ArticlePDF Available

Candidate genes revealed by a genome scan for mosquito resistance to a bacterial insecticide: Sequence and gene expression variations

Authors:
  • French National Centre for Scientific Research (Grenoble, France)

Abstract and Figures

Genome scans are becoming an increasingly popular approach to study the genetic basis of adaptation and speciation, but on their own, they are often helpless at identifying the specific gene(s) or mutation(s) targeted by selection. This shortcoming is hopefully bound to disappear in the near future, thanks to the wealth of new genomic resources that are currently being developed for many species. In this article, we provide a foretaste of this exciting new era by conducting a genome scan in the mosquito Aedes aegypti with the aim to look for candidate genes involved in resistance to Bacillus thuringiensis subsp. israelensis (Bti) insecticidal toxins. The genome of a Bti-resistant and a Bti-susceptible strains was surveyed using about 500 MITE-based molecular markers, and the loci showing the highest inter-strain genetic differentiation were sequenced and mapped on the Aedes aegypti genome sequence. Several good candidate genes for Bti-resistance were identified in the vicinity of these highly differentiated markers. Two of them, coding for a cadherin and a leucine aminopeptidase, were further examined at the sequence and gene expression levels. In the resistant strain, the cadherin gene displayed patterns of nucleotide polymorphisms consistent with the action of positive selection (e.g. an excess of high compared to intermediate frequency mutations), as well as a significant under-expression compared to the susceptible strain. Both sequence and gene expression analyses agree to suggest a role for positive selection in the evolution of this cadherin gene in the resistant strain. However, it is unlikely that resistance to Bti is conferred by this gene alone, and further investigation will be needed to characterize other genes significantly associated with Bti resistance in Ae. aegypti. Beyond these results, this article illustrates how genome scans can build on the body of new genomic information (here, full genome sequence and MITE characterization) to finally hold their promises and help pinpoint candidate genes for adaptation and speciation.
Content may be subject to copyright.
BioMed Central
Page 1 of 11
(page number not for citation purposes)
BMC Genomics
Open Access
Research article
Candidate genes revealed by a genome scan for mosquito
resistance to a bacterial insecticide: sequence and gene expression
variations
Aurélie Bonin*, Margot Paris, Guillaume Tetreau, Jean-Philippe David and
Laurence Després
Address: Laboratoire d'Ecologie Alpine, CNRS-UMR 5553, Université Joseph Fourier, BP 53, 38041 Grenoble cedex 09, France
Email: Aurélie Bonin* - abonin@indiana.edu; Margot Paris - margot.paris@e.ujf-grenoble.fr; Guillaume Tetreau - guillaume.tetreau@e.ujf-
grenoble.fr; Jean-Philippe David - jean-philippe.david@ujf-grenoble.fr; Laurence Després - laurence.despres@ujf-grenoble.fr
* Corresponding author †Equal contributors
Abstract
Background: Genome scans are becoming an increasingly popular approach to study the genetic
basis of adaptation and speciation, but on their own, they are often helpless at identifying the
specific gene(s) or mutation(s) targeted by selection. This shortcoming is hopefully bound to
disappear in the near future, thanks to the wealth of new genomic resources that are currently
being developed for many species. In this article, we provide a foretaste of this exciting new era by
conducting a genome scan in the mosquito Aedes aegypti with the aim to look for candidate genes
involved in resistance to Bacillus thuringiensis subsp. israelensis (Bti) insecticidal toxins.
Results: The genome of a Bti-resistant and a Bti-susceptible strains was surveyed using about 500
MITE-based molecular markers, and the loci showing the highest inter-strain genetic differentiation
were sequenced and mapped on the Aedes aegypti genome sequence. Several good candidate genes
for Bti-resistance were identified in the vicinity of these highly differentiated markers. Two of them,
coding for a cadherin and a leucine aminopeptidase, were further examined at the sequence and
gene expression levels. In the resistant strain, the cadherin gene displayed patterns of nucleotide
polymorphisms consistent with the action of positive selection (e.g. an excess of high compared to
intermediate frequency mutations), as well as a significant under-expression compared to the
susceptible strain.
Conclusion: Both sequence and gene expression analyses agree to suggest a role for positive
selection in the evolution of this cadherin gene in the resistant strain. However, it is unlikely that
resistance to Bti is conferred by this gene alone, and further investigation will be needed to
characterize other genes significantly associated with Bti resistance in Ae. aegypti. Beyond these
results, this article illustrates how genome scans can build on the body of new genomic information
(here, full genome sequence and MITE characterization) to finally hold their promises and help
pinpoint candidate genes for adaptation and speciation.
Published: 21 November 2009
BMC Genomics 2009, 10:551 doi:10.1186/1471-2164-10-551
Received: 25 July 2009
Accepted: 21 November 2009
This article is available from: http://www.biomedcentral.com/1471-2164/10/551
© 2009 Bonin et al; licensee BioMed Central Ltd.
This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0),
which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
BMC Genomics 2009, 10:551 http://www.biomedcentral.com/1471-2164/10/551
Page 2 of 11
(page number not for citation purposes)
Background
In the past few years, evolutionary biologists have increas-
ingly bet on population genomics approaches to study the
genetic basis of adaptation and speciation. Genome scans
have flourished in the literature, providing valuable
insight into the genetics of local adaptation [1,2], sympat-
ric speciation [3], host race or ecotype differentiation [4-
7], and response to climate change or exotic invasions
[8,9], among others. In most cases, however, population
genomics alone fell short of pinpointing the specific
gene(s) or mutation(s) targeted by selection during the
adaptation or speciation process [10,11].
One of the main reasons for this recurring setback is the
present lack of genomic resources for most examined spe-
cies. For example, due to the absence of more powerful
alternatives, many population genomics studies rely on
genetic markers such as AFLPs (Amplified Fragments
Length Polymorphisms) [11,12], which can generally be
obtained easily for any organism. Unfortunately, such
markers present the double disadvantage of being anony-
mous and of falling predominantly in non-coding regions
of the genome, i.e. far from potential candidate regions
for adaptation and speciation [10]. Moreover, when
genomic sequences are scarce or poorly annotated, identi-
fying candidate genes in the vicinity of markers showing a
selection signature represents a daunting task that few
researchers have undertaken so far (but see [13]). Yet, we
predict that these limitations are bound to disappear in
the near future, especially with the increasing use of next-
generation sequencing technologies. In this article, we
aim at providing a foretaste of this exciting new era by
illustrating how new genomic tools can help unravel the
genetic basis of mosquito resistance to Bacillus thuringien-
sis subsp. israelensis (Bti) insecticidal toxins.
Considered as a safe alternative to chemical insecticides,
the bio-insecticide Bti is widely used worldwide for mos-
quito control [14]. Bti toxicity is mainly conferred by three
Cry toxins (Cry4A, Cry4B and Cry11A) and one Cyt toxin
(Cyt1A), which aggregate in a proteic crystal produced
during sporulation of the bacteria [15]. Bti is usually
sprayed in mosquito breeding sites as a mixture of spores
and toxins, which is ingested together with organic detri-
tus by developing larvae. In the larval midgut, Bti toxins
are first activated by protease/trypsin-like enzymes or ami-
nopeptidases [16]. Then, they bind to specific receptors of
the midgut cells (Cry toxins), or directly interact with the
cell membrane (Cyt1A toxin), ultimately causing pore for-
mation and cell lysis [16]. Cy1A is also known to act syn-
ergistically with Cry toxins, increasing the overall toxicity
of the Bti mixture [17,18]. Due to the complexity of Bti
toxicity mechanisms, some have argued that resistance to
Bti would likely require adaptive mutations in several
genes [19,20]. This argument is reinforced by the fact that
only a handful of studies have observed evidence of labo-
ratory or natural resistance to Bti in mosquito [19,21-23].
Because of the probable multilocus nature of Bti resist-
ance, population genomics appears to be an approach of
choice to identify genes involved in this process.
Here, we describe the application of population genomics
to the search for candidate genes for resistance to a toxic
leaf litter containing Bti spores. With the mosquito Aedes
aegypti as a model, we first conducted a genome scan rely-
ing on about 500 MITE (Miniature Inverted-repeat Trans-
posable Element)-derived markers expected to occur
frequently in gene-rich regions [24,25]. By combining the
results of this genome scan with data from the publicly
available genome sequence of Ae. aegypti, we were then
able to localize two good candidate genes for Bti resist-
ance. Finally, these two genes were further analyzed at the
sequence and gene expression levels in order to determine
if selection was indeed a driving force in their evolution.
Results
Genome scan and identification of outlier loci presumably
influenced by selection
Our search for candidate genes linked to Bti resistance was
conducted in two Aedes aegypti strains differing drastically
in their susceptibility to Bti as well as to individual Bti tox-
ins (see the Methods part for resistance ratios). The
genome of Ae. aegypti was screened using a variant of the
DArT (Diversity Arrays Technology) procedure, where
motifs of a particular MITE (Miniature Inverted-repeat
Transposable Element) family called Pony served as
primer anchors for PCR amplification. In total, 476 bial-
lelic dominant markers were surveyed for 29 individuals
in each mosquito strain, revealing a particularly high
genetic differentiation between strains (mean Fst =
0.556). This strong genetic structure was not surprising
given the history of the two strains, and in particular the
recent bottleneck experienced by the resistant strain
(Additional file 1). However, high neutral Fst values are
expected to reduce the power of methods revealing outlier
loci potentially under selection on the basis of an atypi-
cally high genetic differentiation. For example, the appli-
cation of the program Dfdist [26] to our data detected
only one locus departing from neutral expectations for α
= 1%, because the neutral envelope included almost the
entire range of possible differentiation values (Additional
file 2). As a result, we adopted a different strategy and
retained as outliers those loci for which alternative pheno-
types (fragment presence/absence) were fixed or nearly
fixed in each strain. A total of 70 such markers were
sequenced for further analyses.
Outlier sequencing and localization in the genome
Among the outlier sequences obtained, one pair differed
only by a gap and another one by only a mutation, result-
BMC Genomics 2009, 10:551 http://www.biomedcentral.com/1471-2164/10/551
Page 3 of 11
(page number not for citation purposes)
ing in a redundancy rate of 2.86%. After trimming the
primer sequence and the Pony motif, the 68 unique
marker sequences (GenBank accession no. FJ231034-
FJ231090; sequences shorter than 50 bp could not be
deposited) had an average size of 185.2 bp (range 16-868
bp). Of these unique sequences, 41 could be assigned to a
unique position in the Aedes aegypti genome, and all but
two of these positions were found on different supercon-
tigs (Additional file 3). Six sequences were situated on the
same supercontig as a candidate gene for Bti resistance
(DArT_102, DArT_318, DArT_400, DArT_415, DArT_432
and DArT_467), and two of them (DArT_432 and
DArT_467) co-localized with the same gene (a cadherin).
It had to be noted that physical distances between candi-
date genes and outlier markers situated on the same
supercontig were considerable, ranging from 97907 bp
(cadherin and DArT_467) to more than 300 Mbp (glyco-
syltransferase and DArT_415). According to these results,
the cadherin gene (CAD, VectorBase Gene ID
AAEL001196) turned out to be a serious candidate for Bti
resistance because two outlier markers pointed towards it,
one of them at the shortest distance recorded in this study.
This gene, which codes for a possible toxin-binding recep-
tor [27], was thus selected for further investigation at the
sequence and expression levels. We also focused on the
leucine aminopeptidase gene (LAP, VectorBase Gene ID
AAEL001649) because of its potential implication in Bti
toxin activation [28].
Candidate gene sequence analysis
A total of 1,467 bp and 1,657 bp were sequenced for the
CAD and LAP genes, respectively. All sequences were
deposited into GenBank, with accession numbers
GU066340 to GU066385. As shown in Table 1, the two
candidate genes shared similarities: both had a higher
genetic diversity in the susceptible than in the resistant
strain even if more individuals were sequenced in this lat-
ter. This observation was consistent with the fact that both
strains are separated by 18 generations of selection, and
that the resistant strain experienced a strong bottleneck at
generation 10 (Additional file 1). Both genes also showed
similar numbers of haplotypes in each strain (13 and 12
haplotypes in the susceptible strain for CAD and LAP,
respectively; and 5 haplotypes in the resistant strain for
both genes). On the other hand, the two candidate genes
differed by their overall level of genetic diversity and inter-
strain differentiation. For example, the Fst value was 0.186
only for CAD vs. 0.321 for LAP. Likewise, the nucleotide
diversity estimated for LAP in the susceptible strain was
about four times lower than that estimated for CAD, this
ratio reaching 17 when considering the resistant strain.
However, it has to be noted that polymorphisms were not
uniformly distributed along the CAD gene sequence,
some regions showing globally more variation (e.g. sub-
domain SD4 and interdomain SD4-SD5 of the protein;
Figure 1).
These patterns of genetic diversity translated differently at
the codon level. We found three non-synonymous sites in
the CAD sequence, of which none was diagnostic of one
strain in particular, whereas for the LAP sequence, only
two of nine non-synonymous mutations were present in
the resistant strain. Six non-synonymous sites, among
which five could only be found in the susceptible strain,
were noticeably situated in 5' and 3' untranslated regions
and could thus affect transcript stability and/or transla-
tion.
Finally, the examination of haplotype repartition between
strains allowed this picture to be completed. For the CAD
gene, one haplotype present in all 12 sampled resistant
individuals (18 haplotypes out of 24) existed only in 2
susceptible individuals (2 haplotypes out of 20). For the
LAP gene, the most frequent haplotype in the resistant
strain (11 haplotypes out of 24) was absent from the sus-
ceptible strain.
Neutrality tests
For both candidate genes, we tested for deviation from
neutral evolution in the resistant strain by applying differ-
ent neutrality tests to the observed polymorphism data
(Table 2). Tajima's D and Fay and Wu's H highlight a skew
in the frequency distribution of variants, H giving more
weight to high-frequency polymorphisms, whereas Fu
and Li's D* and F* detect a discrepancy between either the
total number of mutations (D*) or the average number of
differences between two sequences (F*) and the number
of singletons. As a result, each one of these tests is based
on different diversity parameters and gives different infor-
mation on the type of selection presumably in action. In
this study, every test except Fay and Wu's H gave positive
values for both genes (Table 2), which is consistent with
the recent bottleneck experienced by the resistant strain.
None of the tests was significant for LAP, whereas D*, F*
and H remained significant for CAD even when account-
ing for the particular demographic history of the strain
(Table 2). The positive values of D* and F* suggested a
deficiency of recent (i.e. rare) mutations, whereas the neg-
ative H indicated an excess of high compared to interme-
diate frequency mutations.
Gene expression analyses
Real-time reverse-transcription PCR (RT-PCR) analyses
revealed that levels of gene expression were reduced in the
resistant strain compared to the susceptible strain for both
candidate genes (2.17-fold and 1.68-fold under-expres-
sion for the CAD and the LAP genes, respectively; Figure
2). Non-parametric Mann-Whitney tests indicated that
reduction in expression levels was significant for both
genes (p = 0.017 and p = 0.002 for the CAD and LAP genes,
respectively). However, only the expression fold change
observed for the CAD gene (2.17) was higher than the
BMC Genomics 2009, 10:551 http://www.biomedcentral.com/1471-2164/10/551
Page 4 of 11
(page number not for citation purposes)
two-fold change conservatively used as a significant
threshold in expression studies [29].
Discussion
A MITE-based genome scan to search for candidate genes
The goal of this study was to identify genes conferring
resistance to the bio-insecticide Bacillus thuringiensis
subsp. israelensis (Bti) in an Aedes aegypti mosquito strain
selected for several generations with a toxic leaf litter con-
taining Bti spores. The genetic basis of Bti resistance is
likely to be multigenic [19,20], and in Ae. aegypti, we listed
as many as 160 serious candidates by virtue of their
known function and/or proven association with Bti resist-
ance in other species (e.g. [20,30,31]). As a result, we
chose to tackle this study by adopting a population
genomics approach. The underlying idea was to examine
many MITE-derived DArT markers scattered in the
genome to get an accurate estimate of the overall back-
ground (i.e. neutral) genetic differentiation between the
selected (Bti-resistant) and control (Bti-susceptible)
strains. This in turn allowed detecting markers with an
atypically high inter-strain genetic differentiation, and
thus possibly linked to a gene under positive selection due
to Bti resistance. Our genome scan revealed an overall
high level of neutral genetic differentiation between the
two strains (mean Fst = 0.556), which is not unexpected
in the light of the history of the resistant strain, and espe-
cially of the bottleneck it experienced at generation 10
Table 1: Measures of genetic diversity and divergence obtained for the two candidate genes
Gene/Domain Strain Fragment size (bp) NHap S SiNS Hd WkKFst
Cadherin
(CAD)
Whole
sequence*
Susceptible
Resistant
1467 20
24
13
5
67
65
0
1
3
3
0.932
0.435
0.023
0.017
0.013
0.012
32.363
25.033
35.254 0.186
SD2 Susceptible
Resistant
75 22
24
3
2
2
2
0
0
0
0
0.567
0.391
0.013
0.010
0.007
0.007
0.961
0.783
1.136 0.233
SD2-SD3 Susceptible
Resistant
69 22
24
3
2
2
1
1
0
0
0
0.558
0.391
0.009
0.006
0.008
0.004
0.610
0.391
0.568 0.119
SD3 Susceptible
Resistant
268 22
24
3
2
12
8
4
0
1
0
0.541
0.391
0.016
0.012
0.012
0.008
4.264
3.130
4.705 0.214
SD3-SD4 Susceptible
Resistant
77 22
24
2
2
3
3
0
0
2
1
0.485
0.391
0.019
0.015
0.011
0.010
1.455
1.174
1.705 0.229
SD4 Susceptible
Resistant
279 22
24
4
2
18
14
4
0
0
0
0.571
0.391
0.026
0.020
0.018
0.013
7.273
5.478
8.080 0.211
SD4-SD5 Susceptible
Resistant
69 22
24
3
2
6
5
1
0
0
0
0.541
0.391
0.038
0.028
0.024
0.019
2.602
1.957
2.795 0.185
SD5 Susceptible
Resistant
246 10
22
4
4
17
14
1
0
1
1
0.733
0.398
0.034
0.020
0.024
0.016
8.422
5.039
8.164 0.176
SD5-SD6 Susceptible
Resistant
63 10
22
2
2
1
1
0
0
0
0
0.533
0.368
0.008
0.006
0.006
0.004
0.533
0.368
0.555 0.187
SD6 Susceptible
Resistant
279 10
22
2
2
15
15
0
0
1
1
0.533
0.368
0.029
0.020
0.019
0.015
8.000
5.519
8.318 0.187
SUP6 Susceptible
Resistant
42 10
22
2
2
1
1
0
0
0
0
0.533
0.368
0.013
0.009
0.008
0.007
0.533
0.368
0.555 0.187
Leucine aminopeptidase (LAP) Susceptible
Resistant
1657 22
24
12
5
29
6
0
0
9
2
0.900
0.688
0.006
0.001
0.005
0.001
10.069
2.025
8.909 0.321
N, number of alleles sampled; Hap, number of different haplotypes; S, number of segregating sites; Si, number of singletons; NS, number of non-
synonymous mutations; Hd, haplotype diversity; , nucleotide diversity (per site); W, Watterson's mutation parameter θ estimated from S (per site);
k; average number of nucleotide differences within strain; K, average number of nucleotide differences between strains; Fst, genetic differentiation
between strains.
* Some measures (e.g. S and NS) are different between the whole sequence and the total of all sequences because of a different number of sampled
haplotypes.
BMC Genomics 2009, 10:551 http://www.biomedcentral.com/1471-2164/10/551
Page 5 of 11
(page number not for citation purposes)
(see Additional File 1). Yet, the Fst value obtained
between the same two mosquito strains using AFLP mark-
ers is substantially lower (mean Fst = 0.114; Paris, pers.
comm.). AFLP and DArT markers are biallelic, dominant
and, to a certain extent, randomly distributed in the
genome so they should provide similar genetic differenti-
ation estimates. However, the DArT markers developed
for the purpose of this study are intimately associated with
a specific family of MITEs called Pony, whose characteris-
tics might explain the discrepancy between the two Fst
measures. Pony elements constitute about 1.1% of the
genome of Ae. aegypti [25], and although their transposi-
tion mechanism is still unclear, it could be triggered by
unfavorable environmental conditions as was shown for
other MITEs in plants for example [24,32,33]. Several
authors have even underlined the potential contribution
of MITEs to rapid adaptations and, ultimately, genome
evolution [24,34]. One can thus speculate that the envi-
ronmental stress imposed by toxic leaf litter selection
stimulated Pony transpositions, hence inflating estimates
of inter-strain genetic differentiation as measured by our
Pony-associated markers. Like many MITEs, Pony motifs
are also known to frequently occur in the non-coding
regions of genes [25]. Because of this last characteristic of
Pony elements and of their high mutational potential pos-
sibly enhanced by stress, Pony-based DArTs are ideal ran-
dom markers to explore the genome of Ae. aegypti and
search for genes conferring resistance to Bti.
DArT marker sequences and identification of candidate
genes
We decided to sequence the MITE-based DArT markers
showing the highest inter-strain differentiation (outliers),
as those are the most likely to be linked to genes confer-
ring resistance to Bti. Of the 70 sequences obtained, 41
(68.6%) matched to unique locations in the genome of
Comparison between lepidopteran cadherin-like proteins and the Aedes aegypti cadherin studied hereFigure 1
Comparison between lepidopteran cadherin-like pro-
teins and the Aedes aegypti cadherin studied here.
Cadherin-like proteins are constituted of different domains:
SD, subdomain; MPED, membrane-proximal extracellular
domain; TM, transmembrane domain; CYTO, cytoplasmic
domain. Only features present in the mature form of the
protein are outlined here. Known and putative Cry binding
sites characterized in lepidopterans are indicated by paren-
theses. B. mori, Bombyx mori; H. armigera, Helicoverpa armig-
era; H. virescens, Heliothis virescens; M. sexta, Manduca sexta;
Ae. aegypti, Aedes aegypti. Adapted from Figure 10 in [27].
Table 2: Results of the neutrality tests for the resistant strain
Neutrality test Value Significance according to coalescent sim-
ulations based on a large constant popu-
lation sizea
Significance according to coalescent simu-
lations based on the known demographic
history of the resistant strainb
Cadherin (CAD)
Tajima's D1.717 p < 0.01 n.s.
Fu and Li's D* 1.698 p < 0.001 p < 0.01
Fu and Li's F* 2.000 p < 0.001 p < 0.05
Fay and Wu's H-50.370 p < 0.001 p < 0.01
Leucine aminopeptidase (LAP)
Tajima's D0.788 n.s. n.s.
Fu and Li's D* 1.233 n.s. n.s.
Fu and Li's F* 1.280 n.s. n.s.
Fay and Wu's H1.181 n.s. n.s.
a, performed following the method of Hudson [70] implemented in DnaSP; b, performed using the program ms [68].
BMC Genomics 2009, 10:551 http://www.biomedcentral.com/1471-2164/10/551
Page 6 of 11
(page number not for citation purposes)
Ae. aegypti, while the rest consisted of repeated sequences.
As less than 40% of the genome of this species is consti-
tuted of single or low copy sequences [35], this result con-
firms the fact that MITE-based DArTs tend to fall near low-
copy, possibly coding, sequences. The 41 genomic loca-
tions involved were all but two situated on different
supercontigs of the Ae. aegypti full genome sequence, sug-
gesting that there was no real bias in the distribution of
our markers in the genome, at least for highly differenti-
ated ones. Unfortunately, the Ae. aegypti genome has a
weak gene density [35], and most of our outlier sequences
(85.4%) fell in supercontigs with few or no putative
genes. Furthermore, we recently performed a differential
transcriptome analysis of Ae. aegypti larvae exposed to dif-
ferent xenobiotics. This study, based on the sequencing of
millions of cDNA tags using next-generation sequencing,
detected more than 2000 loci situated outside predicted
genes and showing a significant transcription signal
(David, pers. comm.). There is thus an obvious need for a
better gene annotation in the Ae. aegypti genome, and one
efficient strategy to tackle this task would be to concen-
trate the effort on regions of interest, like those showing
signatures of selection.
The selection with toxic leaf litter was a recent event in the
history of the resistant strain so linkage disequilibrium is
probably extensive in the vicinity of selected genes [36]. In
those conditions, detecting selection signatures in a
genome is somewhat easier even with a relatively small
marker density [37], but on the other hand, the existence
of large haplotype blocks makes it more difficult to pin-
point the exact gene(s) under selection. In this study, the
distance between outlier markers and candidate genes was
considerable (between 100 and 3,000 Kb) and might
exceed the window of linkage disequilibrium around
selected genes. In other words, we might have overlooked
closer genes genuinely responsible for the signatures of
selection, but not already annotated or absent from our
list of candidates.
The cadherin and leucine aminopeptidase as candidate
genes
Among the five candidate genes discovered in the proxim-
ity of outlier markers, we decided to examine further the
cadherin (CAD) and leucine aminopeptidase (LAP) at the
sequence and expression levels, to confirm or infirm the
selection footprint. This step was all the more crucial since
selection but also particular demographic histories can
generate an atypically high genetic differentiation [38,39].
Neutrality tests are often used to verify the influence of
selection in intraspecific sequence data [40,41]. However,
they have often been criticized for their lack of power and
their sensitivity to demographic events like bottlenecks or
population expansions which can mimic selective effects
[38,42]. In this study, this possible bias was overcome by
performing tailor-made coalescence simulations based on
the known demographic history of the resistant strain, in
order to assess the significance of the tests. No firm evi-
dence of selection was found at the nucleotide or expres-
sion level for the LAP gene, which rules out its implication
in Bti resistance, at least as a gene with major effects. For
the CAD gene, Fu and Li's D* and F* tests highlighted a
deficit in rare mutations, which is usually the trademark
of balancing selection [43]. Fay and Wu's H statistics indi-
cated an excess of high compared to intermediate fre-
quency mutations, which on the contrary suggests the
spread of an advantageous mutation at a linked site (i.e.
positive selection) [44]. Although those conflicting results
might be hard to interpret at first glance, one has to keep
in mind that the selection with the toxic leaf litter started
only 20 generations ago. In addition, the resistant strain
experienced a recent bottleneck which probably further
eliminated low frequency variants. New mutations cer-
tainly have not had time yet to appear in the population,
which would explain the positive D* and F* values. For
the same reason, Tajima's D can be transiently positive
right after a bottleneck [45]. As for the H statistics, it is not
as sensitive to the loss of rare variants, so it is presumably
more reliable here and we can reasonable think that the
CAD gene shows genuine signs of positive selection. This
evidence is further reinforced by the significant under-
expression of this gene in the resistant strain. Cadherins
are indeed known to bind to Cry toxins in other insect
species [27] and to be involved in many cases of Cry resist-
ance in insects (e.g. [46-48]). Here, under-expression of
the cadherin gene could reduce the number of a certain
type of Cry receptors and thus hinder pore forming, ulti-
Results of the expression analyses for the two candidate genesFigure 2
Results of the expression analyses for the two candi-
date genes. This figure illustrates, for each candidate gene,
the mean gene expression in the resistant strain relative to
that in the susceptible strain.
BMC Genomics 2009, 10:551 http://www.biomedcentral.com/1471-2164/10/551
Page 7 of 11
(page number not for citation purposes)
mately limiting susceptibility to Bti. Similarly, resistance
to the toxin Cry1Ac has been shown to be linked to
reduced levels of membrane receptors in the cotton pest
Heliothis virescens [49], or in the cabbage moth Plutella
xylostella [50]. Nevertheless, cadherins are usually specific
to one particular Cry toxin [27], and it is unlikely that the
CAD gene alone is responsible for resistance to Bti which
is a mixture of several Cry toxins. One can thus hypothe-
size that this gene's effects supplement those of one or sev-
eral other genes that remain to be identified.
Conclusion
Beyond their implications for the understanding of the
genetic mechanisms of Bti mosquito resistance, these
results illustrate how genome scans can build on the body
of new genomic information (here, full genome sequence
and MITE characterization) to finally hold their promises
and help pinpoint candidate genes for adaptation and
speciation [10]. In the near future, a wealth of genomics
tools will be available for a much wider range of species,
mostly thanks to the rapid development of next-genera-
tion sequencing technologies [51]. We predict this new
knowledge will boost in many respects the use of popula-
tion genomics for the study of the genetic bases of adapta-
tion and speciation. Several limitations of current genome
scans will certainly be soon overcome, with for example
(1) the development of new genetic markers allowing
screening the genome more finely (e.g. [52]) or specifi-
cally targeting coding regions (e.g. [53]); (2) an easier
access to outlier sequences as well as full genomic
sequences serving as references to locate outlier loci and
identify nearby candidate genes; and (3) a better gene
annotation. In short, population genomics will at last
have the means to meet our expectations when it comes to
identify genes under natural or artificial selection.
Methods
Biological material
Two Aedes aegypti laboratory strains were compared for the
purpose of this study: the Bora Bora reference strain,
known to be susceptible to most insecticides, and a strain
artificially selected for resistance to a decomposed tree leaf
litter showing a high toxicity for mosquito larvae. This leaf
litter had been collected in a mosquito pond in Eastern
France three months after treatment for mosquito control,
and has been proved to contain Bacillus thuringiensis
subsp. israelensis (Bti) spores from commercial origin [54].
The use of this Bti-contaminated leaf litter in the selection
experiments allowed mimicking the evolution of resist-
ance to Bti in a situation close to field conditions. The sus-
ceptible and resistant strains were separated by 18
generations of selection (and by 20 generations in total),
with a strong bottleneck at generation 10 (see Additional
File 1 for more details on the demographic history of the
resistant strain). The selection experiment and the bio-
assays implemented to monitor the evolution of resist-
ance are described in [55]. At each generation, the lethal
dose for 50% of the individuals after a 24 h-exposure (24
h-LD50) was determined for each strain using the Probit
software [56]. After 18 generations of selection, the resist-
ance ratios RR of the resistant strain (i.e., the ratio
between the 24 h-LD50 values for the resistant and the
susceptible strains, respectively) were 3.4-fold, 30.2-fold,
13.7-fold, 6.3-fold and 3-fold for the toxic leaf litter,
Cry4A, Cry4B, Cry11A and Cyt1A toxins, respectively.
DNA extraction and genome scan
The genomic DNA used for all subsequent molecular
work was extracted from fresh fourth-instar mosquito lar-
vae using the Qiagen DNeasy Tissue Kit and protocol
(Qiagen). Prior to extraction, the larvae midgut was
removed carefully to avoid bacterial contamination.
The classical protocol of the Diversity Array Technology
(DArT) [57] was slightly modified so as to provide hun-
dreds of good-quality markers scattered in the genome of
Ae. aegypti and possibly associated to gene-rich regions
[55]. Briefly, in a first step, genomic DNA was digested
with restriction enzyme Bsp1286I and a specific adaptor
was ligated to compatible ends. Restriction fragments
including a particular Ae. aegypti MITE called Pony were
PCR-amplified using a primer annealing to the adaptor
sequence and a primer complementary to a conserved
motif of the Pony element. PCR products obtained for all
individuals of the two strains were pooled together and
cloned to construct a DArT library containing a total of
6144 MITE-based clones. In a second step, a labelled tar-
get produced for each Ae. aegypti individual as described
in the first step was hybridized to the library fragments
spotted on a glass slide in order to reveal the polymorphic
ones. Details of the protocol, in particular the adaptor and
primer sequences used and the reproducibility rates, can
be found in [55].
Identification and sequencing of outlier loci potentially
under selection
For each DArT marker obtained, allelic frequencies were
estimated with the Bayesian method with non-uniform
prior distribution [58] implemented in AFLP-SURV 1.0
[59]. Among those markers, we tracked those for which
alternative phenotypes (fragment presence/absence) were
fixed or nearly fixed in the two strains (for example, frag-
ment present/absent for all individuals or all individuals
except one). Our assumption was that such a pattern of
extreme inter-strain genetic differentiation could be
explained by the spread, in the resistant strain, of an
advantageous allele initially present at low or intermedi-
ate frequency in the susceptible strain (standing varia-
tion), from which it is eventually purged by genetic drift
and/or because it is slightly deleterious.
BMC Genomics 2009, 10:551 http://www.biomedcentral.com/1471-2164/10/551
Page 8 of 11
(page number not for citation purposes)
For 70 such markers, bacterial cultures were sent to
Genome Express® http://www.genome-express.com for
insert amplification and sequencing with M13 forward
and M13 reverse primers. Raw sequence files were edited
with BioEdit 7.0.9 [60] and purged from Pony and primer
sequences. The obtained sequenced were blasted against
the full genomic sequence of Ae. aegypti (consisting of
4758 supercontigs and available at http://aaegypti.vector
base.org/GetData/Downloads?type=Genome).
Identification of candidate genes
Although mechanisms of resistance to Bti are still
unknown in dipterans, resistances to several Cry toxins
have been intensively studied, especially in lepidopteran
pests resistant to transgenic crops expressing Bacillus thur-
ingiensis Cry toxins genes [20,30]. Because these toxins
share similar three-dimensional structures, similar modes
of action and resistance mechanisms can be expected
between lepidopteran and dipteran insects [16,61]. We
therefore considered as candidate genes for Bti resistance
those belonging to families previously proved to be
involved in Cry resistance [20,30,31]. To this list, we
added genes potentially implicated in activation of Bti
toxins (aminopeptidases, e.g. [28]; and trypsins and chy-
motrypsins, e.g. [62]), in toxin binding (alkaline phos-
phatases, e.g. [49,63]; aminopeptidases, e.g. [64];
cadherins, e.g. [48]; galactosidases and glycosyltrans-
ferases, e.g. [20,65]); or immune defense (mitogen-acti-
vated protein kinases, e.g. [31]). A keyword search was
conducted in the VectorBase database http://aaegypti.vec
torbase.org/index.php and a total of 160 candidates
located on 98 different supercontigs were identified out of
the 15,419 putative genes (16789 transcripts in total) ref-
erenced in the Aedes aegypti genome.
Cadherin (CAD) and leucine aminopeptidase (LAP) gene
sequencing
The complete genomic sequences of the CAD and LAP
genes (VectorBase Gene IDs AAEL001196 and
AAEL001649, respectively) were downloaded from the
VectorBase website to help design sequencing primers
(Additional file 4) with the software package Lasergene
7.2 (DNASTAR Inc.). The sequencing strategy for the CAD
gene targeted exon 5 and more specifically the membrane-
proximal subdomains (subdomains 4 to 6) of the protein
which are the preferential binding sites of Bti Cry toxins in
lepidopterans (Figure 1). For the LAP gene, three different
primer pairs were selected to amplify the two main exons.
PCR amplifications were conducted for each gene in a 25-
μl total volume with 2 mM MgCl2, 0.1 mM of each dNTP
(Roche), 0.2 μM of each primer, 5 μg of BSA, 0.6 U of
AmpliTaq Gold DNA polymerase (Applied Biosystems)
and 10-30 ng of DNA. The PCR program included an ini-
tial 10-min denaturation step at 95°C; 40 cycles of dena-
turation at 95°C for 45s, annealing at the optimal
temperature indicated in Additional file 4 for 45s and
elongation at 72°C for 60s; followed by a final extension
step at 72°C for 5 min. PCR products were purified with
the QIAquick PCR purification kit (Qiagen) and sequenc-
ing reactions were performed in both directions using the
amplification primers and the BigDye Terminator Cycle
Sequencing Kit 3.1 (Applied Biosystems), following the
manufacturer's indications. Fluorescently labelled
sequencing products were run on an ABI PRISM 3100 cap-
illary DNA sequencer (Applied Biosystems) and
sequences were analyzed with SeqMan Pro 7.1.0 (DNAS-
TAR Inc.). Overall, we obtained sequences for 11 and 12
individuals of the susceptible and resistant strains, respec-
tively.
Sequence analysis and neutrality tests
The software DnaSP 5.0 [66] was used to infer haplotype
phase and to assess a variety of genetic diversity and dif-
ferentiation parameters (e.g., nucleotide diversity p, hap-
lotype diversity Hd, number of segregating sites S, Fst, etc.)
for each gene. Several statistics were also calculated based
on the observed polymorphism data to test for deviation
from neutral evolution in the resistant strain, including
Tajima's D [67], Fu and Li's D* and F* [43], and Fay and
Wu's H [44]. To assess whether these statistics signifi-
cantly departed from a neutral scenario of evolution given
the known demographic history of the resistant strain, we
performed coalescent simulations using the program ms
[68]. This program generates random independent sam-
ples according to a Wright-Fisher neutral model allowing
population size changes in the past. For each gene, the
mutation rate μ was estimated from the per-locus muta-
tion parameter ? observed for the susceptible strain (θ =
4Neμ and Ne = 6000) and used as the starting value for the
simulations (that is, as the value at present). Then 1000
neutral samples consisting of 24 haplotypes were simu-
lated based on the known demographic history of the
resistant strain (Additional file 1).
RNA extraction and gene expression analyses
For the two candidate genes, real-time reverse-transcrip-
tion PCR (RT-PCR) analyses were performed on three bio-
logical replicates for each strain, with each replicate
consisting of 30 larvae reared in standard insectary condi-
tions up to the fourth-instar stage (5 days). Total RNAs
were extracted using TRIzol (Invitrogen) following the
manufacturer's instructions and their quality was assessed
with a 2100 Bioanalyzer (Agilent) after DNase I (Invitro-
gen) treatment. Four micrograms of total RNA were
digested with DNase I (Invitrogen) and then used for first-
strand cDNA synthesis with SuperScript III (Invitrogen)
reverse transcriptase and oligo-dT20 primers for 60 min at
50°C, according to the manufacturer's instructions. Real-
time RT-PCR reactions were performed on an iQ5 system
BMC Genomics 2009, 10:551 http://www.biomedcentral.com/1471-2164/10/551
Page 9 of 11
(page number not for citation purposes)
(Bio-Rad) in a 25-μL total reaction volume with 0.3 μM of
each primer, 12.5 μL of iQ SYBR Green supermix (Bio-
Rad) and 5 μl of cDNA diluted 25 times. The real-time RT-
PCR program included an initial 3-min denaturation step
at 95°C and 40 cycles of denaturation at 95°C for 15s and
annealing for 30s at the optimal temperature indicated in
Additional file 4. For each gene, real-time RT-PCR effi-
ciency was estimated from a serial dilution of cDNA (5 to
500 times) and taken into account in the data analysis
performed with the ΔΔCT method [69]. Two housekeep-
ing genes encoding ribosomal protein L8 (RPL8, GenBank
accession number: DQ440262) and S7 (RPS7, GenBank
accession number: AY380336) were used for normaliza-
tion. Results were represented as mean expression ratios
between Bti-resistant and susceptible larvae (± SE).
Abbreviations
Bti: Bacillus thuringiensis subsp. israelensis; CAD: cadherin;
DArT: Diversity Arrays Technology; LAP: leucine ami-
nopeptidase; MITE: miniature inverted-repeat transposa-
ble element; RT-PCR: reverse-transcription PCR.
Authors' contributions
AB carried out the genome scan, analyzed the candidate
gene sequence data and drafted the manuscript. MP
worked on the outlier sequences, identified the two candi-
date genes, helped with the analyses and wrote substantial
parts of the paper. GT obtained and analyzed the candi-
date gene sequence and expression data and was involved
in the writing. JPD supervised the gene expression study
and helped draft the manuscript. LD conceived the overall
study, performed the demographic simulations and took
part to the data analysis and to the writing. All authors
read and approved the final manuscript.
Additional material
Acknowledgements
The authors would like to thank Sébastien Boyer for help with the insecti-
cide selection and Andrzej Kilian for guidance on the DArT technique. AB
and LD were funded by the Région Rhône-Alpes (grants #0501553401 and
#0501545401, respectively) and MP, JPD and LD benefited from a collabo-
rative grant attributed by the Démoustication Rhône-Alpes. GT was sup-
ported by the French Ministry of Research.
References
1. Bonin A, Taberlet P, Miaud C, Pompanon F: Explorative genome
scan to detect candidate loci for adaptation along a gradient
of altitude in the common frog (Rana temporaria). Mol Biol Evol
2006, 23:773-783.
2. Meyer CL, Vitalis R, Saumitou-Laprade P, Castric V: Genomic pat-
tern of adaptive divergence in Arabidopsis halleri, a model
species for tolerance to heavy metal. Mol Ecol 2009,
18:2050-2062.
3. Savolainen V, Anstett MC, Lexer C, Hutton I, Clarkson JJ, Norup MV,
Powell MP, Springate D, Salamin N, Baker WJ: Sympatric specia-
tion in palms on an oceanic island. Nature 2006, 441:210-213.
4. Nosil P, Egan SP, Funk DJ: Heterogeneous genomic differentia-
tion between walking-stick ecotypes: "Isolation by adapta-
tion" and multiple roles for divergent selection. Evolution
2008, 62:316-336.
5. Emelianov I, Marec F, Mallet J: Genomic evidence for divergence
with gene flow in host races of the larch budmoth. P Roy Soc
B-Biol Sci 2004, 271:97-105.
6. Campbell D, Bernatchez L: Generic scan using AFLP markers as
a means to assess the role of directional selection in the
divergence of sympatric whitefish ecotypes. Mol Biol Evol 2004,
21:945-956.
7. Galindo J, Morán P, Rolán-Alvarez E: Comparing geographical
genetic differentiation between candidate and noncandidate
loci for adaptation strengthens support for parallel ecologi-
cal divergence in the marine snail Littorina saxatilis . Molecular
Ecology 2009, 18:919-930.
8. Jump AS, Hunt JM, Martínez-Izquierdo JA, Peñuelas J: Natural selec-
tion and climate change: temperature-linked spatial and
temporal trends in gene frequency in Fagus sylvatica . Mol Ecol
2006, 15:3469-3480.
9. Mealor BA, Hild AL: Potential selection in native grass popula-
tions by exotic invasion. Mol Ecol 2006, 15:2291-2300.
10. Bonin A: Population genomics: a new generation of genome
scans to bridge the gap with functional genomics. Mol Ecol
2008, 17:3583-3584.
11. Stinchcombe JR, Hoekstra HE: Combining population genomics
and quantitative genetics: finding the genes underlying eco-
logically important traits. Heredity 2008, 100:158-170.
12. Luikart G, England PR, Tallmon D, Jordan S, Taberlet P: The power
and promise of population genomics: from genotyping to
genome typing. Nat Rev Genet 2003, 4:981-994.
Additional file 1
Demographic history of the Aedes aegypti Bti-resistant strain. The
Bti-resistant strain was originally selected from the susceptible standard
Bora-Bora strain. This table presents the effective population size at each
generation of selection.
Click here for file
[http://www.biomedcentral.com/content/supplementary/1471-
2164-10-551-S1.DOC]
Additional file 2
Results of the Dfdist analysis with = 1%. In this plot of inter-strain Fst
values against heterozygosity estimates, each dot represents a DArT
marker. The red lines represent the 99% neutral confidence interval sim-
ulated using the program Dfdist [26]. Markers situated outside this inter-
val diverge from neutral expectations and are thus potentially under
selection. Here, the confidence interval is so large that it includes almost
the entire range of possible Fst values.
Click here for file
[http://www.biomedcentral.com/content/supplementary/1471-
2164-10-551-S2.DOC]
Additional file 3
Outlier markers with a unique localization in the genome of Aedes
aegypti. This table lists the 41 outlier markers with a unique localization
in the genome of Aedes aegypti, as well as the numbers of annotated
genes and candidate genes in the corresponding supercontigs.
Click here for file
[http://www.biomedcentral.com/content/supplementary/1471-
2164-10-551-S3.DOC]
Additional file 4
Primer pairs used for sequencing and real-time RT-PCR analyses. This
table details the different primer pairs used in this study.
Click here for file
[http://www.biomedcentral.com/content/supplementary/1471-
2164-10-551-S4.DOC]
BMC Genomics 2009, 10:551 http://www.biomedcentral.com/1471-2164/10/551
Page 10 of 11
(page number not for citation purposes)
13. Wood HM, Grahame JW, Humphray S, Rogers J, Butlin RK:
Sequence differentiation in regions identified by a genome
scan for local adaptation. Mol Ecol 2008, 17:3123-3135.
14. Lacey LA: Bacillus thuringiensis serovariety israelensis and Bacil-
lus sphaericus for mosquito control. J Am Mosq Control Assoc
2007, 23:133-163.
15. Schnepf E, Crickmore N, Van Rie J, Lereclus D, Baum J, Feitelson J,
Zeigler DR, Dean DH: Bacillus thuringiensis and its pesticidal
crystal proteins. Microbiol Mol Biol Rev 1998, 62:775-806.
16. Bravo A, Gill SS, Soberón M: Mode of action of Bacillus thuring-
iensis Cry and Cyt toxins and their potential for insect con-
trol. Toxicon 2007, 49:423-435.
17. Pérez C, Fernandez LE, Sun JG, Folch JL, Gill SS, Soberón M, Bravo A:
Bacillus thuringiensis subsp. israelensis Cyt1Aa synergizes
Cry11Aa toxin by functioning as a membrane-bound recep-
tor. Proc Natl Acad Sci USA 2005, 102:18303-18308.
18. Poncet S, Delecluse A, Klier A, Rapoport G: Evaluation of syner-
gistic interactions among the CryIVa, CryIVb, and CryIVd
toxic components of Bacillus thuringiensis subsp. israelensis
crystals. J Invertebr Pathol 1995, 66:131-135.
19. Georghiou GP, Wirth MC: Influence of exposure to single ver-
sus multiple toxins of Bacillus thuringiensis subsp. israelensis
on development of resistance in the mosquito Culex quinque-
fasciatus (Diptera: Culicidae). Appl Environ Microbiol 1997,
63:1095-1101.
20. Griffitts JS, Aroian RV: Many roads to resistance: how inverte-
brates adapt to Bt toxins. Bioessays 2005, 27:614-624.
21. Boyer S, Tilquin M, Ravanel P: Differential sensitivity to Bacillus
thuringiensis var. israelensis and temephos in field mosquito
populations of Ochlerotatus cataphylla (Diptera: Culicidae):
toward resistance? Environ Toxicol Chem 2007, 26:157-162.
22. Paul A, Harrington LC, Zhang L, Scott JG: Insecticide resistance in
Culex pipiens from New York. J Am Mosq Control Assoc 2005,
21:305-309.
23. Saleh MS, El-Meniawi FA, Kelada NL, Zahran HM: Resistance devel-
opment in mosquito larvae Culex pipiens to the bacterial
agent Bacillus thuringiensis var. israelensis. J Appl Entomol 2003,
127:29-32.
24. Casacuberta JM, Santiago N: Plant LTR-retrotransposons and
MITEs: control of transposition and impact on the evolution
of plant genes and genomes. Gene 2003, 311:1-11.
25. Tu ZJ: Molecular and evolutionary analysis of two divergent
subfamilies of a novel miniature inverted repeat transposa-
ble element in the yellow fever mosquito, Aedes aegypti . Mol
Biol Evol 2000, 17:1313-1325.
26. Beaumont MA, Nichols RA: Evaluating loci for use in the genetic
analysis of population structure. P Roy Soc B-Biol Sci 1996,
263:1619-1626.
27. Pigott CR, Ellar DJ: Role of receptors in Bacillus thuringiensis
crystal toxin activity. Microbiol Mol Biol Rev 2007, 71:255-281.
28. Matsui M, Fowler JH, Walling LL: Leucine aminopeptidases:
diversity in structure and function. Biol Chem 2006,
387:1535-1544.
29. Shi LM, Jones WD, Jensen RV, Harris SC, Perkins RG, Goodsaid FM,
Guo L, Croner LJ, Boysen C, Fang H, et al.: The balance of repro-
ducibility, sensitivity, and specificity of lists of differentially
expressed genes in microarray studies. BMC Bioinformatics
2008, 9(Suppl 9):10.
30. Ferré J, Van Rie J: Biochemistry and genetics of insect resist-
ance to Bacillus thuringiensis . Annu Rev Entomol 2002, 47:501-533.
31. Huffman DL, Abrami L, Sasik R, Corbeil J, Goot FG van der, Aroian
RV: Mitogen-activated protein kinase pathways defends
against bacterial pore-forming toxins. Proc Natl Acad Sci USA
2004, 101:10995-11000.
32. Barret P, Brinkman M, Beckert M: A sequence related to rice
Pong transposable element displays transcriptional activa-
tion by in vitro culture and reveals somaclonal variations in
maize. Genome 2006, 49:1399-1407.
33. Kikuchi K, Terauchi K, Wada M, Hirano HY: The plant MITE
mPing is mobilized in anther culture. Nature 2003, 421:167-170.
34. Benjak A, Boué S, Forneck A, Casacuberta JM: Recent amplifica-
tion and impact of MITEs on the genome of grapevine (Vitis
vinifera L.). Genome Biol Evol 2009, 2009:75-84.
35. Nene V, Wortman JR, Lawson D, Haas B, Kodira C, Tu ZJ, Loftus B,
Xi ZY, Megy K, Grabherr M, et al.: Genome sequence of Aedes
aegypti, a major arbovirus vector. Science 2007, 316:1718-1723.
36. Przeworski M: The signature of positive selection at randomly
chosen loci. Genetics 2002, 160:1179-1189.
37. Storz JF: Using genome scans of DNA polymorphism to infer
adaptive population divergence. Mol Ecol 2005, 14:671-688.
38. Nielsen R: Molecular signatures of natural selection. Annu Rev
Genet 2005, 39:197-218.
39. Akey JM, Eberle MA, Rieder MJ, Carlson CS, Shriver MD, Nickerson
DA, Kruglyak L: Population history and natural selection shape
patterns of genetic variation in 132 genes. PLoS Biol 2004,
2:1591-1599.
40. Ford MJ: Applications of selective neutrality tests to molecu-
lar ecology. Mol Ecol 2002, 11:1245-1262.
41. Otto SP: Detecting the form of selection from DNA sequence
data. Trends Genet 2000, 16:526-529.
42. Depaulis F, Mousset S, Veuille M: Power of neutrality tests to
detect bottlenecks and hitchhiking. J Mol Evol 2003, 57(Suppl
1):190-200.
43. Fu YX, Li WH: Statistical tests of neutrality of mutations.
Genetics 1993, 133:693-709.
44. Fay JC, Wu CI: Hitchhiking under positive Darwinian selec-
tion. Genetics 2000, 155:1405-1413.
45. Fay JC, Wu CI: A human population bottleneck can account
for the discordance between patterns of mitochondrial ver-
sus nuclear DNA variation. Mol Biol Evol 1999, 16:1003-1005.
46. Bel Y, Siqueira HAA, Siegfried BD, Ferré J, Escriche B: Variability in
the cadherin gene in an Ostrinia nubilalis strain selected for
Cry1Ab resistance. Insect Biochem Mol Biol 2009, 39:218-223.
47. Gahan LJ, Gould F, Heckel DG: Identification of a gene associ-
ated with Bt resistance in Heliothis virescens . Science 2001,
293:857-860.
48. Morin S, Biggs RW, Sisterson MS, Shriver L, Ellers-Kirk C, Higginson
D, Holley D, Gahan LJ, Heckel DG, Carriere Y, et al.: Three cad-
herin alleles associated with resistance to Bacillus thuringien-
sis in pink bollworm. Proc Natl Acad Sci USA 2003, 100:5004-5009.
49. Jurat-Fuentes JL, Adang MJ: Characterization of a Cry1Ac-recep-
tor alkaline phosphatase in susceptible and resistant Helio-
this virescens larvae. Eur J Biochem 2004, 271:3127-3135.
50. Kumaraswami NS, Maruyama T, Kurabe S, Kishimoto T, Mitsui T,
Hori H: Lipids of brush border membrane vesicles (BBMV)
from Plutella xylostella resistant and susceptible to Cry1Ac d-
endotoxin of Bacillus thuringiensis. Comp Biochem Physiol B-Bio-
chem Mol Biol 2001, 129:173-183.
51. Rokas A, Abbot P: Harnessing genomics for evolutionary
insights. Trends Ecol Evol 2009, 24:192-200.
52. van Orsouw NJ, Hogers RCJ, Janssen A, Yalcin F, Snoeijers S, Ver-
stege E, Schneiders H, Poel H van der, van Oeveren J, Verstegen H,
van Eijk MJT: Complexity reduction of polymorphic sequences
(CRoPS): a novel approach for large-scale polymorphism
discovery in complex genomes. PLoS ONE 2007, 2:e1172.
53. Namroud MC, Beaulieu J, Juge N, Laroche J, Bousquet J: Scanning
the genome for gene single nucleotide polymorphisms
involved in adaptive population differentiation in white
spruce. Mol Ecol 2008, 17:3599-3613.
54. Tilquin M, Paris M, Reynaud S, Despres L, Ravanel P, Geremia RA,
Gury J: Long lasting persistence of Bacillus thuringiensis subsp.
israelensis (Bti) in mosquito natural habitats. PLoS ONE 2008,
3:e3432.
55. Bonin A, Paris M, Després L, Tetreau G, David JP, Kilian A: A MITE-
based genotyping method to reveal hundreds of DNA poly-
morphisms in an animal genome after a few generations of
artificial selection. BMC Genomics 2008, 9:459.
56. Raymond M, Prato G, Ratsira D: Probability analysis of mortality
assays displaying quantal response, version 3.3. Praxeme,
Saint-Georges d'Orques, France; 1995.
57. Wenzl P, Carling J, Kudrna D, Jaccoud D, Huttner E, Kleinhofs A, Kil-
ian A: Diversity Arrays Technology (DArT) for whole-
genome profiling of barley. Proc Natl Acad Sci USA 2004,
101:9915-9920.
58. Zhivotovsky LA: Estimating population structure in diploids
with multilocus dominant DNA markers. Mol Ecol 1999,
8:907-913.
59. Vekemans X, Beauwens T, Lemaire M, Roldan-Ruiz I: Data from
amplified fragment length polymorphism (AFLP) markers
show indication of size homoplasy and of a relationship
between degree of homoplasy and fragment size. Mol Ecol
2002, 11:139-151.
Publish with BioMed Central and every
scientist can read your work free of charge
"BioMed Central will be the most significant development for
disseminating the results of biomedical research in our lifetime."
Sir Paul Nurse, Cancer Research UK
Your research papers will be:
available free of charge to the entire biomedical community
peer reviewed and published immediately upon acceptance
cited in PubMed and archived on PubMed Central
yours — you keep the copyright
Submit your manuscript here:
http://www.biomedcentral.com/info/publishing_adv.asp
BioMedcentral
BMC Genomics 2009, 10:551 http://www.biomedcentral.com/1471-2164/10/551
Page 11 of 11
(page number not for citation purposes)
60. Hall TA: BioEdit: a user-friendly biological sequence align-
ment editor and analysis program for Windows 95/98/NT.
Nucl Acids Symp Ser 1999, 41:95-98.
61. de Maagd RA, Bravo A, Crickmore N: How Bacillus thuringiensis
has evolved specific toxins to colonize the insect world.
Trends Genet 2001, 17:193-199.
62. Oppert B, Kramer KJ, Beeman RW, Johnson D, McGaughey WH:
Proteinase-mediated insect resistance to Bacillus thuringien-
sis toxins. J Biol Chem 1997, 272:23473-23476.
63. Fernandez LE, Aimanova KG, Gill SS, Bravo A, Soberón M: A GPI-
anchored alkaline phosphatase is a functional midgut recep-
tor of Cry11Aa toxin in Aedes aegypti larvae. Biochem J 2006,
394:77-84.
64. Abdullah MAF, Valaitis AP, Dean DH: Identification of a Bacillus
thuringiensis Cry11Ba toxin-binding aminopeptidase from
the mosquito, Anopheles quadrimaculatus. BMC Biochem 2006,
7:16.
65. Griffitts JS, Huffman DL, Whitacre JL, Barrows BD, Marroquin LD,
Muller R, Brown JR, Hennet T, Esko JD, Aroian RV: Resistance to a
bacterial toxin is mediated by removal of a conserved glyco-
sylation pathway required for toxin-host interactions. J Biol
Chem 2003, 278:45594-45602.
66. Librado P, Rozas J: DnaSP v5: A software for comprehensive
analysis of DNA polymorphism data. Bioinformatics 2009,
25:1451-2.
67. Tajima F: Statistical method for testing the neutral mutation
hypothesis by DNA polymorphism. Genetics 1989, 123:585-595.
68. Hudson RR: Generating samples under a Wright-Fisher neu-
tral model of genetic variation. Bioinformatics 2002, 18:337-338.
69. Pfaffl MW: A new mathematical model for relative quantifica-
tion in real-time RT-PCR. Nucleic Acids Res 2001, 29:e45.
70. Hudson RR: Gene genealogies and the coalescent process. Oxf
Surv Evol Biol 1990, 7:1-44.

Supplementary resources (8)

... 2,[9][10][11][12][13][14][15] Regulatory changes that decrease the abundance of receptor proteins by reducing transcription of the genes encoding them are also associated with resistance to Cry toxins, but have received less attention. [16][17][18][19][20][21][22][23][24][25][26][27] Here we report the first evidence of reduced expression of a receptor protein associated with resistance to a Bt toxin in the pink bollworm, Pectinophora gossypiella (Saunders), one of the world's most destructive pests of cotton. 28 Pink bollworm is currently targeted by transgenic cotton producing either Bt toxin Cry1Ac alone in China or Cry1Ac + Cry2Ab toxins in India and elsewhere. ...
... Although this study provides the first evidence of pink bollworm resistance to a Bt toxin associated with down-regulation of receptor expression, this phenomenon has been reported from several other major pests including Aedes aegypti, Diatraea saccharalis, Helicoverpa armigera, Ostrinia furnacalis, Plutella xylostella, Spodoptera exigua, and Trichoplusia ni. [16][17][18][19][20][21][22][23][24][25][26][27] Four of these examples showed that downregulation of cadherin was associated with Cry toxin resistance, 17,18,22,24 but only Yang et al. 22 showed no differences in cadherin cDNA sequence between a Cry1Abresistant and susceptible strain of D. saccharalis, implicating either cis-or trans-regulation of the transcript. Liu et al. 58 also reported that Cry1Ac resistance in a laboratory-selected strain of H. armigera involves cis-mutations in the promoter region of a trypsin gene (HaTryR) that down-regulates HaTryR transcription. ...
... Although this study provides the first evidence of pink bollworm resistance to a Bt toxin associated with down-regulation of receptor expression, this phenomenon has been reported from several other major pests including Aedes aegypti, Diatraea saccharalis, Helicoverpa armigera, Ostrinia furnacalis, Plutella xylostella, Spodoptera exigua, and Trichoplusia ni. [16][17][18][19][20][21][22][23][24][25][26][27] Four of these examples showed that downregulation of cadherin was associated with Cry toxin resistance, 17,18,22,24 but only Yang et al. 22 showed no differences in cadherin cDNA sequence between a Cry1Abresistant and susceptible strain of D. saccharalis, implicating either cis-or trans-regulation of the transcript. Liu et al. 58 also reported that Cry1Ac resistance in a laboratory-selected strain of H. armigera involves cis-mutations in the promoter region of a trypsin gene (HaTryR) that down-regulates HaTryR transcription. ...
Article
Full-text available
BACKGROUND Better understanding of the molecular basis of resistance is needed to improve management of pest resistance to transgenic crops that produce insecticidal proteins from Bacillus thuringiensis (Bt). Here we analyzed resistance of the pink bollworm (Pectinophora gossypiella) to Bt toxin Cry1Ac, which is used widely in transgenic Bt cotton. Field‐evolved practical resistance of pink bollworm to Cry1Ac is widespread in India, but not in China or the United States. Previous work with laboratory‐ and field‐selected pink bollworm indicated that resistance to Cry1Ac is caused by changes in the amino acid sequence of a midgut cadherin protein (PgCad1) that binds Cry1Ac in susceptible larvae. RESULTS Relative to a susceptible strain, the laboratory‐selected APHIS‐R strain had 530‐fold resistance to Cry1Ac with autosomal recessive inheritance. Unlike previous results, resistance in this strain was not consistently associated with insertions or deletions in the expected amino acid sequence of PgCad1. However, this resistance was associated with 79‐ to 190‐fold reduced transcription of the PgCad1 gene and markedly lower abundance of PgCad1 protein. CONCLUSION The ability of pink bollworm and other major pests to evolve resistance to Bt toxins via both qualitative and quantitative changes in receptor proteins demonstrates their remarkable adaptability and presents challenges for monitoring and managing resistance to Bt crops. © 2019 Society of Chemical Industry
... To date, one cadherin was identified as Cry4B receptor and one APN as Cry11B receptor in Anopheles Abdullah et al., 2006;Zhang et al., 2008), and one cadherin as Cry11B receptor, one APN as Cry11A receptor and one other ALP as Cry4B receptor in Aedes (Chen et al., 2009a, Fernandez-Luna et al., 2006, Bayyareddy et al., 2009. Another cadherin was found to be down-regulated and to exhibit genomic signature of selection in a Bti resistant Aedes aegypti strain, suggesting its implication in resistance, but there is not so far validation of its role as a receptor for Cry toxin (Bonin et al., 2009). Zhang et al. (2005) suggested that cell death following Cry toxins binding to membrane receptors is a more complex cellular response than the simple osmotic lysis previously assumed. ...
... The size and disposition of these refuges have been widely modeled but only in situations involving a single, bi-allelic resistance locus (e.g., Tabashnik & Croft, 1982;Caprio, 2001). In the case of resistance to Bti, a mixture of toxins with various modes of action, the resistance is likely to involve many unlinked loci with various levels of dominance (Bonin et al., 2009;Paris et al., 2010). Further characterization of mechanisms underlying resistance to Bti is essential in order to develop an effective resistance management of field populations. ...
... To date, one cadherin was identified as Cry4B receptor and one APN as Cry11B receptor in Anopheles Abdullah et al., 2006;Zhang et al., 2008), and one cadherin as Cry11B receptor, one APN as Cry11A receptor and one other ALP as Cry4B receptor in Aedes (Chen et al., 2009a, Fernandez-Luna et al., 2006, Bayyareddy et al., 2009. Another cadherin was found to be down-regulated and to exhibit genomic signature of selection in a Bti resistant Aedes aegypti strain, suggesting its implication in resistance, but there is not so far validation of its role as a receptor for Cry toxin (Bonin et al., 2009). Zhang et al. (2005) suggested that cell death following Cry toxins binding to membrane receptors is a more complex cellular response than the simple osmotic lysis previously assumed. ...
... The size and disposition of these refuges have been widely modeled but only in situations involving a single, bi-allelic resistance locus (e.g., Tabashnik & Croft, 1982;Caprio, 2001). In the case of resistance to Bti, a mixture of toxins with various modes of action, the resistance is likely to involve many unlinked loci with various levels of dominance (Bonin et al., 2009;Paris et al., 2010). Further characterization of mechanisms underlying resistance to Bti is essential in order to develop an effective resistance management of field populations. ...
... Yellow fever mosquito (Aedes aegypti) has high-frequency mutations in the gene encoding the cadherin receptor, which promotes its resistance against the Bt subsp. israelensis (Bti) strain (Bonin et al., 2009). In sweet potato weevil (Cylas puncticollis), the mutant site that binds with the three toxins Cry3Bb, Cry3Ca, and Cry7Aa in the BBMV can promote the generation of cross-resistance (Hernández-Martínez et al., 2014). ...
Article
Full-text available
In this article, we review the latest works on the insecticidal mechanisms of Bacillus thuringiensis Cry toxins and the resistance mechanisms of insects against Cry toxins. Currently, there are two models of insecticidal mechanisms for Cry toxins, namely, the sequential binding model and the signaling pathway model. In the sequential binding model, Cry toxins are activated to bind to their cognate receptors in the mid-intestinal epithelial cell membrane, such as the glycophosphatidylinositol (GPI)-anchored aminopeptidases-N (APNs), alkaline phosphatases (ALPs), cadherins, and ABC transporters, to form pores that elicit cell lysis, while in the signaling pathway model, the activated Cry toxins first bind to the cadherin receptor, triggering an extensive cell signaling cascade to induce cell apoptosis. However, these two models cannot seem to fully describe the complexity of the insecticidal process of Cry toxins, and new models are required. Regarding the resistance mechanism against Cry toxins, the main method insects employed is to reduce the effective binding of Cry toxins to their cognate cell membrane receptors by gene mutations, or to reduce the expression levels of the corresponding receptors by trans-regulation. Moreover, the epigenetic mechanisms, host intestinal microbiota, and detoxification enzymes also play significant roles in the insects’ resistance against Cry toxins. Today, high-throughput sequencing technologies like transcriptomics, proteomics, and metagenomics are powerful weapons for studying the insecticidal mechanisms of Cry toxins and the resistance mechanisms of insects. We believe that this review shall shed some light on the interactions between Cry toxins and insects, which can further facilitate the development and utilization of Cry toxins.
... This result was consistent with earlier studies. For instances, downregulation of Cad has been reported in lepidopteran species including Helicoverpa armigera (Hübner) (Wang et al. 2005), Diatraea saccharalis (Fabricius) (Yang et al. 2011), Ostrinia furnacalis (Guenée) (Jin et al. 2014), Pectinophora gossypiella (Saunders) (Fabrick et al. 2019), and also in dipteran species as Aedes aegypti (Linnaeus) (Bonin et al. 2009). Literature proposed that downregulation of Cad was associated with Cry toxin resistance. ...
Article
Full-text available
Background Bacillus thuringiensis ( Bt ) utilization as a biological control agent is highly recommended due to its safety, specificity, and efficiency. Importance of the entomocidal Cry proteins secreted by Bt is dramatically increased subsequent Cry genes transformation into a number of economic crops, rendering them protection against insect attack. In the last decade, insect resistance against transgenic Bt crops is gradually raised in several lepidopteran pests. A better understanding of the processing of Bt Cry1C toxin in the larval midgut of the lepidopteran pest species, the cotton leaf worm, Spodoptera littoralis (Boisd.), is very important to characterize the main regulatory elements of Bt tolerance. Results The present study aimed to define factors that are involved in insect tolerance toward Bt Cry1C through evaluating the mRNA level of trypsin (Try), aminopeptidase N (APN), alkaline phosphatase (ALP), cadherin (Cad), and cytochrome P450 (CYP) in both susceptible and cry1C tolerant strains of S. littoralis . Total RNAs were extracted from susceptible and tolerant strains to construct cDNAs. Quantitative real-time polymerase chain reaction (qPCR) showed a significant upregulation of CYP gene in tolerant strain. In contrast, the levels of expression of Try, ALP, and Cad were significantly downregulated in tolerant strain. APN relative mRNA expression did not show significant differences between susceptible and tolerant strains. Histologically, the midgut of late third-instar larvae of tolerant population S. littoralis showed vacuolization of the epithelium and disruption of both the peritrophic membrane and the striated boarder compared to the susceptible strain. Conclusions Obtained data indicated a relationship between exposing to Bt Cry1C toxin and alteration of CYP, Try, ALP, and Cad expression in midgut of S. littoralis . These results may be an evidence for the important roles of CYP, Try, ALP, and Cad in the resistance development and toxicity to Bt Cry1C . The results are useful for further illustrating of Bt Cry1C processing and S. littoralis tolerance.
... These results confirm outcome of Mantel test in which there was no correlation between genetic and the elevation as well as genetic and geographic distances. However, this separation may be related to the other factors like intensive pest management practices, which can create high inter-population variation (Bonin et al. 2009;Piiroinen et al. 2013). Low intrapopulation and high inter-population variation is a genetic criterion that can explain diversification and fast evolution of the populations under selection pressure (Berteaux et al. 2004) because a high selection pressure driven by intensive agricultural practices in an agroecosystem, especially frequent application of chemical pesticides, can result in local adaptation and formation of genetically diverged populations with low intra-population and relatively high inter-population variation. ...
Article
Ommatissus lybicus de Bergevin (Hemiptera: Tropiduchidae) is a key pest of date palm ( Phoenix dactylifera Linnaeus; Arecaceae) with worldwide distribution and various management strategies. To study genetic diversity of date palm hopper, a series of experiments was conducted on genetic structure and genetic diversity of 15 geographic populations of O . lybicus (Abu Musa, Bam, Bushehr, Behbahan, Tezerj, Fin, Jiroft, Shahdad, Jahrom, Ghire Karzin, Ghasre Shirin, Iran; Pakistan; Oman; Egypt; and Tunisia) by amplified fragment length polymorphism, cytochrome c oxidase subunit I (COI), and 28S rRNA markers. Analysis of molecular variance analysis of amplified fragment length polymorphism data and COI sequences revealed a significant variation among O . lybicus populations (94.12% and 65.08% similarities for amplified fragment length polymorphism and COI, respectively). The 28S rDNA sequences from different populations were identical. Phylogenetic network inferred from amplified fragment length polymorphism data and COI sequences grouped two geographically close populations (Tezerj and Bam) in the two distinct clades while far apart geographical populations bunched in the same or close clades. These two populations experience repeated exposure to heavy pesticide applications annually. In conclusion, study of the genetic structure revealed a considerable variation between O . lybicus populations under intensive chemical strategies.
... In addition, such markers represent an attractive alternative to other methods of diversity reduction such as RAD sequencing (Miller, Dunham, Amores, Cresko, & Johnson, 2007) that could be less efficient in species with a high TE load (Davey et al., 2013) and did not produce satisfying results in Ae. albopictus (Goubert, Minard, Vieira, & Boulesteix, 2016). In mosquitoes, TEs have been shown to be powerful markers for both population structure analysis (Biedler et al., 2003;Boulesteix et al., 2007;Esnault et al., 2008;Santolamazza et al., 2008) and genome scans (Bonin, Paris, Tetreau, David, & Despr es, 2009). ...
Article
Invasive species represent unique opportunities to evaluate the role of local adaptation during colonization of new environments. Among these species, the Asian tiger mosquito, Aedes albopictus, is a threatening vector of several human viral diseases, including dengue and chikungunya, and raises concerns about the Zika fever. Its broad presence in both temperate and tropical environments has been considered the reflection of great “ecological plasticity”. However, no study has been conducted to assess the role of adaptive evolution in the ecological success of Ae. albopictus at the molecular level. In the present study, we performed a genomic scan to search for potential signatures of selection leading to local adaptation in one-hundred-forty field-collected mosquitoes from native populations of Vietnam and temperate invasive populations of Europe. High-throughput genotyping of transposable element insertions led to the discovery of more than 120 000 polymorphic loci, which, in their great majority, revealed a virtual absence of structure between the bio-geographic areas. Nevertheless, 92 outlier loci showed a high level of differentiation between temperate and tropical populations. The majority of these loci segregates at high insertion frequencies among European populations, indicating that this pattern could have been caused by recent adaptive evolution events in temperate areas. An analysis of the overlapping and neighboring genes highlighted several candidates, including diapause, lipid and juvenile hormone pathways.
Preprint
Full-text available
Invasive species represent unique opportunities to evaluate the role of local adaptation during colonization of new environments. Among these species, the Asian tiger mosquito, Aedes albopictus , is a threatening vector of several human viral diseases, including dengue and chikungunya, and raises concerns about the Zika fever. Its broad presence in both temperate and tropical environments has been considered the reflection of great “ecological plasticity”. However, no study has been conducted to assess the role of adaptive evolution in the ecological success of Ae. albopictus at the molecular level. In the present study, we performed a genomic scan to search for potential signatures of selection leading to local adaptation in one-hundred-forty field-collected mosquitoes from native populations of Vietnam and temperate invasive populations of Europe. High-throughput genotyping of transposable element insertions led to the discovery of more than 120 000 polymorphic loci, which, in their great majority, revealed a virtual absence of structure between the bio-geographic areas. Nevertheless, 92 outlier loci showed a high level of differentiation between temperate and tropical populations. The majority of these loci segregates at high insertion frequencies among European populations, indicating that this pattern could have been caused by recent adaptive evolution events in temperate areas. An analysis of the overlapping and neighboring genes highlighted several candidates, including diapause, lipid and juvenile hormone pathways.
Thesis
Full-text available
L’étude des modifications phénotypiques et génomiques associées à l’adaptation des pathogènes aux résistances est une étape fondamentale pour mieux comprendre et anticiper le phénomène de contournement des résistances. Le nématode à kyste Globodera pallida est un important pathogène de la pomme de terre, vis-à-vis duquel un QTL majeur de résistance, GpaVvrn, a été identifié chez Solanum vernei. Cependant, la capacité des populations de G. pallida à s’adapter à cette résistance en quelques générations seulement a été mise en évidence par évolution expérimentale. Dans ce contexte, ce travail de thèse avait pour objectifs (1) d’étudier les traits d’histoire de vie du nématode impactés par l’adaptation, afin de tester l’existence éventuelle d’un coût de virulence, et (2) d’identifier les régions génomiques impliquées dans l’adaptation, par une approche originale combinant évolution expérimentale et scans génomiques sur des lignées virulentes et avirulentes. Contre toute attente, nous avons montré que l’adaptation à la résistance issue de S. vernei entraînait une augmentation de la fitness des individus virulents sur hôte sensible. Nous avons également pu identifier des régions génomiques candidates à l’adaptation à la résistance de la plante hôte, contenant des gènes codant pour des effecteurs, et notamment des SPRYSECs, connus chez les nématodes à kyste pour être impliqués dans la suppression des défenses des plantes mais aussi dans la virulence du nématode. À terme, ces résultats s’avéreront utiles pour la conception de stratégies durables de déploiement de variétés de pommes de terre résistantes.
Article
This chapter focuses on recent advances in the study of insect transposable elements (TEs). A brief introduction on TE classification and transposition mechanisms is given. It also describes the current approaches to studying insect TEs and highlights the impact and the evolutionary dynamics of TEs in insect genomes. It discusses the applications of TEs in genetic and molecular analysis of insects. TEs have the ability to replicate and spread in the genome as primarily "selfish" genetic units. TE insertion and recombination are major sources of potentially detrimental mutations, and the host genomes have evolved sophisticated mechanisms to control TE activity. The same insertional or recombinatory activities by TEs generate a great deal of genetic and genomic plasticity and provide the raw material for adaptive evolution. TEs have been used as tools to genetically manipulate cells/organisms, taking advantage of their ability to integrate cognate DNA in the genome.
Article
Full-text available
Cytolytic pore-forming toxins are important for the virulence of many disease-causing bacteria. How target cells molecularly respond to these toxins and whether or not they can mount a defense are poorly understood. By using microarrays, we demonstrate that the nematode Caenorhabditis elegans responds robustly to Cry5B, a member of the pore-forming Crystal toxin family made by Bacillus thuringiensis. This genomic response is distinct from that seen with a different stressor, the heavy metal cadmium. A p38 mitogen-activated protein kinase (MAPK) kinase and a c-Jun N-terminal-like MAPK are both transcriptionally up-regulated by Cry5B. Moreover, both MAPK pathways are functionally important because elimination of either leads to animals that are (i) hypersensitive to a low, chronic dose of toxin and (ii) hypersensitive to a high, brief dose of toxin such that the animal might naturally encounter in the wild. These results extend to mammalian cells because inhibition of p38 results in the hypersensitivity of baby hamster kidney cells to aerolysin, a pore-forming toxin that targets humans. Furthermore, we identify two downstream transcriptional targets of the p38 MAPK pathway, ttm-1 and ttm-2, that are required for defense against Cry5B. Our data demonstrate that cells defend against pore-forming toxins by means of conserved MAPK pathways.
Article
Full-text available
Population genomics has the potential to improve studies of evolutionary genetics, molecular ecology and conservation biology, by facilitating the identification of adaptive molecular variation and by improving the estimation of important parameters such as population size, migration rates and phylogenetic relationships. There has been much excitement in the recent literature about the identification of adaptive molecular variation using the population-genomic approach. However, the most useful contribution of the genomics model to population genetics will be improving inferences about population demography and evolutionary history.
Article
Positive selection can be inferred from its effect on linked neutral variation. In the restrictive case when there is no recombination, all linked variation is removed. If recombination is present but rare, both deterministic and stochastic models of positive selection show that linked variation hitchhikes to either low or high frequencies. While the frequency distribution of variation can be influenced by a number of evolutionary processes, an excess of derived variants at high frequency is a unique pattern produced by hitchhiking (derived refers to the nonancestral state as determined from an outgroup). We adopt a statistic, H, to measure an excess of high compared to intermediate frequency variants. Only a few high-frequency variants are needed to detect hitchhiking since not many are expected under neutrality. This is of particular utility in regions of low recombination where there is not much variation and in regions of normal or high recombination, where the hitchhiking effect can be limited to a small (<1 kb) region. Application of the H test to published surveys of Drosophila variation reveals an excess of high frequency variants that are likely to have been influenced by positive selection.
Article
Elucidating the genetic basis of adaptive population divergence is a goal of central importance in evolutionary biology. In principle, it should be possible to identify chromosomal regions involved in adaptive divergence by screening genome-wide patterns of DNA polymorphism to detect the locus-specific signature of positive directional selection. In the case of spatially separated populations that inhabit different environments or sympatric populations that exploit different ecological niches, it is possible to identify loci that underlie divergently selected traits by comparing relative levels of differentiation among large numbers of unlinked markers. In this review I first address the question of whether diversifying selection on polygenic traits can be expected to produce predictable patterns of allelic variation at the underlying quantitative trait loci (QTL), and whether the locus-specific effects of selection can be reliably detected against the genome-wide backdrop of stochastic variability. I then review different approaches that have been developed to identify loci involved in adaptive population divergence and I discuss the relative merits of model-based approaches that rely on assumptions about population structure vs. model-free approaches that are based on empirical distributions of summary statistics. Finally, I consider the evolutionary and functional insights that might be gained by conducting genome scans for loci involved in adaptive population divergence.
Article
Use of the real-time polymerase chain reaction (PCR) to amplify cDNA products reverse transcribed from mRNA is on the way to becoming a routine tool in molecular biology to study low abundance gene expression. Real-time PCR is easy to perform, provides the necessary accuracy and produces reliable as well as rapid quantification results. But accurate quantification of nucleic acids requires a reproducible methodology and an adequate mathematical model for data analysis. This study enters into the particular topics of the relative quantification in real-time RT-PCR of a target gene transcript in comparison to a reference gene transcript. Therefore, a new mathematical model is presented. The relative expression ratio is calculated only from the real-time PCR efficiencies and the crossing point deviation of an unknown sample versus a control. This model needs no calibration curve. Control levels were included in the model to standardise each reaction run with respect to RNA integrity, sample loading and inter-PCR variations. High accuracy and reproducibility (<2.5% variation) were reached in LightCycler PCR using the established mathematical model.
Article
Mosquito larvae of Culex pipiens were subjected repeatedly to selection pressure with the bacterial agent Bacillus thuringiensis var. israelensis (B.t.i.) in the laboratory. Only 2.78-fold increase in tolerance to B.t.i. was induced in C. pipiens as a result of 20 generations of selection. The tolerance of C. pipiens to B.t.i. decreased by about 58% after stopping the selection for three generations. Larval selection with B.t.i. caused a reduction in the reproductive potential of mosquito adult survivors but did not affect the adult longevity and the time of blood meal digestion ingested by female mosquitoes.