ArticlePDF Available

Chance caught on the wing: Cis-regulatory evolution and the origin of pigment patterns in Drosophila.

Authors:

Abstract and Figures

The gain, loss or modification of morphological traits is generally associated with changes in gene regulation during development. However, the molecular bases underlying these evolutionary changes have remained elusive. Here we identify one of the molecular mechanisms that contributes to the evolutionary gain of a male-specific wing pigmentation spot in Drosophila biarmipes, a species closely related to Drosophila melanogaster. We show that the evolution of this spot involved modifications of an ancestral cis-regulatory element of the yellow pigmentation gene. This element has gained multiple binding sites for transcription factors that are deeply conserved components of the regulatory landscape controlling wing development, including the selector protein Engrailed. The evolutionary stability of components of regulatory landscapes, which can be co-opted by chance mutations in cis-regulatory elements, might explain the repeated evolution of similar morphological patterns, such as wing pigmentation patterns in flies.
The spot element evolved through the acquisition of sites for both activators and repressors.a, Schematic of the 675-base-pair spot element showing the boundaries of deletion constructs and the location of identified binding sites. b, d–f, Expression of GFP driven by the spot element (b) and related constructs. c, The anterior border of expression of the selector gene engrailed abuts the spot element expression domain. d, A 196-base-pair element drives the spot pattern but is derepressed in the posterior compartment. e, Deletion of bp 425–453 abolishes activity of the spot element, indicating that sites required for activation lie within this region. f, Disruption of two characterized Engrailed binding sites from the spot element derepresses reporter expression in the posterior compartment. g, The two candidate sites are bound specifically by the Engrailed protein in vitro. Increasing amounts of Engrailed homeodomain–GST fusion protein (0.25–5 nM) specifically shift labelled DNA oligonucleotides representing native sequences containing putative binding sites (left part of each gel) but not sequences in which Engrailed sites have been mutated (underlined in the sequences, right part of each gel). Addition of anti-GST antibody supershifts complexes. Addition of specific (spe.) or non-specific (non-spe.) unlabelled competitor DNA (+ , 50 ng; + + , 500 ng) reveals the specificity of the formation of complexes. Supershift and competition experiments were performed in the presence of 5.0 nM protein.
… 
Content may be subject to copyright.
Chance caught on the wing:
cis-regulatory evolution and the origin
of pigment patterns in Drosophila
Nicolas Gompel*†, Benjamin Prud’homme*, Patricia J. Wittkopp, Victoria A. Kassner & Sean B. Carroll
1
Howard Hughes Medical Institute and Laboratory of Molecular Biology, University of Wisconsin, 1525 Linden Drive, Madison, Wisconsin 53706, USA
* These authors contributed equally to this work
Present addresses: Department of Zoology, Cambridge University, Downing Street, Cambridge CB2 3EJ, UK (N.G.); Department of Molecular Biology and Genetics, 227 Biotechnology Building, Cornell University,
Ithaca, New York 14853, USA (P.J.W.)
...........................................................................................................................................................................................................................
The gain, loss or modification of morphological traits is generally associated with changes in gene regulation during development.
However, the molecular bases underlyi ng these evolutionary changes have remained elusive. Here we identify one of the molecular
mechanisms that contributes to the evolutionary gain of a male-specific wing pigmentation spot in Drosophila biarmipes,a
species closely related to Drosophila melanogaster. We show that the evolution of this spot involved modifications of an ancestral
cis-regulatory element of the yellow pigmentation gene. This element has gained multiple binding sites for transcription factors
that are deeply conserved components of the regulatory landscape controlling wing development, including the selector protein
Engrailed. The evolutionary stability of components of regulatory landscapes, which can be co-opted by chance mutations in
cis-regulatory elements, might explain the repeated evolution of similar morphological patterns, such as wing pigmentation
patterns in flies.
The evolution of new morphological features is due predominantly
to modifications of spatial patterns of gene expression. Changes in
the expression of a particular gene can result from alterations either
in its cis-regulatory sequences or in the deployment and function of
the trans-acting transcription factors that control it, or both.
Understanding the evolution of new morphological traits thus
requires both the identification of genes that control trait formation
and the elucidation of the cis- and trans-modifications that account
for gene expression differences.
Evolution of cis-regulatory elements has been proposed to be a
major source of morphological diversification because mutations in
regulatory elements can produce discrete tissue-specific expression
pattern changes while avoiding deleterious pleiotropic effects
1–3
.In
the best-studied cases of gene expression changes underlying
morphological divergence, cis-regulatory modifications have been
proposed
4–6
, occasionally suggested by genetic evidence
7–10
, but have
only rarely been formally demonstrated
11
or analysed at the mol-
ecular level
12,13
. It is currently not known whether the evolution of
new morphological traits occurs largely through the modification of
pre-existing cis-regulatory elements or from the generation of new
elements; neither is it understood how many or what kinds of
modifications are required for a regulatory element to drive a novel
pattern.
To address these issues, we have analysed the evolution of a
conspicuous male-specific wing pigmentation pattern in Drosophila
biarmipes, a species closely related to Drosophila melanogaster
14
(Fig. 1). Wing pigmentation patterns in insects are highly diversified
and have various biological functions including mimicry, camou-
flage, thermoregulation, and mate selection
15
.InD. biarm ipes, the
sexually dimorphic wing pattern is associated with a courtship
behaviour in which males display their wings conspicuously to the
females, suggesting a function for this spot in mate choice
16,17
. This
wing spot has evolved recently in some species of the D. melanoga-
ster group, such as D. biarmipes, and it is absent from close outgroup
species such as D. pseudoobscura
17
(Fig. 1).
Formation of wing pigmentation results from the conversion of
melanin precursors diffusing from the veins into pigment deposits
at specific positions along the wing, wherever converting proteins
are present
18
. The product of the yellow (y) gene is required for the
production of black pigments, and the distribution of its product
prefigures adult pigmentation patterns
11,19
. The Yellow protein is
expressed uniformly at low levels throughout the developing wings
of D. melanogaster and D. pseudoobscura, where it imparts a low
overall level of melanic pigmentation. In contrast, in D. biarmipes,
in addition to the low, uniform expression, Yellow protein is highly
expressed in an anterior distal spot
19
(Fig. 1). This tight correlation
between a novel Yellow expression pattern and a novel pigmentation
Figure 1 Expression of the Yellow protein prefigures adult wing pigmentation. The
conspicuous spot of dark pigmentation present at the tip of the male wing of Drosophila
biarmipes (left) is a new trait evolved among species of the Drosophila melanogaster
group
14,46
(about 15 Myr of divergence; divergence time is 60–80 Myr for the family
Drosophilidae
30
), superimposed on the ancestral pattern of uniform grey shading and
darker veins found both in D. melanogaster and in D. pseudoobscura, a species from the
sister D. obscura group (25 Myr of divergence
29,30
). In all three species the male pupal
distribution of Yellow in the wing, revealed by a specific antibody (right), foreshadows the
adult pigmentation.
articles
NATURE | VOL 433 | 3 FEBRUARY 2005 | www.nature.com/nature 481
© 2005 Nature Publishing Group
pattern prompted us to ask whether regulatory evolution at the y
locus underlies the novel distribution of the Yellow protein in
D. biarmipes, or whether this is due to changes in trans-acting
regulators of y.
Regulatory changes in cis to the yellow locus
To test whether the observed differences in Yellow expression
between D. bi armipes and D. melanogaster (Fig. 1) are due to
changes at the y locus, we transformed D. melanogaster with green
fluorescent protein (GFP)-reporter constructs containing non-
coding DNA from the D. biarmipes y (y
bia
) locus. If relevant
evolutionary changes have occurred in cis, then the reporter gene
might be regulated in D. melanogaster in a manner similar to the
native y gene in D. biarmipes. If, however, the changes have occurred
in trans, the D. biarmipes y regulatory element might drive reporter
expression similar to that of the y gene in D. melanogaster (that is,
uniformly). We found that D. melanogaster transgenic flies carrying
the entire 5
0
region (8 kilobases; Fig. 2a) of the y
bia
gene (5
0
y
bia
)
express GFP in the pupal wings in a pattern similar to the native
D. biarmipes Yellow expression (Fig. 2b). Low levels of GFP are
uniformly distributed across the wing, and higher levels of GFP are
confined to the distal part of the anterior compartment. This result
shows that the transcription factors deployed in the developing
wing of D. melanogaster recognize y
bia
cis-regulatory sequences.
Furthermore, the D. biarmipes-like expression pattern in a
D. melanogaster trans-regulatory context shows that evolutionary
changes in Yellow expression involve primarily cis-regulatory modi-
fications at the y locus, which presumably entail the gain (or loss) of
binding sites for transcription factors.
The 5
0
y
bia
element does not recapitulate the precise restriction of
the native spot of Yellow expression; higher levels of reporter protein
expression extend along the proximal–distal axis, indicating that
additional regulatory differences exist between D. biarmipes and
D. melanogaster. Additional reporter constructs suggest that these
differences are trans effects or are due to cis-acting elements located
outside the region we have tested. The unique intron of the
D. biarmipes y gene does contain another cis-regulatory element
(for all developing sensory bristles) but has no activity in the wing
other than in these sense organs. Furthermore, a transgene contain-
ing the 5
0
non-coding, 5
0
untranslated region, first exon, intron and
second exon sequences (partial locus, Fig. 2a) is expressed in a
similar pattern to that of the 5
0
y
bia
element, indicating that the
differences in y expression are not due to any of these sequences
(data not shown).
Having localized major regulatory differences to the y
bia
5
0
region,wenextinvestigatedwhetherthenovelcis-regulatory
activity of the y
bia
region arose in a pre-existing regulatory element
or evolved de novo in the D. biarmipes lineage.
Figure 2 Cis-regulatory changes at the yellow locus are responsible for species-specific
differences in Yellow distribution. a, The organization of the y locus is similar in Drosophila
melanogaster, Drosophila biarmipes and D. pseudoobscura. b, The entire 5
0
region of
D. biarmipes y, comprising sequences between the coding sequences of y and the closest
predicted gene (CG3777), is sufficient to drive reporter GFP expression in D. melanogaster
at a time and in a pattern similar to those of y expression in native D. biarmipes. The y
bia
intron does not drive wing expression other than in the marginal sensory bristles, and the
partial locus drives expression in a pattern similar to the entire 5
0
region of y
bia
(not
shown). Black boxes, coding sequence; grey boxes, fragments analysed in transgenic
constructs.
articles
NATURE | VOL 433 | 3 FEBRUARY 2005 | www.nature.com/nature482
© 2005 Nature Publishing Group
Evolution of a wing-specific cis-regulatory element
In D. melanogaster, analysis of the y regulatory region has revealed
that an 800-base-pair (bp) element located 1 kilobase upstream of
the transcription start site, named the w ing
mel
element (Fig. 2a), is
sufficient to drive gene expression throughout the pupal wing
(Fig. 3b) and is necessary for adult wing pigmentation
11,20,21
.We
hypothesized that functional modifications of this element might
account for differences in Yellow expression between the wings of
D. melanogaster and D. biarmipes. There is strong sequence con-
servation of this portion of the y locus between the two
species (Fig. 3a and Supplementary Fig. 1). We transformed
D. melanogaster with a GFP-reporter construct containing a 920-
bp fragment from D. biarmipes orthologous to the D. melanogaster
wing element (wing
mel
), termed wing
bia
(Fig. 2a). This fragment
drives a reporter pattern resembling that driven by the 5
0
y
bia
element, with slightly less contrast between the levels of overall
expression in the wing and in the anterior distal area (not shown). A
larger fragment encompassing wing
bia
,namedwing
bia large
(1,542 bp; Fig. 2a) drives a reporter pattern identical to the 5
0
y
bia
element (Fig. 3b). These results indicate that the sequences required
for the strong anterior-distal activation of Yellow expression in
D. biarmipes pupal wings are located within and immediately
adjacent to a wing-specific cis-regulatory element that is ortholo-
gous to the wing
mel
element.
To determine whether the novel wing
bia
sequences evolved within
an ancestral w ing cis-regulatory element, we examined D. pseudo-
obscura, an outgroup species that belongs to a clade generally devoid
of wing pigmentation patterns other than the grey (light black)
homogeneous shading (Fig. 1). Phylogenetic character reconstruc-
tion suggests that the pigmentation spot was present in the common
ancestor of D. biarmipes and D. melanogaster and has been lost in the
D. melanogaster lineage
22,23
. There is substantial sequence conserva-
tion at the y gene between D. biarmipes and D. pseudoobscura (Fig. 3a
and Supplementary Fig. 1), which allowed us to identify a region in
D. pseudoobscura that is orthologous to the wing
bia
element, named
wing
pse
(724 bp). This wing
pse
element drives ubiquitous wing
expression (Fig. 3b), demonstrating that a functional wing element
is ancestral to the D. melanogaster/D. biarmipes lineage and that
sequences within and/or adjacent to this element were modified to
control high levels of expression in the anterior distal part of the
wing in D. biarmipes .
To understand the organization of the wing
bia
element and to
localize its novel functional sequences, we further dissected the
wing
bia
element. We found that the sequences necessary for the
anterior distal expression are separable from those controlling the
general wing expression in D. biarmipes. Two complementary, non-
overlapping sequences of the wing
bia
element, right
bia
and left
bia
(Fig. 2a), drive respectively ubiquitous expression throughout the
wing blade and strong activation in the anterior distal area of the
wing (Fig. 3b) (as do the complementary subfragments from
the wing
bia large
element, right
bia large
(not shown) and left
bia large
;
Fig. 4b). A similar dissection clearly separates two wing-specific
complementary functions in D. pseudoobscura (ubiquitous
expression, and expression around the veins) but yields non-
functional elements in D. melanogaster (Fig. 3b). These results
indicate that sites in both regions of the w ing element are required
for its function in D. melanogaster and D. pseudoobscura, and that
some or all of the novel sequences in D. biarmipes responsible for the
specific anterior distal wing expression of Yellow are located in
the left
bia large
element, hereafter referred to as the spot element. The
distinct and robust activities of the two parts of the wing
bia
element
raise the possibility that the wing element has been subfunctiona-
lized in D. biarmipes into two elements controlling expression
throughout the wing and in the spot, respectively.
Figure 3 The cis-regulatory sequences governing spot formation evolved in the context of
an ancestral wing enhancer. a, Conservation of the wing element sequence between
D. biarmipes (bia) and D. melanogaster (mel)orD. pseudoobscura (pse) determined by
Vista
47
with a 10-base-pair window length; only conservation above 75% is shown as
solid boxes. Arrows show the boundaries of the left and right fragments. b, Reporter
expression driven by the orthologous wing elements and its subfragments left and right
(columns) of D. melanogaster (top), D. biarmipes (middle; the wing
bia large
element is
shown) and D. pseudoobscura (bottom), all expressed in D. melanogaster. The ubiquitous
expression driven by the outgroup species wing
pse
element (expression is present in vein
cells at a lower levels comparable to those in left
pse
) shows that the sequences
responsible for the spot pattern in D. biarmipes have evolved in the context of an ancestral
wing regulatory element. The sequences controlling the spot pattern are separable from
those controlling general expression in D. biarmipes (left and right). Note that the posterior
boundary of activity of the left
bia
construct lies near or at the anterior–posterior
compartment boundary.
articles
NATURE | VOL 433 | 3 FEBRUARY 2005 | www.nature.com/nature 483
© 2005 Nature Publishing Group
Multiple sites evolved in the wing spo t element
In principle, the evolution of the spot pattern could arise through
gaining binding sites for a single transcription factor that is
expressed precisely in the cells that form the spot pattern. Alter-
natively, the spot pattern could result from the evolution of a
combination of binding sites for multiple activators, as well as
potential repressors that might restrict expression to this area.
Resolution of these possibilities bears on the general question of
the number of steps involved in the evolution of new patterns of
gene expression and cis-regulatory element function.
To distinguish between these possibilities, we derived a series of
reporter gene constructs with smaller portions of the 675-bp spot
element. A 196-bp construct (335–530; Fig. 4a) retained activity in
the anterior distal region of the wing, although we noted that
reporter expression now extended into the posterior compartment
(Fig. 4d). This suggested that one or more sites critical for activation
Figure 4 The spot element evolved through the acquisition of sites for both activators and
repressors. a, Schematic of the 675-base-pair spot element showing the boundaries of
deletion constructs and the location of identified binding sites. b, df, Expression of GFP
driven by the spot element (b) and related constructs. c, The anterior border of expression
of the selector gene engrailed abuts the spot element expression domain. d, A 196-base-
pair element drives the spot pattern but is derepressed in the posterior compartment.
e, Deletion of bp 425–453 abolishes activity of the spot element, indicating that sites
required for activation lie within this region. f, Disruption of two characterized Engrailed
binding sites from the spot element derepresses reporter expression in the posterior
compartment. g, The two candidate sites are bound specifically by the Engrailed protein
in vitro. Increasing amounts of Engrailed homeodomain–GST fusion protein (0.25–5 nM)
specifically shift labelled DNA oligonucleotides representing native sequences containing
putative binding sites (left part of each gel) but not sequences in which Engrailed sites
have been mutated (underlined in the sequences, right part of each gel). Addition of anti-
GST antibody supershifts complexes. Addition of specific (spe.) or non-specific (non-spe.)
unlabelled competitor DNA (þ, 50 ng; þþ, 500 ng) reveals the specificity of the
formation of complexes. Supershift and competition experiments were performed in the
presence of 5.0 nM protein.
articles
NATURE | VOL 433 | 3 FEBRUARY 2005 | www.nature.com/nature484
© 2005 Nature Publishing Group
resided in this 196-bp element and that one or more sites necessary
for the restriction of expression from the posterior compartment
resided outside it.
To localize further sequences required for activation of the spot
element, we constructed a series of small deletions spanning the
length of the 196-bp element (Fig. 4a, e, and data not shown). We
found that a fragment lacking the internal sequences from bp 425 to
453 completely lacked reporter expression (Fig. 4e). This indicates
that sequences required for activation in the spot are located within
or overlap with bp 425–453. Together, these results indicate that
sites for both at least one activator and one repressor have evolved in
the spot element.
Direct regulation of the spot element by Engrail ed
We next sought to identify potential trans-acting factors that
regulate the spot element in D. biarmipes. The conspicuous posterior
boundary of gene expression observed with the left
bia
and spot
elements (Figs 3b and 4b) is reminiscent of the compartment
boundary of the wing
24
defined by the anterior border of expression
of the selector transcription factor Engrailed
25
(Fig.4c).The
posterior expansion of GFP expression in the deletion constructs
shown in Fig. 4d would be consistent with posterior repression of
the spot element by Engrailed.
To test whether Engrailed might be a direct regulator of the
wing
bia
element, we searched the spot sequence for putative
Engrailed binding sites
26
and identified several candidate sites.
Two of these sites, clustered within 43 bp, are specific to the
D. biarmipes s pot element (absent from D. melanoga ster and
D. pseudoobscura elements; Supplementary Fig. 1) and one site is
located outside the 196-bp construct that exhibits some reporter
expression in the posterior compartment (Fig. 4d). Gel-shift experi-
ments on the native and mutated versions of these two sites showed
that Engrailed binds specifically to them in vitro (Fig. 4g). Disrup-
tions of these two Engrailed binding sites in the context of the spot
element result in the specific derepression of reporter gene
expression in the posterior compartment (Fig. 4f). These results
show that the selector protein Engrailed directly represses the
expression of the y gene in the posterior compartment of the wing
and is one of the inputs that shapes the contours of the wing spot in
D. biarmipes.
Multistep and multigenic evolution of the spot
Although multiple cis-regulatory modifications at the y locus have
produced a profound evolutionary change in Yellow protein
expression, it is important to ascertain whether changes in this
one gene are sufficient for the evolution of the physical trait or
whether additional evolutionary events are required. We have found
that changes at y alone are not sufficient to create a pigmentation
spot.
D. mela nogaster y mutants carrying the D. biarmipes y gene
(Fig. 2a, partial locus) recover only their species-specific pigment
patterns; no wing spot is generated (not shown). Additional loci
must therefore be involved.
The formation of pigment patterns is a multigenic process, and
evolution at other pigmentation loci could also contribute to
pattern evolution
18,27,28
.ThemalespotofD. biarmipes is also
associated with the localized downregulation of the melanin-
inhibiting product of the ebony (e) gene during wing development
19
(Fig. 5c), in a pattern that is approximately the inverse of Yellow
expression. This suggests that, at least, both the repression of e and
the activation of y are necessary for the formation of a dark spot.
Consistent with this hypothesis is the observation that in
D. melanogaster e mutants carrying the y
bia
partial locus transgene
(Fig. 2a), a slight darkening is observed specifically in the anterior
area of the wing where yellow is strongly expressed (data not
shown). However, this darkening is not comparable to the intense
pigmentation spot of D. biarmipes. Changes in the expression of
other pigmentation genes must also be involved. Furthermore, we
have not been able to test whether changes in the trans-acting
regulatory network of D. biarmipes might also contribute to the
unique patterns of gene expression in the area of the wing spot.
Taken together, these results indicate that the evolution of the novel
pigmentation pattern of D. biarmipes required changes at multiple
loci.
To determine whether the inverse regulation of expression of y
and e is a general mechanism for the evolution of novel wing
pigmentation patterns, we examined the expression of these pro-
teins in D. guttifera, a species that separated from the D. melano-
gaster lineage about 40 Myr ago
29
. This species has independently
evolved a strikingly different and more complex wing pigmentation
pattern (Fig. 5a). We found that the pattern of expression of the two
proteins also exhibits an inverse relationship with higher levels of
Yellow (Fig. 5b) and lower levels of Ebony (Fig. 5d) in the pupal
wing where the eventual adult pigmentation spots will form. This
indicates that the evolution of both y and e expression is involved in
the formation and evolution of novel wing pigmentation patterns in
drosophilids.
Chance caught on the wing: novelty by co-option
In drosophilid flies, the shape of the wings and the pattern of
Figure 5 Concerted changes in the expression of Yellow and Ebony underlie the evolution
of novel wing patterns. a, The distant species D. guttifera (a member of the D. quinaria
group
48
) has evolved a complex pattern of dark spots located at the intersection of wing
veins and where campaniform sensilla form. The grey shading is also reinforced in some
interveins. b, The pupal distribution of Yellow also prefigures this adult pattern. c,In
Drosophila biarmipes, the spatial repression of Ebony is also associated with the formation
of the adult male spot of pigmentation
19
. d, This repression of Ebony associated with
pigmentation patterns seems to be general, because it is also seen in D. guttifera, where
the adult spots will form.
articles
NATURE | VOL 433 | 3 FEBRUARY 2005 | www.nature.com/nature 485
© 2005 Nature Publishing Group
venation have not changed much over 60–80 Myr of evolution
30,31
.
Their development and patterning are largely understood in
D. melanogaster and the regulatory proteins involved are con-
served
32
. One such protein, the selector protein Engrailed, is a
deeply conserved feature of the compartmental organization of
arthropod segments and appendages. In the Drosophila wing,
Engrailed is part of the regulatory circuit that sequentially organizes
the patterning of the anterior–posterior axis
33
. Here we have shown
that the activity of this transcription factor has been co-opted to
control a feature of the novel wing pigmentation pattern in
D. biarmipes through the evolution of specific binding sites within,
or in the immediate vicinity of, a wing-specific regulatory element
of the y gene. Because the expression driven by the spot element is
also spatially modulated in D. melanogaster, this indicates that other
conserved components of the wing trans-regulatory landscape (that
is, one or more activators) have similarly been co-opted by the
evolution of binding sites within the y wing element.
These findings suggest a general means by which novel expression
patterns and characters can arise (Fig. 6a). Specifically, the random
mutation of ancestral cis-regulatory elements (including point and
insertional mutations) generates potential binding sites. If and
when these sites can be recognized by transcription factors
expressed in cells in which the ancestral element is active, the
pattern or level of gene expression may be modified (Fig. 6a), in a
manner similar to the mechanism of gene co-option demonstrated
by the vertebrate crystallin genes
34
. The patterns of expression of the
eligible transcription factors are initially cryptic with respect to the
target gene or trait, but these cryptic ‘prepatterns’ are revealed once
functional binding sites have evolved in target genes. In this sense,
and in this example, evolution is precisely a matter, as Jacques
Monod put it, of ‘chance caught on the wing’
35,36
.
This model has two specific implications for the evolution of
novel wing patterns. First, it explains how the observed diversity of
wing pigmentation patterns might result from combinations of the
numerous transcription factors expressed in the wing. Each of these
combinations might constitute a distinct prepattern for pigmenta-
tion genes such as y or e, provided that the corresponding binding
sites evolve in the proper cis-regulatory context. For instance, some
of the spots on the wing of D. guttifera surround the sensory organs
located on the veins, which form at similar positions in most
drosophilids. This raises the possibility that transcription factors
involved in the positioning of these landmark organs have been co-
opted to change y or e regulation in D. guttifera. Second, this model
might explain the widespread repeated evolution of strikingly
similar pigmentation patterns observed in distantly related species
(for instance, pigmentation patterns similar to those studied here
have evolved independently in other dipterans; Fig. 6b). The
evolutionary stability of the trans-regulatory landscape in droso-
philid wings, reflected by the strong conservation of the wing shape
and venation pattern in the family, suggests that similar pigmenta-
tion patterns might arise in parallel through the repeated evolution
of binding sites for the same transcription factors in cis-regulatory
regions of pigmentation genes. A
Methods
Fly stock and maintenance
Flies were bred at 25 8C on Wheeler–Clayton
37
or cornmeal
38
medium. Constructs were
transformed into D. melanogaster yw mutants as described previously
39,40
. The CantonS
strain was used as wild-type D. melanogaster. Drosophila pseudoobscura, Drosophila
biarmipes and Drosophila guttifera stocks were obtained from the Tucson stock centre
(stock numbers 14011-0121.94, 14023-0361.01 and 15130-1971.10, respectively). All
mature D. biar mipes males of this stock exhibited the wing spot. The en-Gal4 and
UASGFP stocks were obtained from the Bloomington Drosophila stock centre.
Immunochemistry
Pupal wings (70 h after puparium formation), still attached to the fly, were allowed to
unfold in water after removal of the pupal cuticle. Flies were transferred to phosphate-
buffered saline (PBS), in which the wings were cut off with a razor blade. Wings were fixed
flat for 15 min between a slide and a coverslip in 4% formaldehyde PBT (PBS containing
0.03% Triton X-100), transferred on ice to a scintillation vial in the fixing solution for a
further 15 min, sonicated briefly in the fixative with a Branson 200 ultrasonic cleaner, fixed
for a further 30 min, washed with PBT, blocked for 1 h in PBT containing 1% bovine serum
albumin, stained with a rat anti-yellow or a rabbit anti-ebony primary antibody
19
and
revealed respectively with a fluorescein isothiocyanate (FITC)-conjugated anti-rat
antibody or FITC-conjugated anti-rabbit IgG antibody (Jackson Immunoresearch).
Cloning
The D. biarmipes y locus sequence was amplified by direct and inverse polymerase chain
reaction (PCR; details are available from the authors on request). The entire 5
0
region was
amplified by PCR with primers designed in the coding sequences of y and the closest gene
upstream of y in D. melanogaster (CG3777; ref. 41). All y fragments for reporter constructs
were cloned into a customized version of the P-based transformation vector
42
from which
one of the two gypsy insulators had been removed and a new polylinker had been added.
Fragments from D. melanogaster and D. pseudoobscura were amplified by PCR from
genomic DNA and specific primers designed using available genome sequences
43,44
(see
Supplementary Table 1 for primer sequences).
Biochemistry
The D. melanogaster Engrailed homeodomain sequence was cloned into the glutathione
S-transferase (GST) gene fusion vector pGEX-3X (Amersham Bioscience). The GST
fusion protein was purified by affinity chromatography
45
. DNA probes for electrophoretic
mobility-shift assays were double-stranded oligonucleotides labelled with
32
P by end-
filling in at both ends with the Klenow fragment of DNA polymerase I. Single-stranded
oligonucleotides were annealed at a final concentration of 0.1
m
M in 10 mM Tris-HCl
pH 7.5 containing 0.1 M NaCl and 1 mM EDTA. Sequences of the oligonucleotide pairs
were as follows: native sequences, 5
0
-TTTCCGCCTAATTGATG-3
0
and 5
0
-TTTCATCAAT
TAGGCGG-3
0
,5
0
-TTTTGCCAATCATTTTT-3
0
and 5
0
-TTTAAAAATGATTGGCA-3
0
;
mutated versions, 5
0
-TTTCCGCCTcccTGATG-3
0
and 5
0
-TTTCATCAgggAGGCGG-3
0
,
TTTTGCCgggCATTTTT-3
0
and 5
0
-TTTAAAAATGcccGGCA-3
0
. Labelled probes were
purified with G50 Sephadex beads (Sigma) on chromatography columns (Bio-Rad).
DNA-binding assays, competition experiments and gel migrations were performed with
10–15 fmol of labelled probes (about 10
4
c.p.m.) following a published protocol
26
;they
were pre-run for 0.5 h and run for 1.5 h at 4 8C on 8% native polyacrylamide minigels in
0.5 £ Tris/borate/EDTA buffer pH 8.3. Non-specific competitor consisted of herring
sperm DNA (Sigma) and the specific competitor was as used elsewhere
26
.
Figure 6 Cryptic prepatterns and the evolution of novel gene expression patterns through
the evolution of cis-regulatory sequences. a, The upper panel shows a model of the
conserved landscape of transcriptional regulators that pattern and shape the Drosophila
wing (green and pink represent repressor and activator, respectively). The evolution of
binding sites for a subset of these regulators in the yellow wing cis-regulatory element
(coloured stars) co-opts them to modify yellow expression (lower panel). Combined with
other regulatory changes at other loci, the changes at the y locus result in a novel
pigmentation spot. b, Wing pigmentation patterns similar to D. biarmipes (left) or
D. guttifera (right) evolved independently in other fly families (here Otitidae and
Lauxaniidae).
articles
NATURE | VOL 433 | 3 FEBRUARY 2005 | www.nature.com/nature486
© 2005 Nature Publishing Group
Wing imaging
Adult wings were mounted flat in Hoyer’s medium
38
and processed for bright-field
imaging with a 4 £ or 10 £ dry lens on a Zeiss Axiophot microscope equipped with a
Kontron charge-coupled device camera. For all reporter lines, pupal wings 70–90 h after
puparium formation were mounted flat between a slide and a coverslip in PBT, without
fixation, and imaged immediately with an Optiphot confocal microscope (Nikon)
equipped with a 4 £ dry lens and a BioRad 1024 system. Antibody-stained preparations
were mounted in glycerol and imaged.
Received 30 September; accepted 1 December 2004; doi:10.1038/nature03235.
1. Carroll, S. B., Grenier, J. K. & Weatherbee, S. D. From DNA to Diversity: Molecular Genetics and the
Evolution of Animal Design 2nd edn (Blackwell Science, Malden, Massachusetts, 2004).
2. Davidson, E. H. Genomic Regulatory Systems: Development and Evolution (Academic, San Diego, 2001).
3. Stern, D. L. Evolutionary developmental biology and the problem of variation. Evolution 54,
1079–1091 (2000).
4. Averof, M. & Patel, N. H. Crustacean appendage evolution associated with changes in Hox gene
expression. Nature 388, 682–686 (1997).
5. Gompel, N. & Carroll, S. B. Genetic mechanisms and constraints governing the evolution of correlated
traits in drosophilid flies. Nature 424, 931–935 (2003).
6. Yoon, H. S. & Baum, D. A. Transgenic study of parallelism in plant morphological evolution. Proc.
Natl Acad. Sci. USA 101, 6524–6529 (2004).
7. Sucena, E. & Stern, D. L. Divergence of larval morphology between Drosophila sechellia and its sibling
species caused by cis-regulatory evolution of ovo/shaven-baby. Proc. Natl Acad. Sci. USA 97,
4530–4534 (2000).
8. Shapiro, M. D. et al. Genetic and developmental basis of evolutionary pelvic reduction in threespine
sticklebacks. Nature 428, 717–723 (2004).
9. Stern, D. L. A role of Ultrabithorax in morphological differences between Drosophila species. Nature
396, 463–466 (1998).
10. Wang, R. L., Stec, A., Hey, J., Lukens, L. & Doebley, J. The limits of selection during maize
domestication. Nature 398, 236–239 (1999).
11. Wittkopp, P. J., Vaccaro, K. & Carroll, S. B. Evolution of yellow gene regulation and pigmentation in
Drosophila. Curr. Biol. 12, 1547–1556 (2002).
12. Wang, X. & Chamberlin, H. M. Multiple regulatory changes contribute to the evolution of the
Caenorhabditis lin-48 ovo gene. Genes Dev. 16, 2345–2349 (2002).
13. Belting, H. G., Shashikant, C. S. & Ruddle, F. H. Modification of expression and cis-regulation of
Hoxc8 in the evolution of diverged axial morphology. Proc. Natl Acad. Sci. USA 95, 2355–2360 (1998).
14. Bock, I. R. & Wheeler, M. R. in Studies in Genetics (ed. Wheeler, M. R.) 1–102 (Univ. of Texas, Austin,
1972).
15. Majerus, M. E. N. Melanism: Evolution in Action (Oxford Univ. Press, Oxford, 1998).
16. Singh, B. N. & Chatterjee, S. Greater mating success of Drosophila biarmipes males possessing an apical
dark black wing patch. Ethology 75, 81–83 (1987).
17. Kopp, A. & True, J. R. Evolution of male sexual characters in the oriental Drosophila melanogaster
species group. Evol. Dev. 4, 278–291 (2002).
18. True, J. R., Edwards, K. A., Yamamoto, D. & Carroll, S. B. Drosophila wing melanin patterns form by
vein-dependent elaboration of enzymatic prepatterns. Curr. Biol. 9, 1382–1391 (1999).
19. Wittkopp, P. J., True, J. R. & Carroll, S. B. Reciprocal functions of the Drosophila Yellow and Ebony
proteins in the development and evolution of pigment patterns. Development 129, 1849–1858 (2002).
20. Chia, W. et al. Molecular analysis of the yellow locus of Drosophila. EMBO J. 5, 3597–3605 (1986).
21. Geyer, P. K. & Corces, V. G. Separate regulatory elements are responsible for the complex pattern of
tissue-specific and developmental transcription of the yellow locus in Drosophila melanogaster. Genes
Dev. 1, 996–1004 (1987).
22. Kopp, A. & True, J. R. Phylogeny of the oriental Drosophila melanogaster species group: a multilocus
reconstruction. Syst. Biol. 51, 786–805 (2002).
23. Schawaroch, V. Phylogeny of a paradigm lineage: the Drosophila melanogaster species group (Diptera:
Drosophilidae). Biol. J. Linn. Soc. 76, 21–37 (2002).
24. Garcia-Bellido, A., Ripoll, P. & Morata, G. Developmental compartmentalization of the wing disk of
Drosophila. Nat. New Biol. 245, 251–253 (1973).
25. Blair, S. S. Engrailed expression in the anterior lineage compartment of the developing wing blade of
Drosophila. Development 115, 21–33 (1992).
26. Solano, P. J. et al. Genome-wide identification of in vivo Drosophila Engrailed-binding DNAfragments
and related target genes. Development 130, 1243–1254 (2003).
27. Wittkopp, P. J., Williams, B. L., Selegue, J. E. & Carroll, S. B. Drosophila pigmentation evolution:
Divergent genotypes underlying convergent phenotypes. Proc. Natl Acad. Sci. USA 100, 1808–1813
(2003).
28. Llopart, A., Elwyn, S., Lachaise, D. & Coyne, J. A. Genetics of a difference in pigmentation between
Drosophila yakuba and Drosophila santomea. Evolution 56, 2262–2277 (2002).
29. Russo, C. A., Takezaki, N. & Nei, M. Molecular phylogeny and divergence times of drosophilid species.
Mol. Biol. Evol. 12, 391–404 (1995).
30. Powell, J. R. Progress and Prospects in Evolutionary Biology: the Drosophila Model (Oxford Univ. Press,
New York, 1997).
31. Hardy, D. E. Diptera: Cyclorrhapha II, Series Schizophora, Section Acalypterae I, Family Drosophilidae
(Univ. of Hawaii Press, Honolulu, 1965).
32. De Celis, J. F. Pattern formation in the Drosophila wing: The development of the veins. BioEssays 25,
443–451 (2003).
33. Blair, S. S. Compartments and appendage development in Drosophila. BioEssays 17, 299–309 (1995).
34. Gonzalez, P., Rao, P. V., Nunez, S. B. & Zigler, J. S. Jr Evidence for independent recruitment of zeta-
crystallin/quinone reductase (CRYZ) as a crystallin in camelids and hystricomorph rodents. Mol. Biol.
Evol. 12, 773–781 (1995).
35. Monod, J. Le Hasard et la Ne
´
cessite
´
. Essai sur la Philosophie Naturelle de la Biologie Moderne (E
´
ditions
du Seuil, Paris, 1970).
36. Kauffman, S. A. At Home in the Universe: the Search for the Laws of Self-organization and Complexity
(Oxford Univ. Press, New York, 1995).
37. Wheeler, M. R. & Clayton, F. E. A new Drosophila culture technique. Drosophila Inf. Serv. 40, 98
(1965).
38. Ashburner, M. Drosophila. A Laboratory Manual (Cold Spring Harbor Laboratory Press, Cold Spring
Harbor, New York, 1989).
39. Spradling, A. C. & Rubin, G. M. Transposition of cloned P elements into Drosophila germ line
chromosomes. Science 218, 341–347 (1982).
40. Miller, D. F., Holtzman, S. L. & Kaufman, T. C. Customized microinjection glass capillary needles for
P-element transformations in Drosophila melanogaster. Biotechniques 33, 366–375 (2002).
41. Flybase. http://flybase.bio.indiana.edu.
42. Barolo, S., Carver, L. A. & Posakony, J. W. GFP and
b
-galactosidase transformation vectors for
promoter/enhancer analysis in Drosophila. Biotechniques 29, 726–732 (2000).
43. Celniker, S. E. et al. Finishing a whole-genome shotgun: release 3 of the Drosophila melanogaster
euchromatic genome sequence. Genome Biol. 3, RESEARCH0079 (2002).
44. Human Genome Sequencing Center. Drosophila genome project. http://www.hgsc.bcm.tmc.edu/
projects/drosophila/ (2002).
45. Schendel, P. F. in Current Protocols in Molecular Biology (eds Ausubel, F. M. et al.) 16.7.1–16.7.7 (Wiley,
New York, 1993).
46. Bock, I. R. Current status of the Drosophila melanogaster species-group (Diptera). Syst. Entomol. 5,
341–356 (1980).
47. Mayor, C. et al. VISTA: Visualizing global DNA sequence alignments of arbitrary length.
Bioinformatics 16, 1046–1047 (2000).
48. Remsen, J. & O’Grady, P. Phylogeny of Drosophilinae (Diptera: Drosophilidae), with comments on
combined analysis and character support. Mol. Phylogenet. Evol. 24, 249–264 (2002).
Supplementary Information accompanies the paper on www.nature.com/nature.
Acknowledgements We thank J. True, C. E. Nelson, C. M. Walsh and C. T. Hittinger for technical
advice; J. True, S. Blair and members of the Carroll laboratory for discussions; B. L. Williams and
J. Yoder for critical comments on the manuscript; S. Castrezana and T. Markow (Tucson
Drosophila Stock Center) for providing Drosophila stocks; J. P. Gruber for the Euxesta sample; and
S. Barolo for the pH Stinger vector. N.G. was funded by an EMBO long-term postdoctoral
fellowship; B.P. and N.G. are recipients of a Philippe Foundation fellowship. The project was
supported by the Howard Hughes Medical Institute (S.B.C.).
Competing interests statement The authors declare that they have no competing financial
interests.
Correspondence and requests for materials should be addressed to S.B.C. (sbcarrol@wisc.edu).
The D. biarmipes y locus sequence is deposited in GenBank under accession number AY817623.
articles
NATURE | VOL 433 | 3 FEBRUARY 2005 | www.nature.com/nature 487
© 2005 Nature Publishing Group
... For nearly 20 y (3,4), studies of the genetics of adaptation have focused on the relative evolutionary contributions of mutations changing protein-coding sequences directly and mutations influencing the expression of proteins by changing the underlying regulatory machinery. Both types of genetic variation contribute to adaptation (5,6), but changes to gene-expression patterns have, in general, been found to explain a disproportionately higher amount of adaptive phenotypic variation (7)(8)(9)(10). As our understanding of gene regulation has matured and technological advancements have improved the resolution with which we can characterize the organization and function of gene regulatory networks (11)(12)(13), focus is shifting from determining whether regulatory evolution is important to deciphering the genetic mechanisms, mutational processes, and adaptive pathways that make regulatory evolution so prevalent. ...
... We compared our chromosome recovery and identity with the male Crotalus viridis genome (23) and found strong overall agreement across all autosomes and the Z chromosome, with no loci matching to our W chromosome (Fig. 1B), as expected. We recovered both telomeres for chromosomes 2, 3, 5, 9, 10, 13, 16, and Z; we also recovered one telomere for chromosomes 6,7,11,12,14,15, and 17. Putative centromeres were inferred for all chromosomes as the longest nontelomeric, high-order repeat regions (SI Appendix, Figs. ...
Article
Developmental phenotypic changes can evolve under selection imposed by age- and size-related ecological differences. Many of these changes occur through programmed alterations to gene expression patterns, but the molecular mechanisms and gene-regulatory networks underlying these adaptive changes remain poorly understood. Many venomous snakes, including the eastern diamondback rattlesnake ( Crotalus adamanteus ), undergo correlated changes in diet and venom expression as snakes grow larger with age, providing models for identifying mechanisms of timed expression changes that underlie adaptive life history traits. By combining a highly contiguous, chromosome-level genome assembly with measures of expression, chromatin accessibility, and histone modifications, we identified cis-regulatory elements and trans-regulatory factors controlling venom ontogeny in the venom glands of C. adamanteus . Ontogenetic expression changes were significantly correlated with epigenomic changes within genes, immediately adjacent to genes (e.g., promoters), and more distant from genes (e.g., enhancers). We identified 37 candidate transcription factors (TFs), with the vast majority being up-regulated in adults. The ontogenetic change is largely driven by an increase in the expression of TFs associated with growth signaling, transcriptional activation, and circadian rhythm/biological timing systems in adults with corresponding epigenomic changes near the differentially expressed venom genes. However, both expression activation and repression contributed to the composition of both adult and juvenile venoms, demonstrating the complexity and potential evolvability of gene regulation for this trait. Overall, given that age-based trait variation is common across the tree of life, we provide a framework for understanding gene-regulatory-network-driven life-history evolution more broadly.
... Furthermore, many examples of adaptive cis-regulatory mutations tend to focus on trait loss rather than gain (Hoekstra and Coyne 2007). Instances include skeletal armor in three-spine sticklebacks (Shapiro, et al. 2004), pigmentation on Drosophila wings (Gompel, et al. 2005;Prud'homme, et al. 2006), and dorsal bristle density on Drosophila larvae (Sucena and Stern 2000). ...
Preprint
Full-text available
Determining the origins of novel genes and the genetic mechanisms underlying the emergence of new functions is challenging yet crucial for understanding evolutionary innovations. The novel fish antifreeze proteins, exemplifying convergent evolution, represent excellent opportunities to investigate the evolutionary origins and pathways of new genes. Particularly notable is the near-identical type I antifreeze proteins (AFPI) in four phylogenetically divergent fish taxa. This study tested the hypothesis of protein sequence convergence beyond functional convergence in three unrelated AFPI-bearing fish lineages, revealing different paths by which a similar protein arose from diverse genomic resources. Comprehensive comparative analyses of de novo sequenced genome of the winter flounder and grubby sculpin, available high-quality genome of the cunner, and those of 14 other relevant species found that the near-identical AFPI originated from a distinct genetic precursor in each lineage, and independently evolved coding regions for the novel ice-binding protein while retaining sequence identity in the regulatory regions with their respective ancestor. The deduced evolutionary processes and molecular mechanisms is consistent with the Innovation-Amplification-Divergence (IAD) model applicable to AFPI formation in all three lineages, a new Duplication-Degeneration-Divergence (DDD) model we propose for the sculpin lineage, and a DDD model with gene fission for the cunner lineage. This investigation illustrates the multiple ways by which a novel functional gene with sequence convergence at the protein level could evolve across divergent species, advancing our understanding of the mechanistic intricacies in new gene formation.
... For example, enhancers exhibit rapid turnover during mammalian evolution 4,5 , and conserved enhancers have lower cell type specificity 6,7 . By contrast, sequence divergent enhancers have a substantial role in establishing tissue and species-specific traits 8,9 . Such divergent enhancers are often mediated by de novo insertion of transposable elements (TEs) carrying clusters of transcription-factor-binding sites 6,10,11 . ...
Article
Full-text available
Divergence of cis-regulatory elements drives species-specific traits¹, but how this manifests in the evolution of the neocortex at the molecular and cellular level remains unclear. Here we investigated the gene regulatory programs in the primary motor cortex of human, macaque, marmoset and mouse using single-cell multiomics assays, generating gene expression, chromatin accessibility, DNA methylome and chromosomal conformation profiles from a total of over 200,000 cells. From these data, we show evidence that divergence of transcription factor expression corresponds to species-specific epigenome landscapes. We find that conserved and divergent gene regulatory features are reflected in the evolution of the three-dimensional genome. Transposable elements contribute to nearly 80% of the human-specific candidate cis-regulatory elements in cortical cells. Through machine learning, we develop sequence-based predictors of candidate cis-regulatory elements in different species and demonstrate that the genomic regulatory syntax is highly preserved from rodents to primates. Finally, we show that epigenetic conservation combined with sequence similarity helps to uncover functional cis-regulatory elements and enhances our ability to interpret genetic variants contributing to neurological disease and traits.
... Much of this variation is likely due to changes in the gene expression as a result of differences in genomic regulatory elements such as enhancers and promoters. Indeed, evidence indicates both that regulatory regions are subject to evolutionary constraints [1,2] and that mutations in regulatory elements can underlie phenotypic evolution [3][4][5][6]. Moreover, in humans, disease-and phenotype-associated variants identified through genome-wide association studies (GWASs) often are enriched in regulatory regions [7,8], further underscoring the importance of regulatory elements to the origins and evolution of specific phenotypes, including those related to disease susceptibility. ...
Article
Full-text available
Background The evolution of genomic regulatory regions plays a critical role in shaping the diversity of life. While this process is primarily sequence-dependent, the enormous complexity of biological systems complicates the understanding of the factors underlying regulation and its evolution. Here, we apply deep neural networks as a tool to investigate the sequence determinants underlying chromatin accessibility in different species and tissues of Drosophila . Results We train hybrid convolution-attention neural networks to accurately predict ATAC-seq peaks using only local DNA sequences as input. We show that our models generalize well across substantially evolutionarily diverged species of insects, implying that the sequence determinants of accessibility are highly conserved. Using our model to examine species-specific gains in accessibility, we find evidence suggesting that these regions may be ancestrally poised for evolution. Using in silico mutagenesis, we show that accessibility can be accurately predicted from short subsequences in each example. However, in silico knock-out of these sequences does not qualitatively impair classification, implying that accessibility is mutationally robust. Subsequently, we show that accessibility is predicted to be robust to large-scale random mutation even in the absence of selection. Conversely, simulations under strong selection demonstrate that accessibility can be extremely malleable despite its robustness. Finally, we identify motifs predictive of accessibility, recovering both novel and previously known motifs. Conclusions These results demonstrate the conservation of the sequence determinants of accessibility and the general robustness of chromatin accessibility, as well as the power of deep neural networks to explore fundamental questions in regulatory genomics and evolution.
... dopachrome conversion enzyme, DCE) from the yellow gene family advances melanin biosynthesis (38). This gene was previously found to mediate the melanin biosynthesis process in Drosophila, flour beetles (Tribolium) and the assassin bug (Platymeris biguttatus) (40)(41)(42)(43)(44). Prophenoloxidase can be activated into Phenoloxidase (45,46) subgenera. ...
Preprint
Full-text available
The metallic color variation of beetles is a spectacular feature that has inspired diverse human cultures. However, little is known about the genetic basis of this trait or its ecological importance. In this study, we characterize the geographical distribution, optical mechanism, genetic basis, and ecological and evolutionary importance of metallic color variation in the Nebria ingens complex, an alpine ground beetle in the Sierra Nevada, California. We find that elytral color varies continuously across two allopatric species (from black N. ingens to green N. riversi), with hybrid populations showing intermediate coloration, and we demonstrate that the metallic color is generated from multilayer reflectors in the epicuticle of the elytra. By applying association mapping in natural populations (wild-GWAS) using high-density SNPs (1.2 million), we identify five promising candidate genes covarying with metallic variation, with known roles in cuticle formation and pigmentation pathways. Among these five genes, the gene yellow-like exhibits a heightened divergence pattern relative to the background genomic landscape and has been maintained despite gene flow. This finding, together with a significant correlation between color variation and water availability, suggests that metallic variation evolves as a local adaptation to environmental variation in the N. ingens complex.
Preprint
Full-text available
The remarkable diversity of insect pigmentation offers a captivating avenue for exploring evolution and genetics. In tephritid fruit flies, decoding the molecular pathways underlying pigmentation traits also plays a central role in applied entomology. Mutant phenotypes like the black pupae (bp) have long been used as a component of genetic sexing strains, allowing male-only release in tephritid sterile insect technique applications. However, the genetic basis of bp remains largely unknown. Here, we present independent evidence from classical and modern genetics showing that the bp phenotype in the GUA10 strain of the Mexican fruit fly, Anastrepha ludens , is caused by a large deletion at the ebony locus resulting in the removal of the entire protein-coding region of the gene. Targeted knockout of ebony induced analogous bp phenotypes across six tephritid species spanning over 50 million years of divergent evolution. This functionally validated our findings and allowed for a deeper investigation into the role of Ebony in pigmentation and development in these species. Our study offers fundamental knowledge for developing new sexing strains based on the bp marker and for future evolutionary developmental biology studies in tephritid fruit flies.
Preprint
Full-text available
The evolution of insects has been marked by the appearance of key body plan innovations and novel organs that promoted the outstanding ability of this lineage to adapt to new habitats, boosting the most successful radiation in animals. To understand the origin and evolution of these new structures, it is essential to investigate which are the genes and gene regulatory networks participating during the embryonic development of insects. Great efforts have been made to fully understand, from a gene expression and gene regulation point of view, the development of holometabolous insects, in particular Drosophila melanogaster , with the generation of numerous functional genomics resources and databases. Conversely, how hemimetabolous insects develop, and which are the dynamics of gene expression and gene regulation that control their embryogenesis, are still poorly characterized. Therefore, to provide a new platform to study gene regulation in insects, we generated ATAC-seq (Assay for transposase-Accessible Chromatin using sequencing) for the first time during the development of the mayfly Cloeon dipterum. This new available resource will allow to better understand the dynamics of gene regulation during hemimetabolan embryogenesis, since C. dipterum belongs to the paleopteran order of Ephemeroptera, the sister group to all other winged insects. These new datasets include six different time points of its embryonic development and identify accessible chromatin regions corresponding to both general and stage-specific promoters and enhancers. With these comprehensive datasets, we characterised pronounced changes in accessible chromatin between stages 8 and 10 of embryonic development, which correspond to the transition from the last stages of segmentation to organogenesis and appendage differentiation. The application of ATAC-seq in mayflies has contributed to identify the epigenetic mechanisms responsible for embryonic development in hemimetabolous insects and it will provide a fundamental resource to understand the evolution of gene regulation in winged insects.
Article
Full-text available
Quantitative variation in attributes such as color, texture, or stiffness dominates morphological diversification. It results from combinations of alleles at many Mendelian loci. Here, we identify an additional source of quantitative variation among species, continuous evolution in a gene regulatory region. Specifically, we examined the modulation of wing pigmentation in a group of fly species and showed that inter-species variation correlated with the quantitative expression of the pigmentation gene yellow . This variation results from an enhancer of yellow determining darkness through species-specific activity. We mapped the divergent activities between two sister species and found the changes to be broadly distributed along the enhancer. Our results demonstrate that enhancers can act as dials fueling quantitative morphological diversification by modulating trait properties.
Article
Full-text available
Hindlimb loss has evolved repeatedly in many different animals by means of molecular mechanisms that are still unknown. To determine the number and type of genetic changes underlying pelvic reduction in natural populations, we carried out genetic crosses between threespine stickleback fish with complete or missing pelvic structures. Genome-wide linkage mapping shows that pelvic reduction is controlled by one major and four minor chromosome regions. Pitx1 maps to the major chromosome region controlling most of the variation in pelvic size. Pelvic-reduced fish show the same left–right asymmetry seen in Pitx1 knockout mice, but do not show changes in Pitx1 protein sequence. Instead, pelvic-reduced sticklebacks show site-specific regulatory changes in Pitx1 expression, with reduced or absent expression in pelvic and caudal fin precursors. Regulatory mutations in major developmental control genes may provide a mechanism for generating rapid skeletal changes in natural populations, while preserving the essential roles of these genes in other processes.
Book
This book focuses on drosophila as an especially useful model organism for exploring questions of evolutionary biology in the full range of evolutionary studies: population genetics, ecology, ecological genetics, speciation, phylogenetics, genome evolution, molecular evolution, and development. The author presents an integrated view of evolutionary biology as elucidated in this single organism. Special effort is made to point out holes in our knowledge and areas particularly ripe for new investigation.
Article
Here we describe how to generate customized microinjection needles from glass capillary tubes. Controls demonstrate the range of variables and effects on needle tip shape using a standard Flaming/Brown micropipet needle puller. Needles generated with two-cycle pulls provide a wider range of needle shapes in a predictable fashion. We used the needle puller's ramp function for multiple-cycle programs to determine the useful range of heat settings inherent to the glass capillary tube. This article focuses primarily on the preparation of injection needles utilized for P-element-mediated germ-line transformation in Drosophila melanogaster that do not require the dechorionation of the egg. However, these types of needles can be useful for numerous other types of injections, such as RNA interference, homologous recombination mutagenesis, morpholinos, transient gene regulation, drug delivery, and the transfer of cytoplasmic factors that are useful in a wide range of biological systems ranging from plants to vertebrates. Using our standard needle, we correlate the survival of injected D. melanogaster embryos with transformation efficiencies and plasmid construct characteristics.
Article
The domestication of all major crop plants occurred during a brief period in human history about 10,000 years ago. During this time, ancient agriculturalists selected seed of preferred forms and culled out seed of undesirable types to produce each subsequent generation. Consequently, favoured alleles at genes controlling traits of interest increased in frequency, ultimately reaching fixation. When selection is strong, domestication has the potential to drastically reduce genetic diversity in a crop. To understand the impact of selection during maize domestication, we examined nucleotide polymorphism in teosinte branched1, a gene involved in maize evolution. Here we show that the effects of selection were limited to the gene's regulatory region and cannot be detected in the protein-coding region. Although selection was apparently strong, high rates of recombination and a prolonged domestication period probably limited its effects. Our results help to explain why maize is such a variable crop. They also suggest that maize domestication required hundreds of years, and confirm previous evidence that maize was domesticated from Balsas teosinte of southwestern Mexico.
Article
Drosophila biarmipes males possess an apical black patch on their wings. In a laboratory stock of this species, males without such a patch have been found. Mating success of these two types of male was studied in an Elcns-Wattiaux mating chamber. Results show that males possessing the dark black patch have greater mating success than those without. This finding suggests that visual stimulus plays an important role in the mating behaviour of D. biarmipes.
Article
The Drosphila melanogaster species-group, established by Sturtevant (1942) for fourteen species, is now known to contain 115 described species here divided into twelve named subgroups (including one newly proposed), as well as further undescribed species. Three of the species, melanogaster, simulans and ananassae, are cosmopolitan; two others, kikkawai and malerkotliana, are widespread in the southern hemisphere, the latter apparently a recent introduction to South America. The greatest numbers of species otherwise occur in the Oriental region with smaller numbers in the Ethiopian, eastern Palaearctic and Australian regions and in several islands of the South Pacific. D.rajasekari and D.raychaudhurii are synonymized with D.biamipes;also D.andamanensis Parshad & Singh is synonymized with D.andamanensis Gupta & Raychaudhuri.
Article
Although Drosophila melanogaster is a paradigm eukaryote for biology, relationships of this species and the other 174 species in the melanogaster species group are poorly explored and ambiguous. Gene regions of Cytochrome oxidase II (mt:CoII), Alcohol dehydrogenase (Adh) and hunchback (hb) were sequenced and analysed phylogenetically to test prior hypotheses of relationships for the group based on chromosomes, morphology, and 28S rRNA gene sequences. A simultaneous cladistic analysis of the three newly sequenced gene regions produced a single well-resolved phylogeny for 49 exemplar species representing eight subgroups. Monophyly of each of the ananassae, melanogaster, montium, and takahashii subgroups is supported; the suzukii subgroup is polyphyletic. This phylogeny is consistent with variation in significant morphological structures, such as the male sex comb on the fore tarsus. The broad range of morphological variation among these species is interpreted and the applicability to evolution and developmental investigations is discussed. This phylogeny facilitates comparative investigations, such as gene family evolution, transposable element transmission, and evolution of morphological structures. © 2002 The Linnean Society of London, Biological Journal of the Linnean Society, 2002, 76, 21–37.