority of species trees are
shown in this cladogram (heavy black lines), whereas the poorly resolved
silvaniform clade is collapsed as a polytomy (fig. S20). The 500 colored trees
were sampled from 10-kb nonoverlapping windows and constructed with
maximum likelihood. (B and C) High-confidence tree structure (black) and
introgression events (red) are shown as solid lines. Dashed red lines indicate
weakly supported introgression events. Gray branch ends are cosmetic. The
melpomene-silvaniform clade is shown in (B) and the erato-sara clade in
(C). Euclidean lengths of solid black lines are proportional to genetic distance
along the branches. Scale bars are in units of substitutions per site. Breaks at the base
in (B) indicate that the branch leading to H. doris has been shortened for display .
RESEARCH | RESEARCH ARTICLE
on November 23, 2020 http://science.sciencemag.org/Downloaded from

Edelman et al., Science 366, 594599 (2019) 1 November 2019 3of6
Fig. 2. Local evolutionary history in t he
erato-sara clade is h eterogeneous across
the genome. (A) Each bar represents a
chromosome, in terms of the H. erato refe rence
(14). Colored bands represent t ree topologies
of each 50-kb window; colors correspond
to the topologies in (B), with black regions
showing missing data. (B) The eight most
common trees. The value in the top left corner
is the percentage of all 50-kb windows that
recover that topology. (C) Each histogram
corresponds to the topology of the same
color in (B) and shows the distribution of the
number of consecutive 50-kb windows with
that topology. Arrows indicate long blocks
in inversions.
Fig. 3. Chromosomal architecture is strongly
correlated with local topology. Tree 1 is shown in
red and Tree 2 is shown in blue, as in Fig. 2. (A) Tree
1 shows a negative relationship with chromosome
size, whereas Tree 2 shows a positive relationship.
Lines are linear regressions with chromosome 21
excluded. Numbers along the top indicate chromo-
some number. (B) Each chromosome was divided
into 10 equally sized bins, and the occupancy of each
topology in each bin was calculated as the number of
windows that recovered the topology in the bin
divided by the number of windows that recovered the
topology in the chromosome. (C) Windows are
binned by recombination rate, and boxes show
the fraction of each tree in each bin for each
chromosome separately. Numbers above boxes are
the numbers of windows in each bin. (D) Boxes
showing the relationship of tree topology with coding
density. Asterisk denotes significance at the 5% level
(paired t test, p < 0.025). In all boxplots, the
central line is the median, box edges are first and
third quartile, and whiskers extend to the largest
value no farther than 1.5 × (interquartile range).
RESEARCH | RESEARCH ARTICLE
on November 23, 2020 http://science.sciencemag.org/Downloaded from

enriched for Tree 2, suggesting that it may
harbor more incompatibility loci than auto-
somes. Interspecific hybrid females in Heliconius
are often sterile, conforming to Haldanesrule,
and sex chromosomes have been implicated
as being particularly important in generating
incompatibiliti es (8, 2024).
To test whether th e pattern that we ob-
served among chromosomes is related to dif-
ferences in recom bination, we investigated
the relationship between recombination rate
and tree topology within chromosomes. The
recombination rate declines at the ends of
chromosomes (fig. S85), and the species tree
(Tree2)ismoreabundantinthoseregions
Edelman et al., Science 366, 594599 (2019) 1 November 2019 4of6
Fig. 4. Parallel evolution of a major inversion at the cortex supergene
locus. (A) Map of 1.7-Mb region on chromosome 15. Coordinates are in terms of
Hmel 2.5 and ticks are in Mb. Tree topology colors correspond to those in Fig. 2.
Genes are shown as black rectangles; cortex is highlighted in yellow. Each line
shows the mapping of a single contig. Aligned sections of each contig are shown
as thick bars, whereas unaligned sections are shown as dotted lines. Arrows
indicate the strand of the alignment. The H. erato group breakpoints are shown
with red vertical lines and the H. numata breakpoints are shown with green
vertical lines. (B) Evolutionary hypotheses consistent with the topology observed
in this inversion in the context of the previously estimated phylogenetic network.
The three species used in the triplet gene tree method, H. erato , H. telesiphe,
and H. sara, are shown as black lines; lineages not included are shown as gray
lines. (C) Histogram of internal branch lengths (T
2
) in windows with the topology
H. erato (H. telesiphe, H. sara). The inferred ILS distribution is shown as a
dashed line, and the inferred introgression distribution is shown as a dotted line.
The average internal branch length in the inversion is shown as a green vertical
line. (D) Histogram of normalized D
XY
(T
3
) between H. telesiphe and H. sara.
Mean normalized D
XY
in the inversion is shown as a green vertical line.
RESEARCH | RESEARCH ARTICLE
on November 23, 2020 http://science.sciencemag.org/Downloaded from

(Fig. 3B). In addition, when windows were
grouped b y local recombination rate calcu-
lated from population genetic data (9, 14),
we obser ved a strong relationship with the
recovered topology (Fig. 3C). Finally, we ob-
served a minor enrichment of Tree 1 in re-
gions of very low gene density, but this effect
was weak (Fig. 3D) compared with that of
recombination. Taken together, these results
show that tighter linkage on longer chromo-
somes and in lower recombination re gions
within chromosomes leads to removal of more
introgressed variation in those regions. This
very strong correlation is consistent with a
highly polygenic architecture of incompatibilities
between species.
Introgression of a convergent inversion
The topology block size distribution in the erato
clade generally decayed exponentially (Fig. 2C),
but two unusually long blocks contained minor
topologies: one on chromosome 2 (Tree 3, com-
posed of three sub-blocks) and the other on
chromosome 15 (Tree 4). Our study of the
~3-Mb topology block on chromosome 2 con-
firms an earlier finding of an inversion in
H. erat o (13), and we show here t hat its rare
topology is most likely explained by ILS, in-
cluding a long period of ancestral polymor-
phism (fig. S95).
The topology block on chromosome 15 is of
particular interest because it spans cortex,a
genetic hotspot of wing color pattern diver-
sity in Lepidoptera (25, 26). We hypothesized
that this block could be an inversion, as in
H. numata, where the P
1
supergene inver-
sion polymorphism around cortex controls
color pattern switching among mimicry morphs
(27). This block recovers H. telesiphe and
H. hecalesia as a m onophyle tic subclade,
which together are sisters t o the sara clade
(Fig.2B,Tree4).Wesearchedourdenovo
assemblies for contigs that mapped across
topology transitions. Taking H. melpomene as
the standard arrangement, we found clear in-
version breakpoints in H. telesiphe, H. hecalesia,
H. sara,andH. demeter. Conversely, H. erato,
H. himera,andE. tales all contain contigs
that map in their entirety across the break-
points (Fig. 4A), implying that they have the
ancestral H. melpomene arrangement.
This chromosome 15 inversion covers almost
exactly the same region as the 400-kb P
1
in-
version in H. numata (25, 27, 28). However,
de novo contigs from our H. numata assembly
show that the breakpoints of P
1
are close to but
not identical to those of the inversion in the
erato clade (Fig. 4A). Furthermore, in topol-
ogies for H. numata, H. telesiphe, H. erato,and
E. tales across chromosome 15, not a single
window recovered H. numata and H. telesiphe
as a monophyletic subclade, as would be ex-
pected if the erato group inversion were ho-
mologous to P
1
in H. numata.
We used QuIBL with the triplet H. erato
H. telesipheH. sara to elucidate the evo-
lutionary history of this inversion. A small
internal branch would suggest ILS, whereas
a large internal branch would be more con-
sistent with introgression (Fig. 4B). The average
internal branch length in the inversion was
much longer than the genome-wide average,
corresponding to a 79% probability of intro-
gression (Fig. 4C). If the inversion were poly-
morphic in the ancestralpopulationforsome
time, then we could also reco ver a similarl y
long internal branch (Fig. 4B, center). We
distinguished between this longer-term poly-
morphic scenario and introgression by com-
paring the genetic distance (D
XY
) b etween
H. telesiphe and H. sara,representedbyT
3
in Fig. 4B. Normalized D
XY
(as in fig. S95)
within the inversion is ~25% less than in
the rest of the genome. Given that this is a
large genomic block, introgression is therefore
the most parsimonious explanation for the evo-
lutionary history of the inversion (Fig. 4D) (29).
Discussion
Species involved in rapid radiations are prone
to hybridization because of frequent geograph-
ical overlap with closely related taxa. In both
the melpomene and erato clades of Helicon ius,
introgression has overwritten the original bi-
furcation history of several species across large
swathes of the genome, a pattern also observed
in Anopheles mosquitos (30). This observa-
tion is also consistent with genomic analysis
of other rapid radiations characterized by
widespread hybridization and introgression,
including Darwinsfinches(2)andAfricancich-
lids (31). In other radiations, the role of in-
trogressionislessclear:inTamias chipmunks,
widespread introgression of mitochondrial DNA
was identified, in contrast to an absence of evi-
dence for nuclear gene flow (32). With few
genomic comparisons available to date, it is
perhaps too early to say whether introgres-
sion is a major feature of adaptive radiations
in general, but evidence thus far suggests this
to be the case.
Our results raise the question of why some
genomic regions cross species boundaries and
others do not. In the erato clade, we found a
strong correlation between recombination
rate and introgression probability. Similar
associations with topology also exist between
sister species in the melpomene clade (8). As-
sociations between recombination and in-
trogression in hybridizing populations of
fishes and monkey flowers (Mimulus spp.) sup-
port the role of linked selection on a highly
polygenic landscape of interspecific incom-
patibilities (18, 33, 34). Our results establish
that this relationship persists and may indeed
be strengthened with time since introgres-
sion. While hybridization is ongoing, many
introgressed blocks are constantly reintroduced
into the population. If linked to weakly dele-
terious alleles, introgressed loci will finally be
purged by linked selection only long after
introgression ceases.
Recombination rate alone cannot account
for differentia l introgression, so we must delve
into specific regions to elucidate their function
and relevance to speciation. It is critical, there-
fore, to have tools that can confidently identify
introgressed loci, and much effort has gone
into developing such methods (11, 35). Our test
using internal branch lengths in triplet gene
trees is based in coalescent theory and takes
advantage of the discriminator y power of a
property of gene trees not explicitly accounted
for by other methods. QuIBL allows us to as-
sess probability of introgression for each locus
in each species triplet (9). Here, we used this
method to identify the evolutionary origin of a
convergent inversion that has undergone mul-
tiple independent introgression events and to
show that genomic regions with discordant to-
pologies arose mostly through hybridization.
Just
as sex aids adaptation w ithin species,
occasional introgression and recombination
among species can have major long-term ef-
fects on the genome, contributing variation
that could fuel rapid adaptive divergence and
radiation.
REFERENCES AND NOTES
1. D. Schluter, The Ecology of Adaptive Radiation (Oxford Univ.
Press, 2000).
2. S. Lamichhaney et al., Nature 518, 371375 (2015).
3. J. B. Pease, D. C. Haak, M. W. Hahn, L. C. Moyle, PLOS Biol. 14,
e1002379 (2016).
4. K. M. Kozak et al., Syst. Biol. 64, 505524 (2015).
5. K. M. Kozak, O. McMillan, M. Joron, C. D. Jiggins, Genome-wide
admixture is common across the Heliconius radiation.
bioRxiv 414201 [Preprint]. 11 September 2018.
6. A. V. Z. Brower, I. J. G. Orduña, Cladistics 34, 151166 (2018).
7. S. H. Martin et al., Genome Res. 23 , 18171828 (2013).
8. S. H. Martin, J. W. Davey, C. Salazar, C. D. Jiggins, PLOS Biol.
17, e2006288 (2019).
9. Materials and methods are available as supplementary
materials.
10. A. G. Clark et al., Nature 450, 203218 (2007).
11. N. Patterson et al., Genetics 192, 10651093 (2012).
12. D. Wen, L. Nakhleh, Syst. Biol. 61, 439457 (2017).
13. J. W. Davey et al., Evol. Lett. 1, 138154 (2017).
14. S. M. Van Belleghem et al., Nat. Ecol. Evol. 1, 52 (2017).
15. C. D. Jiggins, The Ecology and Evolution of Heliconius
Butterflies (Oxford Univ. Press, 2017).
16. J. A. Coyne, H. A. Orr, Speciation (Sinauer, 2004).
17. D. J. Begun, C. F. Aquadro, Nature 356, 519520 (1992).
18. M. Schumer et al., Science
360,
656660 (2018).
19. J.W.Daveyet al., G3: Genes Genomes Genetics 6,695708 (2016 ).
20. T. Dobzhansky, Genetics and the Origin of Species (Columbia
Univ. Press, 1937).
21. H. A. Orr, M. Turelli, Genetics 143, 613616 (1996).
22. C. D. Jiggins et al., Evolution 55, 16311638 (2001).
23. R. E. Naisbit, C. D. Jiggins, M. Linares, C. Salazar, J. Mallet,
Genetics 161, 15171526 (2002).
24. S. M. Van Belleghem et al., Mol. Ecol. 27, 38523872 (2018).
25. M. Joron et al., PLOS Biol. 4, e303e310 (2006).
26. N. J. Nadeau et al., Nature 534, 106110 (2016).
27. P. Jay et al., Curr. Biol. 28, 18391845.e3 (2018).
28. M. Joron et al., Nature 477, 203206 (2011).
29. F. Roda, F. K. Mendes, M. W. Hahn, R. Hopkins, Mol. Ecol. 26,
23172330 (2017).
30. M. C. Fontaine et al., Science 347, 1258524 (2015).
31. J. I. Meier et al., Nat. Commun. 8, 14363 (2017).
32. J. M. Good, D. Vanderpool, S. Keeble, K. Bi, Evolution 69,
1961
1972 (2015).
33.
Y. Brandvain, A. M. Kenney, L. Flagel, G. Coop, A. L. Sweigart,
PLOS Genet. 10, e1004410 (2014).
Edelman et al., Science 366, 594599 (2019) 1 November 2019 5of6
RESEARCH | RESEARCH ARTICLE
on November 23, 2020 http://science.sciencemag.org/Downloaded from

Citations
More filters
Posted ContentDOI

Widespread introgression across a phylogeny of 155 Drosophila genomes

TL;DR: This work leverage 155 genome assemblies, from 149 species, to generate a fossil-calibrated phylogeny and conduct multilocus tests for introgression across nine monophyletic radiations within the genus Drosophila, providing the first evidence of introgressive events occurring across the evolutionary history of this genus.
Journal ArticleDOI

Genomics of evolutionary novelty in hybrids and polyploids

TL;DR: This work focuses on the types of molecular genetic and genomic processes that might be set in motion by the saltational event of genome merger between two diverged species, either with or without genome doubling, and how these various processes can contribute to novel phenotypes.
References
More filters
Journal ArticleDOI

New Algorithms and Methods to Estimate Maximum-Likelihood Phylogenies: Assessing the Performance of PhyML 3.0

TL;DR: A new algorithm to search the tree space with user-defined intensity using subtree pruning and regrafting topological moves and a new test to assess the support of the data for internal branches of a phylogeny are introduced.
Journal ArticleDOI

IQ-TREE: A fast and effective stochastic algorithm for estimating maximum likelihood phylogenies

TL;DR: It is shown that a combination of hill-climbing approaches and a stochastic perturbation method can be time-efficiently implemented and found higher likelihoods between 62.2% and 87.1% of the studied alignments, thus efficiently exploring the tree-space.
Journal ArticleDOI

The Human Genome Browser at UCSC

TL;DR: A mature web tool for rapid and reliable display of any requested portion of the genome at any scale, together with several dozen aligned annotation tracks, is provided at http://genome.ucsc.edu.
Journal ArticleDOI

BUSCO: assessing genome assembly and annotation completeness with single-copy orthologs

TL;DR: Zdobnov et al. as discussed by the authors proposed a measure for quantitative assessment of genome assembly and annotation completeness based on evolutionarily informed expectations of gene content, and implemented the assessment procedure in open-source software, with sets of Benchmarking Universal Single-Copy Orthologs.
Journal ArticleDOI

ModelFinder: fast model selection for accurate phylogenetic estimates

TL;DR: ModelFinder is presented, a fast model-selection method that greatly improves the accuracy of phylogenetic estimates by incorporating a model of rate heterogeneity across sites not previously considered in this context and by allowing concurrent searches of model space and tree space.
Related Papers (5)

Butterfly genome reveals promiscuous exchange of mimicry adaptations among species

Kanchon K. Dasmahapatra, +83 more
- 05 Jul 2012 - 
Frequently Asked Questions (12)
Q1. What are the contributions mentioned in the paper "Genomic architecture and introgression shape a butterfly radiation" ?

In this paper, Edelman et al. presented a survey of the work of Nathaniel B. Edelman, Paul B. Frandsen, Michael Miyagi, Bernardo Clavijo, John Davey, Rebecca B. Reed, Kanchon K. Dasmahapatra, Marcus Kronforst, Mathieu Joron, Chris D. Jiggins, W. Owen McMillan, Federica Di Palma, Andrew J. Blumberg, John Wakeley, David Jaffe, James Mallet * 

If linked to weakly deleterious alleles, introgressed loci will finally be purged by linked selection only long after introgression ceases. 

Just as sex aids adaptation within species, occasional introgression and recombination among species can have major long-term effects on the genome, contributing variation that could fuel rapid adaptive divergence and radiation. 

Their test using internal branch lengths in triplet gene trees is based in coalescent theory and takes advantage of the discriminatory power of a property of gene trees not explicitly accounted for by other methods. 

The topology block on chromosome 15 is of particular interest because it spans cortex, a genetic hotspot of wing color pattern diversity in Lepidoptera (25, 26). 

Other studies in Heliconius have shown a high level of phylogenetic discordance, arguably a result of rampant introgression (4, 5). 

Colored bands represent tree topologies of each 50-kb window; colors correspond to the topologies in (B), with black regions showing missing data. 

Tree 1 has a strongly negative correlation with chromosome size (r2 = 0.883, t = 11.7, 18 df, p < 0.0001), whereas Tree 2 (concordant with their inferred species tree) has a positive correlation (r2 = 0.726, t = 6.9, 18 df, p < 0.0001). 

In hybrid populations, individuals have genomic regions that originate from different species and may be incompatible with the recipient genome or with their environment (16). 

The authors attempted to reconstruct a bifurcating species tree by estimating relationships using protein-coding genes, conserved coding regions, and conserved noncoding regions. 

The authors therefore used phyloNet (12) to infer reticulate phylogenetic networks of these species based on random samples of 100 10-kb windows across the alignment. 

(B) Each chromosome was divided into 10 equally sized bins, and the occupancy of each topology in each bin was calculated as the number of windows that recovered the topology in the bin divided by the number of windows that recovered the topology in the chromosome.