ArticlePDF Available

Contrasting Patterns of Introgression at X-Linked Loci Across the Hybrid Zone Between Subspecies of the European Rabbit (Oryctolagus cuniculus)

Authors:

Abstract and Figures

Hybrid zones provide an excellent opportunity for studying the consequences of genetic changes between closely related taxa. Here we investigate patterns of genetic variability and gene flow at four X-linked loci within and between the two subspecies of European rabbit (Oryctolagus cuniculus cuniculus and O. c. algirus). Two of these genes are located near the centromere and two are located near the telomeres. We observed a deep split in the genealogy of each gene with the root located along the deepest branch in each case, consistent with the evolution of these subspecies in allopatry. The two centromeric loci showed low levels of variability, high levels of linkage disequilibrium, and little introgression between subspecies. In contrast, the two telomeric loci showed high levels of variability, low levels of linkage disequilibrium, and considerable introgression between subspecies. These data are consistent with suppression of recombination near the centromere of the rabbit X chromosome. These observations support a view of speciation where genomic incompatibilities at different loci in the genome create localized differences in levels of gene flow between nascent species.
Content may be subject to copyright.
Geraldes et al. 2005
1
Contrasting patterns of introgression at X-linked loci across the hybrid zone between subspecies of the
European rabbit (Oryctolagus cuniculus)
A
RMANDO GERALDES
*,†,
NUNO FERRAND
*,†
M
ICHAEL W. NACHMAN
*
CIBIO, Centro de Investigação em Biodiversidade e Recursos Genéticos, Campus Agrário de Vairão,
4485-661 Vairão, Portugal
Departamento de Zoologia e Antropologia, Faculdade de Ciências da Universidade do Porto, 4099-
002 Porto, Portugal
Department of Ecology and Evolutionary Biology, University of Arizona,
Tucson, AZ, 85721
Sequence data from this article have been deposited with the GenBank Data Libraries under accession nos. DQ306315-
DQ306490
Genetics: Published Articles Ahead of Print, published on April 2, 2006 as 10.1534/genetics.105.054106
Geraldes et al. 2005
2
Running Head: Introgression at X-linked loci
Key words: X chromosome, Introgression, Nucleotide variability, Linkage disequilibrium, Speciation
Corresponding Author:
Armando Geraldes
Department of Ecology and Evolutionary Biology, Biosciences West Building, The University of
Arizona, P.O. Box 210088, Tucson, AZ 85721
Phone: (520) 626-4747 Fax: (520) 621-9190
Email: geraldes@email.arizona.edu
Geraldes et al. 2005
3
Hybrid zones provide an excellent opportunity for studying the consequences of genetic
changes between closely related taxa. Here we investigate patterns of genetic variability and gene
flow at four X-linked loci within and between the two subspecies of European rabbit (Oryctolagus
cuniculus cuniculus and O. c. algirus). Two of these genes are located near the centromere and two
are located near the telomeres. We observed a deep split in the genealogy of each gene with the root
located along the deepest branch in each case, consistent with the evolution of these subspecies in
allopatry. The two centromeric loci showed low levels of variability, high levels of linkage
disequilibrium and little introgression between subspecies. In contrast, the two telomeric loci showed
high levels of variability, low levels of linkage disequilibrium, and considerable introgression between
subspecies. These data are consistent with suppression of recombination near the centromere of the
rabbit X chromosome. These observations support a view of speciation where genomic
incompatibilities at different loci in the genome create localized differences in levels of gene flow
between nascent species.
Geraldes et al. 2005
4
A key problem in evolutionary genetics concerns the origin of reproductive isolation between
incipient species. Two important conclusions come from previous studies of the genetics of
reproductive isolation. First, barriers to gene flow often derive from incompatibilities between allelic
variants at two or more loci, i.e. epistasis (B
ATESON 1909; DOBZHANSKY 1936; MULLER 1940, 1942).
Empirical support for epistasis comes from a large body of work in Drosophila, beginning with
D
OBZHANSKY (1936). More recently, specific genes underlying reproductive isolation have been
identified, and all involve epistatic interactions (M
ALITSCHEK et al. 1995; TING et al. 1998; BARBASH
et al. 2003; PRESGRAVES et al. 2003). Second, loci contributing to reproductive isolation tend to be
overrepresented on the X chromosome in groups in which males are heterogametic (C
OYNE and ORR
1989). Evidence for the “large X effect” comes from mapping studies of hybrid sterility and hybrid
inviability (e.g. D
OBZHANSKY 1936; GRULA and TAYLOR 1980; TRUE et al. 1996; PRESGRAVES 2003;
PRESGRAVES et al. 2003; TAO et al. 2003). Moreover, Haldane’s (1922) rule (the sterility or
inviability of heterogametic hybrids) seems to be due largely to incompatibilities involving recessive
X-linked mutations (T
URELLI and ORR 1995, 2000). Finally, in a number of cases where sister species
hybridize in nature, X-linked loci introgress less than autosomal loci (e.g. H
AGEN 1990; SPERLING and
S
PENCE 1991; TUCKER et al. 1992).
The genetic basis of reproductive isolation has been studied both with laboratory crosses and in
natural hybrid zones, and both approaches have advantages and disadvantages. For example,
laboratory crosses make it possible to control the genetic background as well as the environment, and
they are repeatable. Hybrid zones offer the advantage of many generations of recombination, making
fine-scale mapping more feasible. In hybrid zones, it is possible to identify genes contributing to
isolation simply from patterns of gene flow without prior knowledge of the phenotype. Hybrid zones
Geraldes et al. 2005
5
also allow us to study species that cannot be crossed in the laboratory. Finally, hybrid zones provide a
picture of the fitness of hybrid genotypes under natural conditions.
The European rabbit (Oryctolagus cuniculus) provides an opportunity to study the genetic basis
of reproductive isolation between recently evolved taxa. This species consists of two subspecies, O. c.
algirus in the southwestern portion of the Iberian Peninsula and O. c. cuniculus in the northeast of the
Iberian Peninsula and France. These two groups diverged in allopatry during the early Pleistocene and
have subsequently come into secondary contact in central Iberia, forming a contact zone that runs in a
NW-SE direction (Figure 1) (B
RANCO et al. 2000; BRANCO et al. 2002). The two subspecies are well
differentiated with respect to mtDNA (B
RANCO et al. 2000), the Y chromosome (GERALDES et al.
2005), and some allozyme loci (F
ERRAND and BRANCO 2006).
Motivated by the large X-effect documented in other species, here we focus on four X-linked
loci to understand the nature of reproductive isolation in rabbits. Two of these loci are near the
centromere and two are near the telomeres. We address three main questions. First, what are the
levels and patterns of genetic variation at genes on the rabbit X chromosome? Second, are patterns of
variation and introgression heterogeneous among loci, and if so, do the differences correlate with the
physical location of genes on the X chromosome? Third, are the data compatible with a model of
divergence without gene flow? We surveyed nucleotide variability at four X-linked loci, in a sample
of 43 male rabbits representing both subspecies and the area of contact. All four loci showed two
divergent lineages. Despite this deep divergence, there is still evidence of gene flow between
subspecies. Patterns of gene flow and nucleotide variability were heterogeneous among loci, being
low at the centromeric loci and high at the telomeric loci. We hypothesize that the centromeric region
of the X chromosome of the European rabbit may be involved in reproductive isolation between these
two subspecies.
Geraldes et al. 2005
6
MATERIALS AND METHODS
Samples: Forty-three male European rabbits were sampled (Table 1). The samples were
divided into three groups: 20 from the northeast region of the Iberian Peninsula and from France,
corresponding to Oryctolagus cuniculus cuniculus (NE), 14 from the southwest region of the Iberian
Peninsula corresponding to O. c. algirus (SW) and nine from the contact zone (CZ) as defined by
mtDNA variation (B
RANCO et al. 2000). The geographic locations of the populations sampled are
shown in Figure 1, and collecting localities are given in Table 1. Additionally, one male Lepus
granatensis was used as an outgroup.
PCR amplification and sequencing: Genomic DNA was extracted either from blood, muscle,
or liver following S
AMBROOK and RUSSEL (2001). Introns of four X-linked loci were PCR amplified;
two are located near the centromere and two are near the telomeres (Figure 2). Amplification and
sequencing primers are listed in Supplementary Table 1. For each locus, two pairs of amplification
primers were designed. The first was based either on published rabbit sequences (Phka2) (D
AVIDSON
et al. 1992) or on conserved exonic regions between human, mouse and rat. Nested primers were then
designed specifically for the rabbit based on the first sequences obtained. Amplifications were carried
out in 50 µL volumes using Platinum Taq High Fidelity DNA Polymerase (Invitrogen, CA, USA)
following manufacturer recommendations. Cycling temperatures were as follows: an initial
denaturation step at 94º for 1min and 20 sec followed by 35 cycles of 94º for 20 sec, annealing for 20
sec and extension at 68º for 4 min. Annealing temperatures for each PCR are specified in
Supplementary Table 1. PCR products were purified using the QIAquick PCR Purification Kit
(Qiagen, CA, USA) prior to sequencing. Sequencing was carried out using an ABI 3700 automated
Geraldes et al. 2005
7
sequencer. All sequences have been deposited in GenBank under accession numbers DQ306315-
DQ306490.
Data Analyses: Sequences were inspected and concatenated using the computer program
Sequencher (Gene Codes, MI, US) and then aligned manually using the BioEdit software (H
ALL
1999). By sequencing the X chromosome in males we were able to recover haplotypes directly. The
analyses below were based on single nucleotide polymorphisms in introns only.
Basic population genetic parameters, including the number of segregating sites, number of
haplotypes, levels of nucleotide diversity,
(NEI and LI 1979), and the proportion of segregating sites,
θ (W
ATTERSON 1975), were estimated using the program DnaSP 4.00 (ROZAS et al. 2003) for the
entire data set and also for the NE, CZ and SW groups (Figure 1). Phylogenetic relationships among
alleles were estimated using the Median Joining (MJ) algorithm (B
ANDELT et al. 1999) as
implemented in Network v4.1.0.8 (http://www.fluxus-technology.com/).
We estimated divergence three ways. First, divergence between all O. cuniculus alleles and
Lepus granatensis was calculated as the average pairwise distance per nucleotide site, D
xy
(Nei 1987),
and as the number of net nucleotide substitutions per site, D
a
(NEI 1987). D
a
is defined as D
xy
- 0.5 (D
x
+ D
y
), where D
xy
is the average pairwise distance between groups and D
x
and D
y
are the average
pairwise distances within groups. Second, D
xy
and D
a
were calculated between the NE and SW groups
of O. cuniculus. Finally, to estimate the divergence time of the two subspecies of O. cuniculus,
maximum likelihood (ML) net nucleotide distances between L. granatensis and O. cuniculus, and
between the two main lineages found in O. cuniculus (see Results), were calculated using PAUP v 4.0
(S
WOFFORD, 2002). Divergence time between subspecies of O. cuniculus was calculated assuming a
divergence time between L. granatensis and O. cuniculus of 11.8 million years (MY) (M
ATTHEE et al.
2004).
Geraldes et al. 2005
8
The population recombination parameter, R (R= 3Nc for X-linked loci, where c is the
recombination rate per generation and N is the population size) between adjacent sites (H
UDSON
1987), the minimum number of recombination events, Rm (H
UDSON and KAPLAN 1985), and the
number of pairs of sites showing four gametic types were calculated using DnaSP 4.00 (R
OZAS et al.
2003). Another estimator of the population recombination parameter, γ (H
EY and WAKELEY 1997),
was calculated using the software SITES. While Hudson’s R is based on the variance of the number
of base pair differences between DNA sequences, γ is a maximum likelihood estimator developed
using a coalescent model for a sample of four DNA sequences with recombination. Linkage
Disequilibrium (LD) between pairs of polymorphic sites present at a frequency of at least 10% was
calculated within and between all loci, using the statistics D’ (L
EWONTIN 1964) and r
2
(HILL and
R
OBERTSON 1968) as implemented in DnaSP 4.00 (ROZAS et al. 2003).
Tajima’s D (T
AJIMA 1989) and Fu and Li’s D (FU and LI 1993) were calculated to test for
deviations from a neutral equilibrium frequency distribution using DnaSP 4.00 (R
OZAS et al. 2003).
Ratios of polymorphism within O. cuniculus to divergence between O. cuniculus and L. granatensis
were compared with the expectations under a neutral model using the Hudson-Kreitman-Aguadé
(HKA) test (H
UDSON et al. 1987). We performed one four-locus test and six pairwise comparisons
between loci using the HKA software (H
EY and KLIMAN 1993).
At each of the four loci we detected a deep split in the genealogy (see Results). We asked if
the observed pattern of nucleotide polymorphism is compatible with a single panmitic population, as
opposed to some form of population subdivision. If two populations have evolved in allopatry, the
basal branch of a gene genealogy may be longer than in a panmitic population. Furthermore,
mutations arising in an isolated subpopulation are unable to recombine with mutations in a different
subpopulation, resulting in higher levels of LD. W
ALL (2000) suggested two measures based on LD
Geraldes et al. 2005
9
that could be powerful indicators of population subdivision. The first, l
b
, is the number of congruent
sites, defined as the number of mutations that, on a pairwise basis, result in only two haplotypes. The
second, g
d
, is the maximum physical distance between congruent sites. Coalescent simulations of
panmixia were performed with the computer program ms (Hudson 2002). For each locus 50,000
genealogies of 43 individuals were simulated conditioned on the estimated values of θ and γ.
Additionally, for each locus, coalescent simulations were performed using two different values of the
population recombination parameter (3Nc=0.0015 and 3Nc=0.015, per site) chosen to reflect the range
of recombination rates known for other mammals (e.g. D
IETRICH et al. 1994; KONG et al. 2002;
J
ENSEN-SEAMAN et al. 2004). A computer program (GARRIGAN et al. 2005) was used to calculate l
b
and g
d
from the simulated datasets, and the distribution of the two statistics for each set of conditions
were plotted against each other. The probability of obtaining the observed values of l
b
and g
d
was
calculated as the proportion of simulated genealogies for which the values of l
d
and g
d
were greater
than the observed values.
Fst and Nm were calculated using the method of H
UDSON et al. (1992a) implemented in DnaSP
4.00 (R
OZAS et al. 2003). Genetic differentiation was also calculated using the test statistic Ks*
(H
UDSON et al. 1992b), and significance was assessed by performing 1000 permutations. To test for
significant population structure among populations and among groups of populations, analyses of
molecular variance AMOVA (E
XCOFFIER et al. 1992) between the SW and NE groups were performed
using ARLEQUIN (S
CHNEIDER et al. 2000).
One simple model of divergence is an isolation model in which two populations become
separated with no subsequent gene exchange. The HKA model (H
UDSON et al. 1987) takes this form
and further assumes that the ancestral species has a population size that is the average of the two
descendent species. More recent models relax this assumption. For example, W
AKELEY and HEY
Geraldes et al. 2005
10
(1997) proposed a model that is similar to the HKA model but includes an additional parameter, θ
A
,
the population mutation parameter for the ancestral species. While the HKA test only uses the number
of polymorphic sites and divergence, this model also incorporates the total number of polymorphic
positions in the two groups (S), the number of polymorphisms exclusive to one group (Sx
NE
and
Sx
SW
), the number of shared polymorphisms (Ss) and the number of fixed differences (Sf). We tested
the fit of our data to these two models in two different ways. First we performed pairwise
comparisons between all loci, and second we performed tests with all four loci together. The fit of our
data to the W
AKELEY and HEY model of divergence without gene flow was tested using the program
WH (WANG et al. 1997).
These models assume that there has been no gene flow between the two populations since the
initial split. In many cases this is an unrealistic assumption. H
EY and NIELSEN (2004) developed a
model of population divergence that allows for genetic drift (increasing population divergence) and
gene flow (retarding it) to act together, which they call the Isolation with Migration model. The
computer program IM is an implementation of the Markov Chain Monte Carlo method for the
analyses of genetic data under this model. We used IM to estimate the effective population size of O.
c. cuniculus and of O. c. algirus and to estimate migration rates for each locus between subspecies in
each direction. IM assumes that there is no recombination within loci. For Phka2 and Hprt1 we used
the largest region showing no evidence of recombination, following W
ON and HEY (2005). For Phka2
a portion of 687 bp containing 23 polymorphic sites was used, and for Hprt1 a region of 501 bp with
20 polymorphic sites was used. For Smcx all 20 NE and 14 SW individuals were used since the data
are free of recombination. For Msn we removed three recombinant individuals (Vau1, Rsl4 and Rsl10)
from the NE group. We assigned wide prior distributions of the parameters based on preliminary trial
runs. We ran the program under Metropolis Coupled Monte Carlo Markov Chains (MC3), using 10
chains with linear heating. We used a burn-in period of 1,000,000 steps and recorded results every 40
Geraldes et al. 2005
11
steps. To test that the chains were mixing well we ran the program with different random seed
numbers and the results were similar. We ran the program for 25,625,601 steps after the burn-in
period and recorded the results of 625,014 steps.
RESULTS
Levels and patterns of variation: We observed considerable variation at all four genes.
Polymorphic sites for each gene are shown in Figure 3, and summaries of variation are given in Table
2. The number of polymorphic sites varied from 56 at Msn to 151 at Phka2, but the number of
haplotypes observed at each gene was much more constant (from 23 at Smcx to 29 at Phka2). Levels
of polymorphism were high, both in the total sample and within each subspecies. However, nucleotide
diversity (π) at the two centromeric loci was considerably lower (0.52% at Smcx and 0.55% at Msn)
than at the two telomeric loci (0.70% at Phka2 and 1.26% at Hprt1). This contrast was even more
striking in the proportion of segregating sites, where θ at the centromeric loci was roughly half the
value seen at the telomeric loci.
We assessed the amount of LD in our sample in several ways, and all were consistent in
revealing more recombination (less LD) at the telomeric loci than at the centromeric loci (Table 2),
consistent with suppression of recombination near the centromere. The number of pairs of sites
showing all four gametic types was zero at Smcx, 77 at Msn, 390 at Phka2, and 180 at Hprt1. Rm, the
minimum number of recombination events in the history of the sample (H
UDSON and KAPLAN 1985)
was zero at Smcx, intermediate at Msn and Hprt1 (7 and 6 respectively) and highest at Phka2 (17).
Similarly, R (H
UDSON 1987) between adjacent sites was low at the two centromeric loci (Smcx and
Msn) and much higher at the telomeric loci (Phka2 and Hprt1). The values for γ (H
EY and WAKELEY
Geraldes et al. 2005
12
1997), a maximum likelihood estimator of the population recombination parameter, are concordant
with R in showing more recombination at the telomeric loci than at the centromeric loci. Thus,
although there are small differences among estimators, levels of recombination are higher at the
telomeric loci than at the centromeric loci. We assessed the significance of LD through pairwise
comparisons of polymorphic sites present at a frequency of at least 10% using a Fisher’s Exact Test.
From the 8515 pairwise comparisons performed, 2497 were significant in a two-tailed test, and 980
remained significant after a Bonferroni correction for multiple tests. Of these, 197 were between
polymorphic sites at Smcx, 219 between sites at Msn, 93 between sites at Phka2, and 130 between sites
at Hprt1. Interlocus LD was only detected between Smcx and Msn where there were 341 pairs of sites
that showed significant LD.
The distribution of allele frequencies as measured by Tajima’s D and Fu and Li’s D generally
conformed to expectations under a neutral model of molecular evolution (Table 2). For example, in
the total sample, Tajima’s D was either positive (Msn and Hprt1), very close to zero (Smcx) or
negative (Phka2), but not significantly different from zero. When the population groups were
analyzed separately, Tajima’s D was negative (except for Hprt1 in the NE and SW groups), but not
significantly so (P > 0.05 for all tests). We also tested a neutral model of molecular evolution by
comparing ratios of polymorphism within O. cuniculus (in the total sample) to divergence between O.
cuniculus and Lepus granatensis, using the HKA test. We performed one four-locus comparison as
well as six pairwise comparisons between loci, and none of these were significant (P > 0.05 for each).
Divergence and gene flow between subspecies: Fst estimates between O. c. cuniculus and O.
c. algirus are shown in Table 3. Fst was very high at the centromeric loci, Smcx (0.680) and Msn
(0.829), and one order of magnitude lower at the telomeric loci, Phka2 (0.027) and Hprt1 (0.022).
The two subspecies were significantly differentiated, using the Ks* test statistic, at all loci but Hprt1
(Table 3). Similarly, the AMOVA analyses revealed that at the two centromeric loci, Smcx and Msn,
Geraldes et al. 2005
13
most of the genetic variation is partitioned among the two subspecies (64% and 84% respectively)
while at the telomeric loci, Phka2 and Hprt1, most of the observed variation was partitioned among
populations within each subspecies (93% at both loci), and only a marginal proportion (2%) of the
variation was segregating between subspecies.
This differentiation can also be seen in the phylogeny of alleles for each gene (Figure 4). At
each locus there were two divergent groups of haplotypes, and in each case the root fell along the deep
branch separating these two groups. In this analysis, only Smcx was free of homoplasy. The locus
with the most homoplasy was Phka2. This homoplasy may be due to recombination or recurrent
mutation. Evidence for recurrent mutation comes from the observation that at Hprt1 three different
positions have three nucleotides segregating (Figure 3d). Other evidence of recurrent mutation is the
fact that the amount of homoplasy is slightly reduced if CpG sites, which are known to be
hypermutable, are excluded. For example, for Phka2, the consistency index (CI) increased from 0.778
to 0.803 when CpG sites were removed. However, much of the homoplasy is probably due to
recombination, as evidenced by the fact that the CI was 1.0 for Smcx, 0.889 at Msn, but was 0.778 at
and 0.717 at Phka2 and Hprt1, respectively. Moreover, at Phka2 and Hprt1 one and 16 individuals,
respectively, were identified as recombinants between the two divergent lineages based on their
position on the haplotype network and by visual inspection of the table of polymorphism (Figure 3 a
and d).
The degree of introgression between subspecies can also be seen by the concordance (or lack
thereof) between geography and phylogeny. For the two centromeric loci (Smcx and Msn) there was
good concordance between phylogeny and geography, i.e. the two major lineages correspond well
with each subspecies (Figure 4). At Msn we did not detect any introgressed haplotypes and at Smcx
we only observed three NE individuals with haplotypes from the lineage that is otherwise restricted to
the SW and CZ groups. At the two telomeric genes (Phka2 and Hprt1), in contrast, there seems to be
Geraldes et al. 2005
14
little or no concordance between phylogeny and geography. At all four genes, individuals from the
CZ group are scattered throughout the haplotype networks.
The proportion of congruent sites, l
b
, is greater at the two centromeric loci (representing 30%
and 32% of all polymorphic sites at Smcx and at Msn respectively), than at the telomeric loci (11% at
Phka2 and 16% at Hprt1). Similarly, the maximum distance between congruent sites, g
d
, is greater at
Smcx (85% of the total locus length) and at Msn (95%) than at Phka2 (65%) and Hprt1 (16%). We
calculated the probability of observing these values of l
b
and g
d
using coalescent simulations of 50,000
genealogies of 43 individuals evolving neutrally under panmixia with mutation (θ) and recombination
(γ) parameters estimated from the data. Results are shown in Table 4. Under these conditions, the null
model was rejected for Msn (P=0.00214). This test is quite conservative using γ estimated from the
data since population subdivision will increase LD and thus underestimate the true value of
recombination. Therefore, we also conducted simulations with a population size of 10
5
and per site
recombination rates of 0.5X10
-8
and 5X10
-8
, reflecting the range of recombination rates seen in other
mammals (e.g. D
IETRICH et al. 1994; KONG et al. 2002; JENSEN-SEAMAN et al. 2004). At the two
centromeric loci, Msn and Smcx, the null model was rejected using either value of recombination. For
Phka2 the null hypothesis was rejected only with the higher recombination rate, and Hprt1 was
marginally significant (P=0.066) only for the higher recombination rate (Table 4).
Another way of looking at divergence is to quantify the amount of shared and fixed variation
between the two groups (Table 5). The number of shared polymorphisms was low at the centromeric
loci (16% and 6% of all polymorphisms at Smcx and Msn respectively), and high at the telomeric loci
(42% and 64% at Phka2 and Hprt1 respectively). Only Msn showed fixed differences between the
two groups. These patterns of variation suggest that there has been gene flow between O. c. cuniculus
and O. c. algirus at some, but not all, loci. To further test this, we performed an HKA test between
Geraldes et al. 2005
15
NE and SW population groups. A multilocus test between all four loci failed to reject the null model,
but in one pairwise comparison (between Phka2 and Msn) the model was rejected (P=0.035). This
result seems to be mainly driven by the fact that divergence at Msn was much higher than expected
(observed D=26.80, expected D=10.68). The isolation without migration model (W
AKELEY and HEY,
1997) shares most of the assumptions with the HKA model, but estimates the ancestral population size
instead of assuming that it is the average of the population size of the extant populations. A four locus
test using this model also failed to reject the null hypothesis. We also tested the fit of the data using
all pairwise comparisons to Msn. Only the comparisons to Msn were performed because in the other
comparisons there are no fixed differences and the program is unable to simulate the distribution of the
expected values. The comparison between Smcx and Msn failed to reject the null model while the
other two comparisons did reject the null model (Phka2/Msn P
( 2)
=0.032 and P
(WH)
=0.030; Msn/Hprt
P
( 2)
=0.024 and P
(WH)
=0.009).
We used IM (H
EY and NIELSEN 2004) to obtain Maximum Likelihood Estimates (MLE) of the
effective population size for each subspecies. We also estimated migration rates for each locus in each
direction. The average estimate of the effective population size was approximately 882,000 for O. c.
algirus and 422,000 for O. c. cuniculus (Table 6). The probability distribution of the ancestral
population parameter was flat (not shown), as expected if the ancestral population existed long ago
(W
ON and HEY 2005). Similarly the probability distribution of t, the time since divergence, was flat,
but non zero (not shown). This suggests that the two subspecies were isolated in the past, but this
analysis does not provide a reliable estimate of the time of isolation. Gene flow at the telomeric loci
was higher from NE to SW than from SW to NE. For the centromeric loci, introgression of Msn is
quite low in both directions, while Smcx shows some unidirectional introgression from SW to NE.
Thus it seems that levels and patterns of gene flow are very different between centromeric and
telomeric loci.
Geraldes et al. 2005
16
We estimated divergence time between the two subspecies of O. cuniculus using a phylogenetic
approach. Assuming a divergence time of 11.8 MY (M
ATTHEE et al. 2004) between O. cuniculus and
L. granatensis, divergence time between O. c. cuniculus and O. c. algirus was estimated to be on the
order of 2-5 MY ago (Table 7).
DISCUSSION
We documented genetic variation at four X-linked loci in natural populations of the European
rabbit, O. cuniculus. At each locus, we observed a deep split in the phylogeny with the root lying
along the long internal branch. This pattern is consistent with the evolution of each subspecies in
allopatry and subsequent secondary contact. Despite this broad similarity among loci, we detected
heterogeneity among loci in terms of levels of nucleotide polymorphism, recombination, and
introgression between the two subspecies. This heterogeneity corresponds well with the physical
location of the loci on the rabbit X chromosome. The two centromeric loci had lower levels of
nucleotide polymorphism, higher levels of LD and reduced introgression in comparison to the two
telomeric loci. Although we do not have direct estimates of the frequency of crossing-over in rabbits,
these observations are consistent with suppression of recombination near the centromere, as has been
observed in other species (e.g. K
ONG et al. 2002).
Levels and patterns of variation: Across the entire sample (i.e. including both subspecies),
the average heterozygosity among all loci (π=0.76%) was high and roughly one order of magnitude
higher than heterozygosity at X-linked loci in humans (π= 0.081%; H
AMMER et al. 2004) and mice
(π=0.078%; N
ACHMAN 1997). Clearly, this high level of nucleotide variability reflects not only
nucleotide polymorphism within each subspecies but also the divergence between subspecies. One
Geraldes et al. 2005
17
gene (Msn) showed no introgression between subspecies. Levels of nucleotide polymorphism at this
gene were 0.14% for O. c. cuniculus and 0.26% for O. c. algirus, closer to values observed in humans
(H
AMMER et al. 2004) and mice (NACHMAN 1997).
We also observed variation in levels of polymorphism among loci. Interestingly, the two
centromeric loci had lower levels of π and θ than observed at the two telomeric loci, both for the entire
dataset and for each subspecies considered separately. Within each subspecies, this difference may be
explained by different levels of introgression. In other words, Smcx and Msn may be less variable
within each subspecies because they contain relatively few introgressed haplotypes, compared to
Phka2 and Hprt1. However, we also observe less variation at Smcx and Msn in the total sample. This
may be due in part to lower mutation rates at these genes. For example, divergence between
Oryctolagus and Lepus is lower at Smcx (Dxy=1.74%) and Msn (Dxy=4.08%) than at Phka2
(Dxy=6.06%) or Hprt1 (Dxy=4.43%) (Table 2). If recombination is suppressed near the centromere,
these differences in mutation rate may reflect an association between mutation and recombination (e.g.
H
ELLMANN et al. 2003). It is also possible that reduced variation at Smcx and Msn may be due partly
to the effect of either positive or negative selection at linked sites (M
AYNARD-SMITH and HAIGH 1974;
C
HARLESWORTH et al. 1993).
Indirect evidence that recombination is suppressed near the centromere comes from our
observation of increased LD at Smcx and Msn compared to Phka2 and Hprt1. Patterns of LD are
affected by many factors, including selection, mutation, recombination, and changes in population size
(e.g. A
RDLIE et al. 2002). However, in humans, there is good evidence that levels of LD are inversely
correlated with recombination rate over much of the genome (e.g. R
EICH et al. 2001; MCVEAN et al.
2004; M
YERS et al. 2005). Moreover, in many organisms, recombination is suppressed near the
centromeres, particularly in metacentric chromosomes (e.g. K
ONG et al. 2002). Thus, our observation
Geraldes et al. 2005
18
of increased LD at Smcx and Msn relative to Phka2 and Hprt1 is consistent with, but not proof of,
reduced recombination near the rabbit X centromere.
Divergence and gene flow between subspecies: RFLP surveys of mtDNA polymorphism in
the Iberian Peninsula and France have shown that O. cuniculus is composed of two deeply divergent
mtDNA lineages that are thought to have diverged about 2 MYA (B
IJU-DUVAL et al. 1992; BRANCO et
al. 2000). A survey of nucleotide variability at Sry also found evidence for the existence of two
divergent lineages in the Y chromosome (G
ERALDES et al. 2005). These two lineages are associated
with O. c. algirus and O. c. cuniculus (B
RANCO et al. 2000), and are thought to have evolved in
allopatry. Our X chromosome data confirm the existence of two divergent evolutionary units in O.
cuniculus, and we show that in general the data reject the evolution of the two lineages under
panmixia. The divergence time estimated from these loci is in good agreement with divergence time
estimated from mitochondrial genes, and places the origin of these two subspecies at the
Pliocene/Pleistocene boundary. We observed high levels of population differentiation at the two
centromeric loci, but not at the telomeric loci. At the centromeric loci the two divergent lineages
correspond well with the described subspecies and are broadly concordant with the patterns of
differentiation seen at the Y chromosome and at the mtDNA. The same was not observed at the two
telomeric loci, where geography and phylogeny are largely decoupled.
If two populations evolve in allopatry for a sufficiently long time and then come into secondary
contact with little or no gene flow, a high percentage of fixed differences and a small number of
shared polymorphisms are expected. In our data this is seen only at Msn where 36% of all
polymorphisms correspond to fixed differences between groups, and 6% correspond to shared
polymorphisms. At all other loci, there are no fixed differences between subspecies and the
percentage of shared polymorphisms varies from 16% at Smcx (centromeric) to 64% at Hprt1
(telomeric). This heterogeneity among loci is also reflected in the rejection of an isolation without
Geraldes et al. 2005
19
gene flow model using the HKA test between Phka2 and Msn and the rejection of the null model using
the WH test between Phka2 and Msn, and between Hprt1 and Msn.
It is noteworthy that the patterns of reduced introgression seen at Smcx and Msn, which may
experience reduced recombination, are similar to the patterns seen previously at the mtDNA (B
RANCO
et al. 2000) and the Y chromosome (Geraldes, unpublished data), genomic regions with no
recombination. In contrast, a survey of 14 allozyme loci revealed higher, but variable, levels of
introgression (φ
ct
between subspecies ranged from 0 to 0.46), comparable to the patterns observed at X
chromosome loci. Differences among loci in levels of introgression have also been documented in
other organisms. For example, genomic regions with suppressed recombination as a result of
chromosomal rearrangements introgress less than co-linear regions in comparisons between
Drosophila pseudoobscura and D. persimilis (N
OOR et al. 2001; MACHADO et al. 2002) and between
hybridizing sunflowers of the genus Helianthus (R
IESEBERG et al. 1999). Based on such observations,
N
OOR et al. (2001) and RIESEBERG (2001) have argued that chromosomal rearrangements may
promote speciation, not through underdominance directly as in traditional models (e.g. W
HITE 1978),
but by suppressing recombination and thereby extending the effects of isolation genes to linked sites.
Our finding of low levels of introgression in an area of high LD near the X chromosome centromere of
the rabbit is consistent with similar observations in fruit flies and sunflowers. Similarly, in Anopheles
mosquitoes, two (out of three) areas of reduced introgression map to centromeres (T
URNER et al.
2005).
Our observations also have some interesting parallels with studies of hybridization in the house
mice, Mus musculus and M. domesticus. In the house mouse hybrid zone in Western Europe, the Y
chromosome shows reduced introgression (V
ANLERBERGHE et al. 1986; TUCKER et al. 1992; DOD et
al. 1993), and the X chromosome shows lower levels of introgression than do the autosomes (T
UCKER
et al. 1992; D
OD et al. 1993; MUNCLINGER et al. 2002), although there is also considerable variability
Geraldes et al. 2005
20
in levels of introgression among loci on the X chromosome (P
AYSEUR et al. 2004). In a similar
fashion, we observe some X linked loci with much reduced introgression in rabbits, providing further
support for the importance of the X chromosome in reproductive isolation. Interestingly, the
differences among loci in migration estimates (Table 6), may provide some clues to the nature of
incompatibilities underlying reproductive isolation. In particular we note that estimates of the number
of migrants from NE to SW for the centromeric loci are slightly lower than in the opposite direction.
This is in agreement with the expected asymmetric behavior of young Dobzhansky-Muller interactions
(O
RR 1995) and suggests that incompatibilities may derive from interactions between the cuniculus X
chromosome and an algirus genetic background.
One ultimate goal of speciation studies is to determine the identity of genes involved in
reproductive isolation between nascent species. With the completion of the sequence of the rabbit
genome expected in the next few years, it may soon be possible to identify candidate genes for
reproductive isolation in this species. The results presented here suggest that some of these genes may
lie near the centromere of the X chromosome.
Geraldes et al. 2005
21
ACKNOWLEDGEMENTS
This work was supported by Fundação para a Ciência e a Tecnologia (SFRH/BD/4621/2001
PhD grant to AG and Research Project POCTI/BSE/40280/2001, and by a National Science
Foundation grant to MWN. We thank R. Villafuerte for help in collecting rabbit samples, J. Good and
C. Pinho for valuable discussions, D. Garrigan, T. Salcedo and M. Dean for help with ms simulations,
and J. Hey for help with IM. We would also like to thank two anonymous reviewers for suggestions on
a previous version of this manuscript.
Geraldes et al. 2005
22
LITERATURE CITED
A
RDLIE, K. G., L. KRUGLYAK and M. SEIELSTAD, 2002 PATTERNS of linkage disequilibrium in the
human genome. Nat. Rev. Genet.
3: 299-309.
B
ANDELT, H. J., P. FORSTER, and A. ROHL, 1999 Median-Joining networks for inferring intraspecific
phylogenies. Mol. Bio. Evol. 16: 37-48
B
ARBASH , D. A., D. F. SIINO, A. M. TARONE and J. ROOTE, 2003 A rapidly evolving MYB-related
protein causes species isolation in Drosophila. Proc. natl. Acad. Sci. USA
100: 5302-5307.
B
ATESON, W., 1909 Heredity and variation in modern lights, pp. 85-101 in Darwin and modern
science, edited by A. C. Seward. Cambridge Univ. Press, Cambridge, U.K.
B
IJU-DUVAL, C., H. ENNAFAA, N. DENNEBOUY, M. MONNEROT, F. MIGNOTTE et al., 1991
Mitochondrial DNA evolution in Lagomorphs: origin of systematic heteroplasmy and
organization of diversity in European rabbits. J. Mol. Evol. 33: 92-102.
B
RANCO, M., N. FERRAND and M. MONNEROT, 2000 Phylogeography of the European rabbit
(Oryctolagus cuniculus) in the Iberian Peninsula inferred from RFLP analyses of the
cytochrome b gene. Heredity 85: 307-317.
B
RANCO, M., M. MONNEROT, N. FERRAND and A. R. TEMPLETON, 2002 Postglacial dispersal of the
European rabbit (Oryctolagus cuniculus) on the Iberian Peninsula reconstructed from nested
clade and mismatch analyses of mitochondrial DNA variation. Evolution 56: 792-803.
C
HANTRY-DARMON, C., C. ROGEL-GAILLARD, M. BERTAUD, C. URIEN, M. PERROCHEAU et al., 2003
133 new gene localizations on the rabbit cytogenetic map. Cytogenet. Genome Res. 103: 192-
201.
C
HARLESWORTH, B., M. T. MORGAN and D. CHARLESWORTH, 1993 The effect of deleterious
mutations on neutral molecular variation. Genetics 134:
1289-1903.
Geraldes et al. 2005
23
C
OYNE, J. A. and H. A. ORR, 1989 Two rules of speciation, pp. 180-207 in Speciation and its
consequences, edited by D. Otte and J. Endler. Sinauer Associates, Sunderland, MA.
C
OYNE, J. A. and H. A. ORR, 2004 Speciation. Sinauer Associates, Sunderland, MA.
D
AVIDSON, J. J., T. OZCELIK, C. HAMACHER, P. J. WILLEMS, U. FRANCKE, et al., 1992 cDNA cloning
of a liver isoform of the Phosphorylase-kinase alpha-subunit and mapping of the gene to
Xp22.2-p22.1, the region of human X-linked liver glycogenosis. Proc. Natl. Acad. Sci. USA
89: 2096-2100.
D
OBZHANSKY, T. H., 1936 Studies on hybrid sterility. II. Localization of sterility factors in
Drosophila pseudoobscura hybrids. Genetics 21: 113-135.
D
IETRICH, W. F., J. C. MILLER, R. G. STEEN, M. MERCHANT, D. DAMRON et al., 1994 A genetic map
of the mouse with 4,006 single sequence length polymorphisms. Nature Genetics
D
OD, B., L. S. JERMIIN, P. BOURSOT, V. H. CHAPMAN, J. TONNES-NIELSEN et al., 1993
Counterselection on sex-chromosomes in the Mus musculus European hybrid zone. J. Evol.
Biol.
6: 529-546.
E
XCOFFIER, L., P. E. SMOUSE and J. M. QUATTRO, 1992 Analysis of molecular variance inferred from
metric distances among DNA haplotypes - application to Human mitochondrial DNA
restriction data. Genetics 131: 479-491
F
ERRAND, N. and M. BRANCO, 2006 The evolutionary history of the European rabbit (Oryctolagus
cuniculus): major patterns of population differentiation and geographic expansion inferred
from protein polymorphism, in Phylogeography of European refugia, edited by S. Weiss and
N. Ferrand. Kluwer Academics Publishers, Amsterdam. In press.
F
U, Y. X. and W. H. LI, 1993 Statistical tests of neutrality of mutations. Genetics 133: 696-709.
Geraldes et al. 2005
24
G
ARRIGAN, D., Z. MOBASHER, S. B., KINGAN, J. A. WILDER AND M. F. HAMMER, 2005 Deep
haplotype divergence and long-range linkage disequilibrium at Xp21.1 provide evidence that
humans descended from a structured ancestral population. Genetics
170: 1849-1856.
G
ERALDES, A., C. ROGEL-GAILLARD and N. FERRAND, 2005 High levels of nucleotide diversity in the
European rabbit (Oryctolagus cuniculus) SRY gene. Anim. Genet. 36: 349-351.
G
RULA, J. W. and O. R. TAYLOR, 1980 Some characteristics of hybrids derived from the sulphur
butterflies, Colias eurytheme and Colias philodice: phenotypic effects of the X chromosome.
Evolution 34: 673-687.
H
AGEN, R. H., 1990 Population structure and host use in hybridizing subspecies of Papilio glaucus
(Lepidoptera: Papilionidae). Evolution 44: 1914-1930.
H
ALDANE, J. B. S., 1922 Sex-ratio and unisexual sterility in hybrid animals. J. Genet. 12: 101-109.
H
ALL, T. A., 1999 BioEdit: a user friendly biological sequence alignment editor and analyses program
for Windows 95/98/NT. Nucl. Acids Symp. Ser. 41: 95-98.
H
AMMER, M. F., D. GARRIGAN, E. WOOD, J. A. WILDER, Z. MOBASHER et al., 2004 Heterogeneous
patterns of variation among multiple X-linked loci: the possible role of diversity-reducing
selection in Non-Africans. Genetics
167: 1841-1853.
H
AYES, H., C. ROGEL-GAILLARD, C. ZIJLSTRA, N. A. DE HAAN, C. URIEN et. al., 2002 Establishment
of an R-banded rabbit karyotype nomenclature by FISH localization of 23 chromosome-
specific genes on both G- and R-banded chromosomes. Cytogenet. Genome. Res. 98: 199-205.
H
ELLMAN, I., I. EBERSBERGER, S. E. PTAK, S. PAABO and M. PRZEWORSKI, 2003 A neutral
explanation for the correlation of diversity with recombination rates in humans. Am. J. Hum.
Genet.
72: 1527-1535.
Geraldes et al. 2005
25
H
EY, J. and R. M. KLIMAN, 1993 Population genetics and phylogenetics of DNA sequence variation at
multiple loci within the Drosophila melanogaster species complex. Mol. Bio. Evol. 10: 804-
822.
H
EY, J. and J. WAKELEY, 1997 A coalescent estimator of the population recombination rate. Genetics
145: 833-846.
H
EY, J. and R. NIELSEN, 2004 Multilocus methods for estimating population sizes, migration rates and
divergence time, with applications to the divergence of Drosophila pseudoobscura and D.
persimilis. Genetics 167: 747-760.
H
ILL, W. G. and A. ROBERTSON, 1968 Linkage disequilibrium in finite populations. Theor. Appl.
Genet. 38: 226-231.
H
UDSON, R. R., 1987 Estimating the recombination parameter of a finite population model without
selection. Genet. Res. 50: 245-250.
H
UDSON, R. R., 2002 Generating samples under a Wright-Fisher neutral model of genetic variation.
Bioinformatics
18: 337-338.
H
UDSON, R. R. and N. L. KAPLAN, 1985 Statistical properties of the number of recombination events
in the history of a sample of DNA sequences. Genetics 111: 147-164.
H
UDSON, R. R., D. D. BOOS and N. L. KAPLAN, 1992 A statistical test for detecting geographic
subdivision. Mol. Bio.Evol. 9: 138-151.
H
UDSON, R. R., M. KREITMAN and M. AGUA, 1987 A test of neutral molecular evolution based on
nucleotide data. Genetics 116: 153-159.
H
UDSON, R. R., M. SLATKIN and W. P. MADDISON, 1992 Estimation of levels of gene flow from
DNA sequence data. Genetics 132: 583-589.
J
ENSEN-SEAMAN, M. I., T. S. FUREY, B. A. PAYSEUR, Y. LU, K. M. ROSKIN et al., Comparative
recombination rates in the Rat, Mouse and Human genomes. Genome Research
14: 528-538.
Geraldes et al. 2005
26
L
EWONTIN, R. C., 1964 Interaction of selection + Linkage. I. General considerations - Heterotic
models. Genetics 49: 49-67.
K
ONG, A., D. F. GUDBJARTSSON, J. SAINZ, G. M. JONSDOTTIR, S. A. GUDJONSSON et al., 2002 A high-
resolution recombination map of the human genome. Nat. Genet.
31: 241-247.
M
ACHADO, C. A., R. M. KLIMAN, J. A. MARKERT and J. HEY, 2002 Inferring the history of speciation
from multilocus DNA sequence data: the case of Drosophila pseudoobscura and close
relatives. Mol. Bio. Evol.
19: 472-488.
M
ALITSCHEK, B., D. FORNZLER and M. SCHARTL, 1995 Melanoma formation in Xiphophorus: a
model system for the role of receptor tyrosine kinases in tumorigenesis. Bioessays 17: 1017-
1023.
M
ATTHEE , C. A., B. J. VAN VUUREN, D. BELL AND T. J. ROBINSON, 2004 A molecular supermatrix of
the rabbits and hares (Leporidae) allows for the identification of five intercontinental
exchanges during the Miocene. Syst. Biol.
53: 433-447.
M
AYNARD-SMITH, J. and J. HAIGH, 1974 The hitch-hiking effect of a favorable gene. Genet. Res. 23:
23-35.
M
CVEAN, G. A. T., S. R. MYERS, S. HUNT, P. DELOUKAS, D. R. BENTLEY and P. DONNELLY, 2004
The fine-scale structure of recombination rate variation in the human genome. Science 304:
581-584.
M
ULLER, H. J., 1940 Bearing of the Drosophila work on systematics, pp.185-268 in The new
systematics, edited by J. S. Huxley. Clarendon press, Oxford, U.K.
M
ULLER, H. J., 1942 Isolating mechanisms, evolution and temperature. Biol. Symp. 6: 71-125.
M
UNCLINGER, P., E. BOZIKOVA, M. SUGERKOVA, J. PIALEK and M. MACHOLAN, 2002 Genetic
variation in house mice (Mus, Muridae, Rodentia) from the Czech and Slovak Republics. Folia
Zool.
51: 81-92.
Geraldes et al. 2005
27
M
YERS, S., L. BOTTOLO, C. FREEMAN, G. MCVEAN and P. DONNELLY, 2005 A fine-scale map of
recombination rates and hotspots across the human genome. Science 310:
321-324.
N
ACHMAN, M. W., 1997 Patterns of DNA variability at X-linked loci in Mus domesticus. Genetics
147: 1303-1316.
N
EI, M., 1987 Molecular Evolutionary Genetics. Columbia University Press, New York.
N
EI, M. and W. H. LI, 1979 Mathematical model for studying genetic-variation in terms of restriction
endonucleases. Proc. Natl. Acad. Sci. USA 76: 5269-5273.
N
OOR, M. A. F., K. L. GRAMS, L. A. BERTUCCI and J. REILAND, 2001 Chromosomal inversions and
the reproductive isolation of species. Proc. Natl. Acad. Sci. USA.
98: 12084-12088.
O
RR, H. A., 1995 The population genetics of speciation: the evolution of hybrid incompatibilities.
Genetics
139: 1805-1813.
P
AYSEUR, B. A., J. G. KRENZ and M. W. NACHMAN, 2004 Differential patterns of introgression across
the X chromosome in a hybrid zone between two species of house mice. Evolution
58: 2064-
2078.
P
OSADA, D. and K. A. CRANDALL, 1998 MODELTEST: testing the model of DNA substitution.
Bioinformatics 14: 817–818.
P
OSADA, D. and T. R . BUCKLEY, 2004 Model selection and model averaging in phylogenetics:
advantages of Akaike information criterion and Bayesian approaches over Likelihood ratio
tests. Syst. Biol. 53: 793–808.
P
RESGRAVES, D. C., 2003 A fine-scale genetic analysis of hybrid incompatibilities in Drosophila.
Genetics 163: 955-972.
P
RESGRAVES, D. C., L. BALAGOPALAN, S. M. ABYMAYR and H. A. ORR, 2003 Adaptive evolution
drives divergence of a hybrid inviability gene between two species of Drosophila. Nature 423:
715-719.
Geraldes et al. 2005
28
R
EICH , D. E., M. CARGILL, S. BOLK, J. IRELAND, P. C. SABETI et al., 2001 Linkage disequilibrium in
the human genome. Nature
411: 199-204.
R
IESEBERG, L. H., J. WHITTON and K. GARDNER, 1999 Hybrid zones and the genetic architecture of a
barrier to gene flow between two sunflower species. Genetics 152:
713-727.
R
IESEBERG, L. H., 2001 Chromosomal rearrangements and speciation. Trends Ecol. Evol. 16: 351-
358.
R
OZAS, J., J. C. SANCHEZ-DELBARRIO, X. MESSEGUER and R. ROZAS, 2003 DnaSP, DNA
polymorphism analyses by the coalescent and other methods. Bioinformatics 19: 2496-2497.
S
AMBROOK, J. and D. W. RUSSEL, 2001 Molecular cloning: a laboratory manual. Cold Spring Harbor
Laboratory Press, Cold Spring Harbor.
S
CHNEIDER, S., D. ROESSLI and L. EXCOFFIER, 2000 Arlequin: a software program for population
genetics data analysis. Genetics and Biometry Lab, Department of Anthropology, University
of Geneva.
S
PERLING, F. A. H. and J. R. SPENCE, 1991 Structure of an asymmetric hybrid zone between two
water strider species (Hemiptera: Gerridae: Limnoporus). Evolution 45: 1370-1383.
S
WOFFORD, D. L., 2002 PAUP* ver 2.0.b10. Phylogenetic Analysis Using Parsimony and other
methods. Sinauers Associates, Sunderland, Sunderlan MA.
T
AJIMA, F., 1989 Statistical methods for testing the neutral mutation hypothesis by DNA
polymorphism. Genetics 123: 585-595.
T
AMURA, K. and M. NEI, 1993 Estimation of the number of nucleotide substitutions in the control
region of mitochondrial DNA in humans and chimpanzees. Mol. Bio. Evol.
10: 512-526.
Geraldes et al. 2005
29
T
AO, Y., S. CHEN, D. L. HARTL and C. C. LAURIE, 2003 Genetic dissection of hybrid
incompatibilities between Drosophila simulans and D. mauritiana. I. Differential accumulation
of hybrid male sterility effects on the X and autosomes. Genetics 164: 1383-1397.
T
ING, C. T., S. C. TSAUR, M. L. WU and C. I. WU, 1998 A rapidly evolving homeobox at the site of a
hybrid sterility gene. Science 282: 1501-1504.
T
RUE, J. R., B. S. WEIR and C. C. LAURIE, 1996 A genome-wide survey of hybrid incompatibility
factors by the introgression of marked segments of Drosophila mauritiana chromosomes into
Drosophila simulans. Genetics. 142: 819-837.
T
UCKER, P. K., R. D. SAGE, J. WARNER, A. C. WILSON and E. M. EICHER, 1992 Abrupt cline for sex
chromosomes in a hybrid zone between two species of mice. Evolution 46: 1146-1163.
T
URELLI, M. and H. A. ORR, 1995 The dominance theory of Haldane’s rule. Genetics 140: 389-342.
T
URELLI, M. and H. A. ORR, 2000 Dominance, epistasis and the genetic of postzygotic isolation.
Genetics 154: 1663-1679.
T
URNER, T. L., M. W. HAHN and S. V. NUZHDIN, 2005 Genomic islands of speciation in Anopheles
gambiae. PLoS Biol. 3: e285.
V
ANLERBERGHE, F., B. DOD, P. BOURSOT, M. BELLIS and F. BONHOMME, 1986 Absence of Y
chromosome introgression across the hybrid zone between Mus musculus domesticus and Mus
musculus musculus. Genet. Res.
48: 191-197.
W
ANG, R. L., J. WAKELEY and J. HEY, 1997 Gene flow and natural selection in the origin of
Drosophila pseudoobscura and close relatives. Genetics 147: 1091-1106.
W
AKELEY, J. and J. HEY, 1997 Estimating ancestral population parameters. Genetics 145: 847-855.
W
ALL, J. D., 2000 Detecting ancient admixture in Humans using sequence polymorphism data.
Genetics 154:
1271-1279.
Geraldes et al. 2005
30
W
ATTERSON, G. A., 1975 On the number of segregating sites in genetical models without
recombination. Theor. Popul. Biol. 7: 256-276.
W
HITE, M. J. D., 1978 Modes of speciation. W. H. Freeman, San Francisco.
W
ON, Y. J. and J. HEY, 2005 Divergence population genetics of chimpanzees. Mol. Bio. Evol. 22:
297-307.
W
RIGHT, S., 1951 The genetical structure of populations. Ann. Eugenics 15: 323-354.
Geraldes et al. 2005
31
Table 1 - Individuals sampled and their geographic locations
Population Sample Size Population no.
a
Group Individual ID
Versailles 1 1 NE Ver1827
Vaulx-en-Velin 1 2 NE Vau1
Carlucet 1 3 NE Cau19
Perpignan 1 4 NE Pep18
Zaragoza 2 5 NE Zrg16, Zrg20
Castelló 2 6 NE Rsl4, Rsl10
Benavente 1 7 NE Bnv3
Zamora 2 8 NE Zam1, Zam20
La Rioja 2 9 NE Lrj3, Lrj6
Madrid 1 10 NE Mdr7
Alicante 3 11 NE Alic1, Alt107, Alt120
Cartagena 1 12 NE Cat12
Cuenca 2 13 NE Cue1, Cue3
Galicia 1 14 CZ Gal25c3
Bragança 2 15 CZ Bra1, Bra13
Toledo 3 16 CZ Tol25, Tol50, Tol64
Ciudad Real 2 17 CZ Cre1, Vdm12
Las Amoladeras 1 18 CZ Amo2
Córdoba 3 19 SW Luc4, Luc9, Luc17
Sevilla 3 20 SW Pfr1, Pfr5, Pfr7
Doñana 1 21 SW Don6
Vila Real 3 22 SW Vrl1, Vrl4, Vrl7
Idanha-a-nova 1 23 SW Id85
Elvas 2 24 SW Elv3, Elv6
Geraldes et al. 2005
32
Vila Viçosa 1 25 SW VV1_1/94
a
Population numbers from Figure 1.
Geraldes et al. 2005
33
Table 2 - Levels of polymorphism, allele frequency spectrum tests of neutrality, divergence and recombination
Polymorphism
Frequency Spectrum Tests of
Neutrality
Divergence (%)
Recombination
Between NE and SW
Between O. cuniculus
and L. granatensis
n L
a
K
b
S
c
π (%) θ (%) Tajima's D Fu and Li's D Da
d
Dxy
e
Da
f
Dxy
g
γ
h
R
i
Rm
j
4gt
k
All 43 3168 29 151 0.699 1.102
-1.337 -0.995
0.008 0.659
5.719 6.059
0.0051 0.0025 17 390
NE 20 3168 14 105 0.574 0.934 -1.584 -2.106 0.0019 0.0018 10 95
CZ 9 3168 9 86 0.924 0.999 -0.385 0.121 0.0052 0.0028 3 25
Phka2
SW 14 3168 10 95 0.708 0.945 -1.109 -1.109 0.0074 0.0032 10 137
All 43 2709 23 60 0.517 0.512 0.035 -1.499 0.531 0.782 1.475 1.735 0 0.0012 0 0
NE 20 2709 13 44 0.341 0.458 -1.021 -0.647
0 0 0 0
CZ 9 2709 6 29 0.390 0.394 -0.054 0.667 0 0.0003 0 0
Smcx
SW 14 2709 8 21 0.160 0.244 -1.438 -2.172 0 0.0020 0 0
All 43 2825 25 56 0.553 0.458 0.733 -0.849 0.786 0.949 3.803 4.083 0.0023 0.0011 7 77
NE 20 2825 8 14 0.079 0.140 -1.591 -1.390
0.0003 0.0002 1 3
CZ 9 2825 8 36 0.462 0.469 -0.073 0.423 0 0.0005 0 0
Msn
SW 14 2825 11 23 0.246 0.256 -0.161 -0.649 0.0016 0.0117 2 7
All 43 1473 28 68 1.256 1.115 0.454 -0.381 0.027 1.256 3.848 4.428 0.0088 0.0086 6 180
NE 20 1473 13 47 1.129 0.899 1.025 0.978
0.0071 0.0034 2 58
CZ 9 1473 8 49 1.228 1.249 -0.087 0.134 0.0094 0.0109 4 104
Hprt1
SW 14 1473 12 50 1.328 1.132
0.760 0.642
0.0079 0.0120 4 111
a
Length of the sequence in bp.
b
Number of haplotypes.
Geraldes et al. 2005
34
c
Number of polymorphic sites.
d
Net nucleotide divergence per site (NEI 1987) between NE and SW (CZ was excluded from this analysis).
e
Average pairwise nucleotide substitutions per site (NEI 1987) between NE and SW (CZ was excluded from this
analysis).
f
Net nucleotide divergence (NEI 1987) between the haplotype found in Lepus granatensis and all O. cuniculus
sequences.
g
Average pairwise nucleotide substitutions per site (NEI 1987) between the haplotype found in Lepus granatensis
and all O. cuniculus sequences.
h
Maximum likelihood estimate of the population recombination parameter between adjacent sites (HEY and
W
AKELEY 1997).
i
Hudson’s (1987) estimator of the population recombination parameter between adjacent sites.
j
Minimum number of recombination events in the history of the sample (HUDSON and KAPLAN 1985).
k
Number of the pairs of sites that show all four gametic types.
Geraldes et al. 2005
35
Table 3 - Genetic differentiation between NE and SW groups at four X-linked loci
Fst
a
Nm
b
φ
ct
c
φ
st
d
φ
sc
e
Da(%)
f
Phka2 0.0266* 12.22 0.02 0.07 0.05 0.008
Smcx 0.6796*** 0.16 0.64 0.80 0.43 0.531
Msn 0.8286*** 0.07 0.84 0.86 0.11 0.786
Hprt1 0.0218 14.99 0.02 0.07 0.05 0.027
a
Fst was calculated using the method proposed by HUDSON, SLATKIN and MADDISON (1992).
Statistical significance for the estimation of Fst between the two groups was obtained with
the Kst* statistic (H
UDSON et al. 1992) (*P<0.05; **P<0.01; ***P<0.005).
b
Nm was calculated according to WRIGHT’s (1951) island model of population structure,
using the expression (Fst=1/(1+3Nm)) for X-linked loci.
c
φ
ct
is the fixation index for the amount of variation segregating between NE and SW groups,
calculated using the AMOVA framework (E
XCOFIER et al. 1992). For the NE group the
thirteen populations studied were pooled into seven subgroups (NE1 - populations 1, 2, 3 and
4; NE2 - population 5; NE3 - population 6; NE4 - populations 7 and 8; NE5- populations 9
and 10; NE6 - populations 11 and 12; and NE7 - population 13). For the SW group, the seven
populations studied were pooled into four subgroups (SW1 - population 19; SW2 -
populations 20 and 21; SW3 - population 22; and SW4 - populations 23, 24 and 25).
Geraldes et al. 2005
36
d
φ
st
is the fixation index for the amount of variation segregating within each subgroup,
calculated using the AMOVA framework (E
XCOFIER et al. 1992).
e
φ
sc
is the fixation index for the amount of variation segregating among subgroups within
each group, calculated using the AMOVA framework (E
XCOFIER et al. 1992).
f
Da is the net nucleotide distance per base pair between populations (NEI 1987).
Geraldes et al. 2005
37
Table 4- Probabilities of observing the number of congruent sites, l
b
, and maximum distance between
congruent sites, g
d
, under a single panmitic population.
γ estimated from data
a
γ =0.0015
b
γ =0.015
c
l
b
g
d
l
b
and
gd
l
b
g
d
l
b
and
gd
l
b
g
d
l
b
and
gd
Phka 2 0.35364 0.96288 0.34002 0.76470 0.99352 0.76296 0.04484 0.8919 0.04032
Smcx 0.39566 0.79952 0.37298 0.09558 0.33292 0.04248 0.00010 0.10748 0.00004
Msn 0.03470 0.03466 0.00214 0.06722 0.04908 0.00638 0.00006 0.01082 0.00000
Hprt 1 0.17324 0.99916 0.17302
0.62354 0.99992 0.62354
0.06626 0.99742 0.06626
Probabilities were calculated as the proportion of simulated genealogies with values of l
b
, g
d
, or both,
equal or greater than those observed in our data.
a
For Phka2 γ=0.0051 per site, for Smcx γ=0 per site, for Msn γ=0.0023 per site and for Hprt1 γ=0.0088
per site.
b
γ=3Nc, where N=1X10
5
and c=0.5X10
-8
per site.
c
γ=3Nc, where N=1X10
5
and c=5X10
-8
per site Table 4 - Shared and fixed variation between NE and SW
groups at 4 X-linked loci.
Geraldes et al. 2005
38
Table 5- Shared and fixed variation between NE and SW groups at four X-linked loci
a
S
b
Sx
NE
c
Sx
SW
d
Ss
e
Sf
f
Phka2 140 45 (64.6) 35 (56.5) 60 (41.1) 0 (10.9)
Smcx 56 35 (16.9) 12 (14.7) 9 (10.7) 0 (2.8)
Msn 53 11 (15.2) 20 (13.3) 3 (9.7) 19 (2.6)
Hprt1 59 9 (20.3) 15 (17.7) 38 (12.9) 0 (3.4)
a
The expected values under the population parameters estimated with software WH
(W
AKELEY and HEY 1997) are shown between parentheses
b
S- Number of polymorphic positions.
c
Sx
NE
- number of exclusive polymorphisms in the NE group.
d
Sx
SW
- number of exclusive polymorphisms in the SW group.
e
Ss- Number of shared polymorphisms.
f
Sf- number of fixed differences.
Geraldes et al. 2005
39
Table 6 - Maximum-Likelihood Estimates (MLE) and the 90% Highest Posterior Density (HPD) Intervals
a
of
demographic parameters
Ne
SW
Ne
ΝΕ
Nm from NE to SW
Nm from SW to NE
Phka2 Smcx Msn Hprt1
Phka2 Smcx Msn Hprt1
MLE 882,675 422,149 10.215 0.008* 0.008* 3.470 0.281 0.404 0.012 0.951
HPD 520,833-1,737,939 202,851-685,307 2.600-16.092 0.008-6.480 0.008-0.411 0.733-14.695
0.004-7.696
+
0.004-3.569 0.004-0.358 0.004-6.687
*The estimated value of Nm is at the lower limit of resolution; i. e. 0.008 corresponds to the first bin of the parameter space
surveyed for the migration parameter from SW to NE.
+
The actual interval was larger than this and could not be estimated reliably, because the likelihood surface was relatively
flat.
a
The 90% HPD intervals contain 90% of the probability density for each estimate.
Geraldes et al. 2005
40
Table 7- Uncorrected and corrected
a
net nucleotide (Da)
b
divergences between O. cuniculus and
L. granatensis and between subspecies of O. cuniculus, and estimates of divergence time (MY)
between subspecies of O. cuniculus.
O. cuniculus/L. granatensis
O. c. cuniculus/O. c. algirus
Uncorrected
Corrected
Uncorrected
Corrected
Da (%)
Da (%)
Da (%)
Divergence Time
(MY)
Da (%)
Divergence Time
(MY)
Phka2 5.719
8.483
c
1.153 2.38
1.263
c
1.76
Smcx 1.475
1.529
d
0.735 5.88
0.750
d
5.79
Msn 3.803
4.634
e
0.711 2.21
0.819
e
2.08
Hprt1 3.848
5.313
f
1.633 5.01
1.911
f
4.24
a
Appropriate models of nucleotide substitution to correct for multiple hits were selected for each
gene using MODELTEST 3.06 (P
OSADA and CRANDALL 1998) with the Akaike Information
Criterion (P
OSADA and BUCKLEY, 2004).
b
Pairwise distances (Dxy) per site were calculated using PAUP v 4.0 (SWOFFORD, 2002) with
locus specific estimated models of substitution. Net nucleotide divergences (Da) per site were
calculated as Dxy - 0.5 (Dx +Dy). Recombinant haplotypes were excluded from this analysis.
c
TAMURA-NEI model (1993) with proportion of invariable sites of 0.6578 and estimated
α parameter describing the gamma distribution of 1.0022.
d
TAMURA-NEI model (1993) with proportion of invariable sites of 0.5536.
e
Transversion Model with estimated α parameter describing the gamma distribution of 0.2204.
f
TAMURA-NEI model (1993) with proportion of invariable sites of 0.8524.
Geraldes et al. 2005
41
Figure legends
Figure 1 - Populations of European rabbit sampled and their geographic locations. Dark grey area
indicates SW populations, light grey indicates CZ populations and white indicates NE populations.
The name and number of samples from each population, and population names are specified in Table
1.
Figure 2 - Chromosomal location of the four X-linked loci used in the present study. Modified from
C
HANTRY-DARMON et al. 2003 and HAYES et al. 2002.
Figure 3- Polymorphisms for the 4 X-linked loci. a) Phka2, b) Smcx, c) Msn and d) Hprt1. The
nucleotide found in the Lepus granatensis sample was used to infer the ancestral state for each
position. Numbers in the columns indicate the position of the polymorphism in the alignment. Each
row represents one individual. A dot (.) represents the ancestral state, and letters (A, C, G, and T)
represent the derived nucleotide. The name of the haplotype defined by each sequence is shown in
parentheses. Positions polymorphic for bases other than the one found in Lepus granatensis are
indicated with an asterisk (*) and positions segregating for three different bases in O. cuniculus are
indicated with the symbol (#). Whenever a polymorphic site in O. cuniculus corresponded to a
deletion in L. granatensis we treated the most frequent base in our sample of O. cuniculus as the
ancestral state.
Figure 4- Median-Joining haplotype networks representing the phylogenetic relationships between all
the alleles found in the European rabbit. a) Phka2, b) Smcx, c) Msn and d) Hprt1. The size of the
circles is proportional to the frequency of each haplotype. The population group of the individuals
that are represented in each haplotype is denoted by black (SW), grey (CZ) and white (NE). The point
in the network from which the outgroup sequence of Lepus granatensis stems is represented by an
arrow. Haplotype IDs correspond to Figure 3.
Geraldes et al. 2005
42
Figure 1
14
1
2
3
5
4
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
25
24
Geraldes et al. 2005
43
Figure 2
S
m
cx
Msn
P
hk
a2
13
11
12
21.1
23.3
25
15
23.1
21.5
21.3
H
pr
t1
Geraldes et al. 2005
44
Figure 3 a)
1111111111111111111111111111111111111111112222222222222222222222222222222222222222222222222222222233333
11112222234444445556667777777888889999999990001111222333444455556666677788888888999990001111111122222333333444444555555556666677777788889999900011
1277902250117880155683663997778999235770333567781172355244568222403481227234123478881334869901266899011480245893477888122244691467724707811450277902704
617999086419833801724706336367307968246723491143076682967887~824965333127848462059023614722038570681880239538264191704928957937663592047959072707934732
* * *
L. granatensis ATCAGGCTGTGCAAGTGCATTCAGAGAATGCGGTAGGCTGTCCAGCGGGGTACAAGCGTCCTCCGGGGGCGACGAGGTTCCGCGACTGTGCGTTAAGGGGGCGGCGCTCGGGACTACCCGTGCAGAAGCCCACAGCTCAAGCGGCGGCGGG
Ver1827 (P1) ....C.....A.....C........A..C.....TA.....TGG.....A.G.C......T........T..T..AA.......T.......C....A..C......C.......G....C..CAG.............GA..A....C..
Vau1 (P1) ....C.....A.....C........A..C.....TA.....TGG.....A.G.C......T........T..T..AA.......T.......C....A..C......C.......G....C..CAG.............GA..A....C..
Cau19 (P1) ....C.....A.....C........A..C.....TA.....TGG.....A.G.C......T........T..T..AA.......T.......C....A..C......C.......G....C..CAG.............GA..A....C..
Pep18 (P2) ....C.....A........C.....A..C.....TA.....TGG.......G.C..T...T........T..T..AA.......T.......C.......C.......G........T..CA.C.G............GGA.......CA.
Zrg16 (P3) ..T.CA....A.......G..T...A..CA....TA.....TGG.....A.G.C....G.T.......AT..T..AA.......T...C...C.......C.AA................C..C.G..............A.......C..
Zrg20 (P4) ....CA....A.......G..T...A..CA....TA.....TGG.T.A.A.G.C....G.T.......AT..T..AA.......T.......C.G.....C..A................C.TC.G.............GA.......C..
Rsl4 (P1) ....C.....A.....C........A..C.....TA.....TGG.....A.G.C......T........T..T..AA.......T.......C....A..C......C.......G....C..CAG.............GA..A....C..
Rsl10 (P5) .C..C.....A..........T...A..CA....TA.....TGG...A.A.G.C.A..G.T...A...AT..T..AA.......T....A..C.G.....C..A................C.TC.G.....C..A....G........C..
Bnv3 (P6) ..T.CA....A....C.....T...A..CA....TA.....TGG...A.A.G.C....G.T.......AT..T..AA.....T.T.......CC......C..A..........A.....C..C.............T..A.......CA.
Zam1 (P7) .......C..................GGC.....C...G.....A...A.....G................TTAG..........TC........GA..AC...............G.T.C..T.G....T........GA....A..C..
Zam20 (P1) ....C.....A.....C........A..C.....TA.....TGG.....A.G.C......T........T..T..AA.......T.......C....A..C......C.......G....C..CAG.............GA..A....C..
Lrj3 (P8) ...GC.....A..............A..C...A.TA.....TGG.....A.G.C......T........T..TA.AA.......T.......C....A..C......C.......G....C..C.G.............GA..A...TC..
Lrj6 (P9) ..T.C.....AT...C.....T...A..CA....TA.....TGG...A.A.GGC....G.T.T.....AT..T..AA.......T.......CC......C..A................C..C...............GA.......C..
Mdr7 (P10) ...GC.....A..............A..C.....TA.....TGG.....A.G.C......T........T..TA.AA.......T.......C....AA.C......C.......G....C..C.G......T......GA..A...TC..
A
lt1 (P11) ..T.CAT...A..........T...A..CA..A.TA...ACTGG...A.ACG.C....G.T.......AT..T..AA.......T...C...C.......CGAA..T.............C..C.G.............GA.......C..
A
lt107 (P1) ....C.....A.....C........A..C.....TA.....TGG.....A.G.C......T........T..T..AA.......T.......C....A..C......C.......G....C..CAG.............GA..A....C..
A
lt120 (P12) GC..C.....A..G...........A..C.....TA.....TGG.......G.C..T...T........T..T..AA....A..T.......C.......C.......G........T..CA.C.G............GGA.......CA.
Cat12 (P1) ....C.....A.....C........A..C.....TA.....TGG.....A.G.C......T........T..T..AA.......T.......C....A..C......C.......G....C..CAG.............GA..A....C..
Cue1 (P13) .C..C....CA.G........T...A..CA....TA.....TGG...A.A.G.C....G.T.......AT..T..AA.......T...C...C.......T..AT.T.............C..C.G.............GA...T...C.A
Cue3 (P14) ....C.....A..............A..C.....TAA....TGG.....A.G.C......T........T..TA.AA.......T..A.A..C....A..C......C...A...G....C..C.G...T.........GA..A....C..
Gal25C3 (P15) .......C..................GGC.TT..C...G.....A...A.....G....T.......A...TTAG..........TC........GA...C............T..G.T.C..T.G.............GA.......C..
Bra1 (P16) ....C.....A.....C........A..C.....TA.....TGG.....A.G.C....G.T.......AT..T..AA.......T.......C....A..C......C.......G....C..CAG.............GA..A....C..
Bra13 (P1) ....C.....A.....C........A..C.....TA.....TGG.....A.G.C......T........T..T..AA.......T.......C....A..C......C.......G....C..CAG.............GA..A....C..
Tol25 (P6) ..T.CA....A....C.....T...A..CA....TA.....TGG...A.A.G.C....G.T.......AT..T..AA.....T.T.......CC......C..A..........A.....C..C.............T..A.......CA.
Tol50 (P17) ....C.....A...AC.....T...A..CA....TA.....TGG...A.A.G.C....G.T.......AT..T..AA......AT.......CC......C..A.A...........T.AC..C.G.............GA.......C..
Tol64 (P18) .......C..................GGC.....C...G.....A...A.....G................TTAG..........TC........GA..AC...............G.T.C..T.G....T........GA.......C..
Cre1 (P13) .C..C....CA.G........T...A..CA....TA.....TGG...A.A.G.C....G.T.......AT..T..AA.......T...C...C.......T..AT.T.............C..C.G.............GA...T...C.A
Vdm12 (P19) ..T.CA....A....C.....T...A..C.....TAA....TGG...A.A.G.C....G.T.T.....AT..T..AA.......T.......CC......C..A................C..C.........C.....GA.......C..
A
mo2 (P20) ..T.CA....A..........T...A..CA..A.TA...ACTGG...A.ACG.C....G.T.......AT..T..AA...T...T...C...C.......C.AA..T.............C..C.G.............GA.......C..
Luc4 (P21) ....C.....A.........C....A........TA.....TGG.....A.G.C......TC......AT..T..AA.......T.......C.......C........A..T.......C..C.G.............GA.A.....C..
Luc9 (P22) .C..C.....A......T....TA.A..C.....C..T...TGG.......G.C.................T.AG..........TC.............C.....T...A.....G.T.C..T.G.T..T........G.T......T..
Luc17 (P23) ....C.....A..............A..C.....TAA....TGG.....A.G.C......T........T..TA.AA.......T.......C....A..C......C.......G....C..C.G.............GA..A....C..
Pfr1 (P24) ....CA....A....C.....T...A..CA....TA.....TGG..AA.A.G.C....G.T.......AT..T..AA.......T.......C.......C..............G....CA.C.G.........T..GGAT......C..
Pfr5 (P25) .C..C.....A..........T...A..CA....T......TGG...A.A...C.A.AG.T..T.A...T..TA.AA.C.....T.......C.......C...................C..C.GT.T.......C..GA.......C..
Pfr7 (P26) ..T.CA....AT...C.....T...A..CA....TA.....TGG...A.A.G.C....G.T.T.....AT..T..AA.......T.......CC......C..A................C..C....T..........GA.......C..
Don6 (P25) .C..C.....A..........T...A..CA....T......TGG...A.A...C.A.AG.T..T.A...T..TA.AA.C.....T.......C.......C...................C..C.GT.T.......C..GA.......C..
Vrl1 (P27) ....C.....A..............A..C.....TAA....TGG.....A.G.C......T........T..TA.AA.......T.....T.C....A..C......C.......G....C..C.G.............GA..A....C..
Vrl4 (P27) ....C.....A..............A..C.....TAA....TGG.....A.G.C......T........T..TA.AA.......T.....T.C....A..C......C.......G....C..C.G.............GA..A....C..
Vrl7 (P18) .......C..................GGC.....C...G.....A...A.....G................TTAG..........TC........GA..AC...............G.T.C..T.G....T........GA.......C..
Id85 (P28) .C..C.....A..............A..C....GTA.....TGG.....A.G.C......T.......AT..TA..A..T....T.......C....A..C..............G....C..C.G.............GA..A....C..
Elv3 (P27) ....C.....A..............A..C.....TAA....TGG.....A.G.C......T........T..TA.AA.......T.....T.C....A..C......C.......G....C..C.G.............GA..A....C..
Elv6 (P27) ....C.....A..............A..C.....TAA....TGG.....A.G.C......T........T..TA.AA.......T.....T.C....A..C......C.......G....C..C.G.............GA..A....C..
VV1_1/94 (P29) ....C...A.A.............GA..C.....C..T...TGG.........C......T.....A..TA.T...AC......T......AC.......C...........T..........C.G.............GA.....A.C..
Geraldes et al. 2005
45
Figure 3 b)
111111111111111111111111111122222222222
12222444456666669999000000111122333344455566678901224455666
872578023630357992679044478111123478945806747992656074517577
791227657304237563981314743136867116226267558886109309012445
*
L. granatensis GTAGTGAGGAAGTGGTATCTAGGCCCGCGAGAGAGGCGCGGTACAGCAGACCGCCTTTCA
Ver1827 (S1) ...A..G.A...C..G.C.CG.AT..A.......C..C.A.CG.....A...........
Vau1 (S2) ...A..G.A...C..G.C.CG.AT..A.......C......CG.....A...........
Cau19 (S3) ...A..G.A...C..G.C.CG.AT..A.......C..C...CG.....A...........
Pep18 (S2) ...A..G.A...C..G.C.CG.AT..A.......C......CG.....A...........
Zrg16 (S4) ...A..G.....C..G.C.CG.AT..A.C.....C......................G..
Zrg20 (S5) ..C..A.....TCT.A...........T...GT............A...G.G..TC....
Rsl4 (S6) ...A..G.....C..G.C.CG.AT..A.......C.........................
Rsl10 (S7) ...A..GA....C..G.C.CG.AT..A.......C.........................
Bnv3 (S6) ...A..G.....C..G.C.CG.AT..A.......C.........................
Zam1 (S6) ...A..G.....C..G.C.CG.AT..A.......C.........................
Zam20 (S8) ...A..G..G..C..G.C.CG.AT..A.......C.........................
Lrj3 (S9) ...A..G.....C..G.CTCG.AT..A.......C...............G.........
Lrj6 (S10) ...A..GA....C..G.C.CG.AT..A..TA...C...A.....................
Mdr7 (S6) ...A..G.....C..G.C.CG.AT..A.......C.........................
Alic1 (S6) ...A..G.....C..G.C.CG.AT..A.......C.........................
Alt107 (S2) ...A..G.A...C..G.C.CG.AT..A.......C......CG.....A...........
Alt120 (S11) ...A..G.....C..G.C.CGAAT..A.......C.........................
Cat12 (S3) ...A..G.A...C..G.C.CG.AT..A.......C..C...CG.....A...........
Cue1 (S12) ..C..AG....TCT.A........T..T....T.............AG.G....TCC...
Cue3 (S13) ..C..A.....TCTAA...........T....T..........T.A...G....TC....
Gal25c3 (S14) A.C..A.....TCT.A...........T....T............A...G....TC....
Bra1 (S6) ...A..G.....C..G.C.CG.AT..A.......C.........................
Bra13 (S6) ...A..G.....C..G.C.CG.AT..A.......C.........................
Tol25 (S15) ..C..A.....T.T.AG..........T....T............A...G....TC....
Tol50 (S16) ..C..A.....TCTAA...........T....T............A...G....TC....
ToL64 (S17) ..C..AG....TCT.A...........T....T...........C..G.G....TCC...
Cre1 (S15) ..C..A.....T.T.AG..........T....T............A...G....TC....
Vdm12 (S15) ..C..A.....T.T.AG..........T....T............A...G....TC....
Amo2 (S5) ..C..A.....TCT.A...........T...GT............A...G.G..TC....
Luc4 (S5) ..C..A.....TCT.A...........T...GT............A...G.G..TC....
Luc9 (S5) ..C..A.....TCT.A...........T...GT............A...G.G..TC....
Luc17 (S18) ..C..AG....TCT.A...........T....T............A...G....TC....
Pfr1 (S18) ..C..AG....TCT.A...........T....T............A...G....TC....
Pfr5 (S16) ..C..A.....TCTAA...........T....T............A...G....TC....
Pfr7 (S5) ..C..A.....TCT.A...........T...GT............A...G.G..TC....
Don6 (S19) ..C..A.....TCT.A...........T...GT...T........A...G.G..TC....
Vrl1 (S20) ..C..A.....TCTAA...........T....T............A...G...TTC....
Vrl4 (S16) ..C..A.....TCTAA...........T....T............A...G....TC....
Vrl7 (S21) .AC..AG....TCT.A...........T....T.......A....A...G..A.TC....
ID85 (S18) ..C..AG....TCT.A...........T....T............A...G....TC....
Elv3 (S22) ..C..A....GTCT.A...........T...GTT...........A...G.G..TC..AG
Elv6 (S5) ..C..A.....TCT.A...........T...GT............A...G.G..TC....
V
V
1
_
1/94 (S23) ..C.AAG....TCT.A........TT.T....T..A..........AG.G....TCC...
Geraldes et al. 2005
46
Figure 3 c)
11111111111111111111111122222222222222
112344566677889900012234445566777778889912333455567788
67228067002568463456714872693956566778990933344505770400
42461247745510585032319375099132613195582964905593460557
L. granatensis TCCGCACCCGGCAGTTCCGGCTCTTGCTAGTCGCATTTTTACTAGGACGGTACGGC
Ver1827 (M1) C...G..T.A...A...T....T...TC.AC...GC.C.C....CT....C..A..
Vau1 (M2) C...G..T.A.TTA...T.A..T...TC.AC...GC.C.C....CT....C..A..
Cau19 (M1) C...G..T.A...A...T....T...TC.AC...GC.C.C....CT....C..A..
Pep18 (M1) C...G..T.A...A...T....T...TC.AC...GC.C.C....CT....C..A..
Zrg16 (M1) C...G..T.A...A...T....T...TC.AC...GC.C.C....CT....C..A..
Zrg20 (M1) C...G..T.A...A...T....T...TC.AC...GC.C.C....CT....C..A..
Rsl4 (M3) C...G..T.A...A...T....T....C..C..TGC.C.C..C.CT....C..A.A
Rsl10 (M4) C...G....A...A...T....T....C..C..TGC.C.C..C.CT....C..A.A
Bnv3 (M1) C...G..T.A...A...T....T...TC.AC...GC.C.C....CT....C..A..
Zam1 (M5) C...G....A...A...T....T....C..C....C.C.C....CT.T..C..A..
Zam20 (M6) C...G..T.A...A...T....T...TC.AC...GC.CCC....CT....C..A..
Lrj3 (M1) C...G..T.A...A...T....T...TC.AC...GC.C.C....CT....C..A..
Lrj6 (M1) C...G..T.A...A...T....T...TC.AC...GC.C.C....CT....C..A..
Mdr7 (M1) C...G..T.A...A...T....T...TC.AC...GC.C.C....CT....C..A..
Alic1 (M7) C...G..TGA...A...T....T....C..C...GC.C.C....CT....C..A..
Alt107 (M1) C...G..T.A...A...T....T...TC.AC...GC.C.C....CT....C..A..
Alt120 (M1) C...G..T.A...A...T....T...TC.AC...GC.C.C....CT....C..A..
Cat12 (M1) C...G..T.A...A...T....T...TC.AC...GC.C.C....CT....C..A..
Cue1 (M1) C...G..T.A...A...T....T...TC.AC...GC.C.C....CT....C..A..
Cue3 (M8) C...G..T.AC..A...T....T...TC.AC...GC.C.C....CT....C..A..
Gal25c3 (M9) .T.A.G.........CA....G........C...G.C.....CC.....T......
Bra1 (M1) C...G..T.A...A...T....T...TC.AC...GC.C.C....CT....C..A..
Bra13 (M10) .T.A.G.........CA....G......G.C...G.C.....CC.....T......
Tol25 (M11) .T.A.G.........CA....G........C...G.C.....CC............
Tol50 (M12) .....G.........CA....G.A......C...G.C.....C.............
Tol64 (M9) .T.A.G.........CA....G........C...G.C.....CC.....T......
Cre1 (M13) .....G.....T...CA..A.G.........G..G.C...................
Vdm12 (M14) .....G.........CA....G........C...G......T..............
Amo2 (M15) C.T.G..T.A...A...T....T....C..C..TGC.C.C....CT....C..A..
Luc4 (M16) .....G.....T...CA..A.G..C......G..G.C...................
Luc9 (M16) .....G.....T...CA..A.G..C......G..G.C...................
Luc17 (M12) .....G.........CA....G.A......C...G.C.....C.............
Pfr1 (M17) .....GG....T...CA..A.G.........G..G.C...................
Pfr5 (M18) ...A.G.........CA....G........C.A.G.C...T.C.....C.......
Pfr7 (M19) .....G.........CA....G...A....C...G.C.........G....G..A.
Don6 (M20) .....G........ACA....G........C...G.C.....C.............
Vrl1 (M21) .T.A.G.........CA....G........C...G.C.....CC..G..T......
Vrl4 (M22) .....G.....T...CA.AA.G.........G..G.C.........G.....T...
Vrl7 (M21) .T.A.G.........CA....G........C...G.C.....CC..G..T......
Id85 (M23) .....G.....T...CA..A.G.........G..G.C...............T...
Elv3 (M21) .T.A.G.........CA....G........C...G.C.....CC..G..T......
Elv6 (M24) .....G.........CA....G........C...G.C.....C...G.........
VV1
_
1/94 (M25) .T.A.G.........CA...TG........C...G.C.....CC..G..T......
Geraldes et al. 2005
47
Figure 3 d)
111111111111111
11122222233334444556777777777777888888888888899999000012222333344
15734703488904490224374001122334458000133345889902355146712557445601
75321245157570474459160592837281907145818901190618112406290589160621
# * # #
L. granatensis GAACTCGGATAGCCTTCATCCCTCCAGGCGCTTGGAACCGAGCCCTGACGGTTCAAACAAGGAAAGGG
Ver1827 (H1) ......A.........T.G....AA......CCAT..A.A..TA....T.....GG....A...G...
Vau1 (H1) ......A.........T.G....AA......CCAT..A.A..TA....T.....GG....A...G...
Cau19 (H2) ......A.........T......AA......CCAT..A.A..TA....T.....GG....A...G...
Pep18 (H3) ........G...............A......CCAT..A.A..TA...GT......G..GT.....A..
Zrg16 (H4) ......A...G.....T.GT......A.T..CG..C........TC.G..TGG.GG..GT.....AAA
Zrg20 (H5) ........G...............A......CCAT..A.A..TA...GTA.....G..GT.....A..
Rsl4 (H6) ........G...............A..A..TCCAT..A.A..TA...GT......G..GT.....A..
Rsl10 (H6) ........G...............A..A..TCCAT..A.A..TA...GT......G..GT.....A..
Bnv3 (H7) ........G...............A......CCAT..A.A..TA...GT......G..GT...G.A..
Zam1 (H8) ......A...G.....T..T......A.T..CG..C........TC.G..TGG.GG..GT.....AAA
Zam20 (H4) ......A...G.....T.GT......A.T..CG..C........TC.G..TGG.GG..GT.....AAA
Lrj3 (H9) .......A.C.A...........AA......CCAT..A.A..TA...GT....TGG....A...G...
Lrj6 (H4) ......A...G.....T.GT......A.T..CG..C........TC.G..TGG.GG..GT.....AAA
Mdr7 (H10) ........G.............C.A......CCAT..A.A..TA...GT......G..GT.....A..
Alic1 (H11) .......A.C..A..GT......AA....A.CCAT..A.A..TA...GT.....G.....A...G...
Alt107 (H12) .......A.C.............AA....A.CCAT..A.A..TA...GT.....G..T..A...G...
Alt120 (H1) ......A.........T.G....AA......CCAT..A.A..TA....T.....GG....A...G...
Cat12 (H13) .......A.C..............A......CCAT..A.A..TA...GT......G..GT.....A..
Cue1 (H4) ......A...G.....T.GT......A.T..CG..C........TC.G..TGG.GG..GT.....AAA
Cue3 (H9) .......A.C.A...........AA......CCAT..A.A..TA...GT....TGG....A...G...
Gal25c3 (H14) .......A.C.............AA......CCAT..A.A..TA...GTA....G.....A...G...
Bra1 (H4) ......A...G.....T.GT......A.T..CG..C........TC.G..TGG.GG..GT.....AAA
Bra13 (H15) ...T..A......T..TTG.......A.T..CG.T....T....TC.G...GG.GG....A...G...
Tol25 (H16) ......A.........T.G..T...GA.T...G.T...T.....TCAG...GG.GG..GT.....A..
Tol50 (H17) ........G...............A......CCAT..A.A...A...GT......G..GT.A...A..
Tol64 (H18) ........GC......T.G.....A......CCAT..A.AG.TA...GT......G..GT.....A..
Cre1 (H19) ......A.........T.G.T....GA.T...G.T.........TC.G...GG.GG..GT.....A..
Vdm12 (H19) ......A.........T.G.T....GA.T...G.T.........TC.G...GG.GG..GT.....A..
Amo2 (H20) ....C.....G.....T.G.......A.T...G.TC........TC.G...GG.GG..GT.....A.A
Luc4 (H21) A.C.............T.G.....A.A.T..CG.T....T.A..TC.G...GG.GG..GT.....A..
Luc9 (H22) ........G..............AA....A.CCAT..A.A..TA...GT.....G.....A...G...
Luc17 (H4) ......A...G.....T.GT......A.T..CG..C........TC.G..TGG.GG..GT.....AAA
Pfr1 (H23) .G......G...............A......CCAT..A.A..TA...GT......G..GT..T..A..
Pfr5 (H5) ........G...............A......CCAT..A.A..TA...GTA.....G..GT.....A..
Pfr7 (H24) ...T..A......T..TTG.......A.T..CG.T.T..T....TC.G...GG.GG....A...G...
Don6 (H25) .....T.A.C.A............A......CCAT..A.A..TA...GT.....GGG...A...G...
Vrl1 (H14) .......A.C.............AA....C.CCAT..A.A..TA...GTA....G.....A...G...
Vrl4 (H15) ...T..A......T..TTG.......A.T..CG.T....T....TC.G...GG.GG....A...G...
Vrl7 (H14) .......A.C.............AA....C.CCAT..A.A..TA...GTA....G.....A...G...
Id85 (H26) .............T..TTG.......A.T..CG.T....T....TC.G...GG.GG..GT.....A..
Elv3 (H27) ........G.....A.........A......CCAT..A.A..TA...GT......G..GT.....A..
Elv6 (H28) ...T.........T..TTG.......A.T..CG.T.T..T....AC.G...GG.GG....A...G...
VV1
_
1/94 (H4) ......A...G.....T.GT......A.TCGCC..GCA.CGAG...CGGC.GG.GG..GT.....
A
A
A
Geraldes et al. 2005
48
Figure 4 a)
Lepus granatensisLepus granatensis
Geraldes et al. 2005
49
Figure 4 b)
Lepus
granat ensis
Lepus
granat ensis
Geraldes et al. 2005
50
Figure 4 c)
Lepus granatensisLepus granatensis
Geraldes et al. 2005
51
Figure 4 d)
Lepus granatensisLepus granatensis
Geraldes et al. 2005
52
Supplementary Table 1- X-chromosome loci details, PCR and Sequencing Primers
a
Primer no. Primer name Type
b
Primer sequence (5'-3') Temp.
c
1 Phka2e2f Pcr CCTGGGTGCGAGACAACATCTACAG
2 Phka2e4r Pcr CCAGGAACAGGAGAAAGAGGGAGGT
66º
3 NPhka2e2f NPcr CATGGCCTACCGCAAGAATG
4 NPhka2e4r NPcr GCAGGTGGCAGTGTTGTACTT
62º
5 P730r S TTTCGAGGCTCCCTACTTCA
6 P640f S GATTCCCACCAAAGCAAAGA
7 P1900f S ACCTCCTCCCCTCAACCTTA
8 P2015r S TGTGGAAAACACAACCTGGA
9 P2630f S TAGAGGGTCACAGTCTCGGG
10 SPhka2_e2f2
S TCAGAAGCCAAGATTCCCAC
11 Sphka2_e3r S CCTCATCATGCACTGGAGAA
12 Sphka2_e3f S GAAGCTGATGCGAGGTCTTC
13 Sphka2_e3r2 S GCACATTACTTCCCGTGTCC
14 Sphkf S ACGAACACCTCACCTGCTTAT
15 Slehf S GCCATCCTCCACTGCACTC
16 SP2919r S ATCAGTGTGTTCTTTGCCCC
17 SP851f S TGTCAGGTCTAGGAAGAGT
18 LP2000r S CTCCTGCACCAGCCCGCC
19 SP1410r S TCACTGGGAGATGGCCACGT
20 Smcxe2f Pcr GGACTGGCAGCCACCCTTTGCTGTG
21 Smcxe4r Pcr CACAAGGTTGGCTCCAGACTGGTACA
64º
22 NSmcxe2f NPcr/S
GTAACAGGGTTGGGAAAACGGATACC
23 Nsmcxe4r NPcr/S
TGTTTTGCCTGGTGGATAGTTGAGG
62º
24 SSmcxf2 S GGGAAGTGAGAGTGAGTGCC
25 SS1580r S TTCACTTCCTCATGCACCAG
Geraldes et al. 2005
53
26 Ssmcxr2 S TCTCCAAATGAACCGAAACC
27 Msne4f2 Pcr CCCAGCGCCTGTTCTTTCTGCAAG
28 Msne7r3 Pcr/S CCTGATTTCACTCCAAGGGAAGCC
62º
29 M105f NPcr GGTCTGTGCCTCTCCATTAGGG
30 M3226r NPcr/S
TGGTTAGTTGCCAACACTGAGGG
62º
31 M847r S ACCCAGATCTACTACACTGT
32 M769f S TAGGGACAGTAAGCTGTGC
33 M1513f S ACAGAGATGATGCTGGCACT
34 M2099f S TATCTGAAGATTGCTCAAGA
35 M3300r S AATAGAATAGAGTAAGAGT
36 Msne4f3 S TCAACAAGGAAGTGCACAAGTGTGG
37 Msne5f1 S CACAAACTCAACAAGGACCAGTGGG
38 Msne6r1 S ATTCTGCTCATAGATGTTGAGACCC
39 Smsnf S AGTGAACTACTTCAGCATTA
40 Smsnr S AGTGCCAGCATCATCTCTGT
41 Hprt11e2f Pcr GCGATGATGAACCAGGTTATGACC
42 Hprt11e3r Pcr TAGCTCTTCAGTCTGATAAAATCTACAG
64º
43 NpHprt1_f NPcr/S
ATTACGTCGAGGACTTGGAAAGG
44 NpHprt1_r NPcr/S
AGCAGGTCAGCAAAGAACTTATAGCC
68º
45 Hprt1r2 S GGATTCTAGTCCCGGTTGCT
46 Sh654r S TCCAATGAATGTTAGAACTA
47 Sh520f S CTCTCCTCAGCTTCCTGTA
a
Primers 1 through 19 were used for the region between exons two and four of Phka2
[Phosphorylase kinase alpha 2 (liver)] located on O. cuniculus Xp15 (C
HANTRY-DARMON et al.
2003). Primers 20-26 were used for the region between exons two and four of Smcx [Smcy
homolog, X-linked (mouse)] located on O. cuniculus Xp11 (C
HANTRY-DARMON et al. 2003).
Geraldes et al. 2005
54
Primers 27 through 40 were used in the study of the region between exons six and nine of Msn
(Moesin) located on O. cuniculus Xq12prox (C
HANTRY-DARMON et al. 2003). For this gene the
number of exons in the three species compared is not the same. The primers used for the
amplifications are in exons six and nine of humans, which correspond to exons three and six in
the house mouse and exons four and seven in the rat. Finally, primers 41 through 47 were used
for the region between exons two and three of Hprt1 [Hypoxanthine phosphoribosyltransferase 1
(Lesch-Nyhan syndrome)] located on O. cuniculus Xq23 (H
AYES et al. 2002).
b
Purpose for which the primers were used: PCR- primers used for PCR, NPCR- primers used for
Nested PCR and S- primers used for most sequencing reactions, SA- alternative sequencing
primers used to sequence individuals in which the S primers did not work.
c
Annealing temperature used in the respective PCR or Nested PCR reactions.
... Such variation was directly linked with their chromosomal location. Different patterns of recombination between loci near centromeres and loci near telomeres between subspecies of the European rabbit (Oryctolagus cuniculus) were found for autosomal and sex chromosomes (Geraldes et al. 2006;Carneiro et al. 2009). It was found low levels of gene flow at two loci near the centromere and high levels at two loci near the telomere, suggesting that the centromeric region of the X chromosome may be involved in reproductive isolation between the two subspecies (Geraldes et al., 2006). ...
... Different patterns of recombination between loci near centromeres and loci near telomeres between subspecies of the European rabbit (Oryctolagus cuniculus) were found for autosomal and sex chromosomes (Geraldes et al. 2006;Carneiro et al. 2009). It was found low levels of gene flow at two loci near the centromere and high levels at two loci near the telomere, suggesting that the centromeric region of the X chromosome may be involved in reproductive isolation between the two subspecies (Geraldes et al., 2006). Similarly, analyzing autosomal loci, those near telomeres showed little differentiation between the subspecies, while most of those near centromeres showed strong differentiation, supporting the view of speciation in which regions of low recombination (near centromeres) can facilitate species divergence (Carneiro et al. 2009). ...
Article
Full-text available
Satellites are an abundant source of repetitive DNAs that play an essential role in the chromosomal organization and are tightly linked with the evolution of sex chromosomes. Among fishes, Triportheidae stands out as the only family where almost all species have a homeologous ZZ/ZW sex chromosomes system. While the Z chromosome is typically conserved, the W is always smaller, with variations in size and morphology between species. Here, we report an analysis of the satellitome of Triportheus auritus (TauSat) by integrating genomic and chromosomal data, with a special focus on the highly abundant and female-biased satDNAs. In addition, we investigated the evolutionary trajectories of the ZW sex chromosomes in the Triport-heidae family by mapping satDNAs in selected representative species of this family. The satellitome of T. auritus comprised 53 satDNA families of which 24 were also hybridized by FISH. Most satDNAs differed significantly between sexes, with 19 out of 24 being enriched on the W chromosome of T. auritus. The number of satDNAs hybridized into the W chromosomes of T. signatus and T. albus decreased to six and four, respectively, in accordance with the size of their W chromosomes. No TauSat probes produced FISH signals on the chromosomes of Agoniates halecinus. Despite its apparent conservation, our results indicate that each species differs in the satDNA accumulation on the Z chromosome. Minimum spanning trees (MSTs), generated for three satDNA families with different patterns of FISH mapping data, revealed different homogeniza-tion rates between the Z and W chromosomes. These results were linked to different levels of recombination between them. The most abundant satDNA family (TauSat01) was exclusively hybridized in the centromeres of all 52 chromosomes of T. auritus, and its putative role in the centromere evolution was also highlighted. Our results identified a high differentiation of both ZW chromosomes regarding satellites composition, highlighting their dynamic role in the sex chromosomes evolution.
... Несмотря на неотличимость кариотипов этих двух подвидов, предположительно центральная область Х-хромосомы (перицентромерные районы) вовлечена в репродуктивную изоляцию между ними (нарушения конъюгации гомологов в мейозе), что приводит к ограниченности генетических потоков между подвидами. Предковой для домашнего кролика является французская форма Oryctolagus c. cuniculus (48). Эта ограниченность настолько велика, как в Австралии, куда были также завезены представители испанского подвила (O. ...
... Несмотря на неотличимость кариотипов этих двух подвидов, предположительно центральная область Х-хромосомы (перицентромерные районы) вовлечена в репродуктивную изоляцию между ними (нарушения конъюгации гомологов в мейозе), что приводит к ограниченности генетических потоков между подвидами. Предковой для домашнего кролика является французская форма Oryctolagus c. cuniculus (48). Эта ограниченность настолько велика, как в Австралии, куда были также завезены представители испанского подвила (O. ...
... Несмотря на неотличимость кариотипов этих двух подвидов, предположительно центральная область Х-хромосомы (перицентромерные районы) вовлечена в репродуктивную изоляцию между ними (нарушения конъюгации гомологов в мейозе), что приводит к ограниченности генетических потоков между подвидами. Предковой для домашнего кролика является французская форма Oryctolagus c. cuniculus (48). Эта ограниченность настолько велика, как в Австралии, куда были также завезены представители испанского подвила (O. ...
... Only five of the 15 hybrids identified had dog content above 20%, confirming low levels of dog introgression and supporting the high genetic differentiation between Iberian wolves and dogs. Uneven distribution of dog genomic blocks across the 38 autosomes of backcross hybrids suggests different capacities to retain or purge introgressed alleles, as previously seen in several mammals, including the grey wolf (Frantz et al., 2015;Geraldes et al., 2006;Good et al., 2010;Pilot et al., 2021;Turner & Harr, 2014), birds (Carling & Brumfield, 2008;Runemark et al., 2018) and fishes (Schumer et al., 2018). Interestingly, the chromosomes of wolf-dog hybrids with significant excess of dog genomic content partially overlap with those previously found for Italian wolf-dog hybrids (Galaverni et al., 2017), suggesting either more permeable regions to introgression or regions with high ancestral shared variation. ...
Article
Full-text available
After decades of intense persecution, the Iberian wolf subspecies faced a severe bottleneck in the 1970s that considerably reduced its range and population size, nearly leading to its extinction in central and southern Iberian Peninsula. Such population decline could have impacted the genetic diversity of Iberian wolves through different processes, namely genetic drift and dynamics of hybridization with domestic dogs. By contrasting the genomes of 68 contemporary with 54 historical samples spanning the periods before and immediately after the 1970s bottleneck, we found evidence of its impact on genetic diversity and dynamics of wolf–dog hybridization. Our genome‐wide assessment revealed that wolves and dogs form two well‐differentiated genetic groups in Iberia and that hybridization rates did not increase during the bottleneck. However, an increased number of hybrid individuals was found over time during the population re‐expansion, particularly at the edge of the wolf range. We estimated a low percentage of dog ancestry (~1.4%) in historical samples, suggesting that dog introgression was not a key driver for wolf extinction in central and southern Iberia. Our findings also unveil a significant decline in genetic diversity in contemporary samples, with the highest proportion of homozygous segments in the genome being recently inherited. Overall, our study provides unprecedented insight into the impact of a sharp decline on the Iberian wolf genome and refines our understanding of the ecological and evolutionary drivers of wolf–dog hybridization in the wild.
... Sex chromosomes are disproportionately associated with genomic divergence and reproductive isolation (Coyne and Orr, 1989;Haldane, 1922;Presgraves, 2018), and the X chromosome is frequently found to show steeper clines across hybrid zones than other genomic regions (Geraldes et al., 2006;Hooper et al., 2018;Moran et al., 2018;Tucker et al., 1992). These effects also appear in ZW sex chromosome systems (such as for birds and butterflies), but here we focus on XY systems. ...
Article
Full-text available
Natural hybrid zones provide opportunities for studies of the evolution of reproductive isolation in wild populations. Although recent investigations have found that the formation of neo‐sex chromosomes is associated with reproductive isolation, the mechanisms remain unclear in most cases. Here, we assess the contemporary structure of gene flow in the contact zone between largely allopatric cytotypes of the dioecious plant Rumex hastatulus, a species with evidence of sex chromosome turn‐over. Males to the west of the Mississippi river, USA, have an X and a single Y chromosome, whereas populations to the east of the river have undergone a chromosomal rearrangement giving rise to a larger X and two Y chromosomes. Using reduced‐representation sequencing, we provide evidence that hybrids form readily and survive multiple backcross generations in the field, demonstrating the potential for ongoing gene flow between the cytotypes. Cline analysis of each chromosome separately captured no signals of difference in cline shape between chromosomes. However, principal component regression revealed a significant increase in the contribution of individual SNPs to inter‐cytotype differentiation on the neo‐X chromosome, but no correlation with recombination rate. Cline analysis revealed that the only SNPs with significantly steeper clines than the genome average were located on the neo‐X. Our data are consistent with a role for neo‐sex chromosomes in reproductive isolation between R. hastatulus cytotypes. Our investigation highlights the importance of studying plant hybrid zones for understanding the evolution of sex chromosomes.
... For example, the impact of hybridization and introgression on evolution can be diverse, from redistributing adaptive genetic variation 23,69,70 to generating negative epistasis between alleles that have evolved in different genomic backgrounds (Fishman and Sweigart, 71 Maheshwari and Barbash, 72 and Nosil and Schluter; 73 reviewed in Hedrick, 15 Suarez-Gonzalez et al., 16 Baack and Rieseberg, 74 and Moran et al. 75 ). The number of introgressed alleles that remain in a hybrid lineage depends on their selection coefficients, [76][77][78] their location in the genome (i.e., sex chromosomes versus autosomes [79][80][81] ), levels of divergence between the hybridizing species, 9,82,83 and recombination rates among loci. 6,84 Previous studies have, for example, shown that Drosophila hybrids often show maladaptive phenotypes. ...
Article
Genome-scale sequence data have invigorated the study of hybridization and introgression, particularly in animals. However, outside of a few notable cases, we lack systematic tests for introgression at a larger phylogenetic scale across entire clades. Here, we leverage 155 genome assemblies from 149 species to generate a fossil-calibrated phylogeny and conduct multilocus tests for introgression across 9 monophyletic radiations within the genus Drosophila. Using complementary phylogenomic approaches, we identify widespread introgression across the evolutionary history of Drosophila. Mapping gene-tree discordance onto the phylogeny revealed that both ancient and recent introgression has occurred across most of the 9 clades that we examined. Our results provide the first evidence of introgression occurring across the evolutionary history of Drosophila and highlight the need to continue to study the evolutionary consequences of hybridization and introgression in this genus and across the tree of life.
... With incompatibilities accumulating among their genomes, the crossing of parental species often affects the fertility of hybrids ultimately leading to their sterility [8,9]. Hybrid sterility may have various causes [10,11] and its molecular underpinning is still little understood. However, an important cause of hybrid sterility is the improper pairing and recombination between orthologous chromosomes from two different parental species during meiotic prophase, leading to the abruption of meiosis and/or the formation of aneuploid gametes [1,6,7]. ...
Article
Full-text available
The transition from sexual reproduction to asexuality is often triggered by hybridization. The gametogenesis of many hybrid asexuals involves premeiotic genome endoreplication leading to bypass hybrid sterility and forming clonal gametes. However, it is still not clear when endoreplication occurs, how many gonial cells it affects and whether its rate differs among clonal lineages. Here, we investigated meiotic and premeiotic cells of diploid and triploid hybrids of spined loaches (Cypriniformes: Cobitis) that reproduce by gynogenesis. We found that in naturally and experimentally produced F1 hybrids asexuality is achieved by genome endoreplication, which occurs in gonocytes just before entering meiosis or, rarely, one or a few divisions before meiosis. However, genome endoreplication was observed only in a minor fraction of the hybrid’s gonocytes, while the vast majority of gonocytes were unable to duplicate their genomes and consequently could not proceed beyond pachytene due to defects in bivalent formation. We also noted that the rate of endoreplication was significantly higher among gonocytes of hybrids from natural clones than of experimentally produced F1 hybrids. Thus, asexuality and hybrid sterility are intimately related phenomena and the transition from sexual reproduction to asexuality must overcome significant problems with genome incompatibilities with a possible impact on reproductive potential.
Article
Full-text available
Understanding the biogeography of species in space and time is essential for the development of evidence‐based conservation and management plans. In this paper we propose a biogeographical spatial modelling approach based on the favourability function, and developed under a fuzzy logic framework, to unravel the historical biogeography of the two European wild rabbit subspecies, Oryctolagus cuniculus algirus ( Oca ) and Oryctolagus cuniculus cuniculus ( Occ ), in the Iberian Peninsula (IP). We first reviewed published and unpublished information (PhD theses, scientific papers, technical reports, etc.) on the occurrence of each rabbit subspecies throughout the IP. We compiled data from 201 Iberian rabbit populations and from genetic information of 4348 rabbits that was used to identify subspecies. Only populations in which all rabbits surveyed belonged to one subspecies were considered in the modelling procedure. We modelled rabbit subspecies' distribution separately for populations in which nuclear and mitochondrial DNA sequences were available. We employed a trend surface analysis developed by logistic regressions, which applied the favourability function and fuzzy logic operations. Using our approach we indentify the expansion cores from which both rabbit subspecies would have expanded after isolation during the last glaciations. Furthermore, we reveal the possible existence of a competitive exclusion zone between both rabbit subspecies that may have prevented their further expansion. Finally, our study shows that the Oca subspecies is distributed in north‐western areas previously attributed to Occ . This assessment of the actual and historical distribution of each rabbit subspecies may allow more specific conservation interventions, as the two subspecies are not just genetically distinct but also ecologically and behaviourally different. Our methodological approach could be useful in unravelling the historical biogeography of other lesser‐known species.
Preprint
Full-text available
Natural hybrid zones provide opportunities for studies of the evolution of reproductive isolation in wild populations. Although several recent investigations have found that the formation of neo-sex chromosomes is associated with reproductive isolation, the mechanisms remain unclear in most cases. Here, we assess the contemporary structure of gene flow in the contact zone between largely allopatric cytotypes of the dioecious plant Rumex hastatulus, a species in which there is evidence of sex chromosome turn-over. Males to the west of the Mississippi river, USA, have an X and a single Y chromosome, whereas populations to the east of the river have undergone a chromosomal rearrangement giving rise to a larger X and two Y chromosomes. Using reduced-representation sequencing, we provide evidence that hybrids form readily and survive multiple backcross generations in the field, demonstrating the potential for ongoing gene flow between the cytotypes. At the scale of chromosomes, cline analysis of each chromosome separately captured no signals of difference in cline shape between chromosomes. However, when comparing SNPs, principal component regression revealed a significant increase in the contribution of individual SNPs to inter-cytotype differentiation on the neo-X, but no correlation with recombination rate. Cline analysis revealed that the only SNPs with significantly shallower clines than the genome-average were located on the neo-X. Our data are consistent with a role for the neo-sex chromosome in reproductive isolation between R. hastatulus cytotypes. Our investigation highlights the importance of studying plant hybrid zones in species with sex chromosomes for understanding mechanisms of reproductive isolation and for understanding the role of gene flow in governing the spread of the neo-X chromosomes.
Article
Full-text available
Genetic variation among populations of commensal house mice was studied across the territories of the Czech and Slovak Republics and in some adjacent areas of Germany. We used six diagnostic allozyme loci (Es-2, Gpd-1, Idh-1, Mpi, Np, Sod-1) and the following molecular markers: B1 insertion in the Btk gene (X chromosome), Zfy2 18-bp deletion (Y chromosome), BamHI restriction site in the nit-Nd1 gene (mtDNA) and Hba-4ps 16-bp insertion (diagnosing the presence of t haplotypes). In total, 544 individuals taken from 49 localities were examined. Almost the entire territories of the Czech Republic and Slovakia were found to be occupied by Mus musculus, the only exception being the westernmost parts of the Czech Republic, where M. musculus meets the range of M. domesticus and forms a narrow belt of hybrid populations. Despite this, domesticus-type alleles of some allozyme markers (notably Es-2) were also found at sites well within the range of M. musculus, either tens or hundreds of kilometres behind the hybrid zone. This provides evidence of either: (1) introgression of some markers into the species' genome due to free gene flow through the zone, or (2) human-mediated long-distance migrations, or (3) incomplete lineage sorting. Conversely, variants of molecular markers typical for M. domesticus in Btk, Zfy2 and mt-Nd1 were only found in the westernmost populations studied. t haplotypes were quite frequent in some populations, irrespective of whether M. domesticus, M. musculus or their hybrids, yet no t/t homozygotes were found. The mean frequency of t/+ heterozygotes found within the study populations was 13%.
Article
Full-text available
The divergence of Drosophila pseudoobscura from its close relatives, D. persimilis and D. pseudoobscura bogotana, was examined using the pattern of DNA sequence variation in a common set of 50 inbred lines at 11 loci from diverse locations in the genome. Drosophila pseudoobscura and D. persimilis show a marked excess of low-frequency variation across loci, consistent with a model of recent population expansion in both species. The different loci vary considerably, both in polymorphism levels and in the levels of polymorphisms that are shared by different species pairs. A major question we address is whether these patterns of shared variation are best explained by gene flow or by persistence since common ancestry. A new test of gene flow, based on patterns of linkage disequilibrium, is developed. The results from these, and other tests, support a model in which D. pseudoobscura and D. persimilis have exchanged genes at some loci. However, the pattern of variation suggests that most gene flow, although occurring after speciation began, was not recent. There is less evidence of gene flow between D. pseudoobscura and D. p. bogotana. The results are compared with recent work on the genomic locations of genes that contribute to reproductive isolation between D. pseudoobscura and D. persimilis. We show that there is a good correspondence between the genomic regions associated with reproductive isolation and the regions that show little or no evidence of gene flow.
Article
Genetic analyses of reproductive barriers represent one of the few methods by which theories of speciation can be tested. However, genetic study is often restricted to model organisms that have short generation times and are easily propagated in the laboratory. Replicate hybrid zones with a diversity of recombinant genotypes of varying age offer increased resolution for genetic mapping experiments and expand the pool of organisms amenable to genetic study. Using 88 markers distributed across 17 chromosomes, we analyze the introgression of chromosomal segments of Helianthus petiolaris into H. annuus in three natural hybrid zones. Introgression was significantly reduced relative to neutral expectations for 26 chromosomal segments, suggesting that each segment contains one or more factors that contribute to isolation. Pollen sterility is significantly associated with 16 of these 26 segments, providing a straightforward explanation of why this subset of blocks is disadvantageous in hybrids. In addition, comparison of rates of introgression across colinear vs. rearranged chromosomes indicates that close to 50% of the barrier to introgression is due to chromosomal rearrangements. These results demonstrate the utility of hybrid zones for identifying factors contributing to isolation and verify the prediction of increased resolution relative to controlled crosses.
Article
We compare the utility of two methods for estimating the average levels of gene flow from DNA sequence data. One method is based on estimating FST from frequencies at polymorphic sites, treating each site as a separate locus. The other method is based on computing the minimum number of migration events consistent with the gene tree inferred from their sequences. We compared the performance of these two methods on data that were generated by a computer simulation program that assumed the infinite sites model of mutation and that assumed an island model of migration. We found that in general when there is no recombination, the cladistic method performed better than FST while the reverse was true for rates of recombination similar to those found in eukaryotic nuclear genes, although FST performed better for all recombination rates for very low levels of migration (Nm = 0.1).
Article
The neutral theory of molecular evolution predicts that regions of the genome that evolve at high rates, as revealed by interspecific DNA sequence comparisons, will also exhibit high levels of polymorphism within species. We present here a conservative statistical test of this prediction based on a constant-rate neutral model. The test requires data from an interspecific comparison of at least two regions of the genome and data on levels of intraspecific polymorphism in the same regions from at least one species. The model is rejected for data from the region encompassing the Adh locus and the 5′ flanking sequence of Drosophila melanogaster and Drosophila sechellia. The data depart from the model in a direction that is consistent with the presence of balanced polymorphism in the coding region.
Article
In the famous last paragraph of On the Origin of Species, Darwin compared the “fixed law of gravity”, which causes the Earth to orbit the Sun, with the evolution of species by natural selection. This may be the first recorded case in biology of “physics envy”: the view that the proper task of the life sciences—in Darwin's case, evolutionary biology—is to emulate physics by establishing general laws and working out their consequences. Although we don't speak of “laws” in biology, we do have lawlike generalizations, including the near‐universality of the genetic code and the mechanism for translating DNA and RNA into proteins. This article is protected by copyright. All rights reserved.
Article
We compared the patterns of movement of sex chromosomal and autosomal loci along a 160 km transect across a zone of hybridization between M. domesticus and M. musculus in southern Germany and western Austria using seven genetic markers. These included one Y-specific DNA sequence (YB10), two X-specific loci (DXWas68 and DXWas31), and four autosomal isozyme loci (Es-10, Es-1, Mpi-1, and Np-1). Random effects logistic regression analysis enabled us to examine the relationship between M. domesticus allele frequency and geographic distance from the western edge of the hybrid zone and allowed statistical evaluation of differences in cline midpoint and width among loci. More limited movement was observed for all three sex chromosomal markers across the zone compared with three of the four autosomal markers. If differential movement reflects fitness differences of specific alleles (or alleles at closely linked loci) on a hybrid background, then alleles that move to a limited extent across a hybrid zone may contribute to hybrid breakdown between two species. The limited flow of both X- and Y-specific alleles suggest that sex chromosomes have played an important role in Mus speciation.
Article
Abstract A complete understanding of the speciation process requires the identification of genomic regions and genes that confer reproductive barriers between species. Empirical and theoretical research has revealed two important patterns in the evolution of reproductive isolation in animals: isolation typically arises as a result of disrupted epistatic interactions between multiple loci and these disruptions map disproportionately to the X chromosome. These patterns suggest that a targeted examination of natural gene flow between closely related species at X-linked markers with known positions would provide insight into the genetic basis of speciation. We take advantage of the existence of genomic data and a well-documented European zone of hybridization between two species of house mice, Mus domesticus and M. musculus, to conduct such a survey. We evaluate patterns of introgression across the hybrid zone for 13 diagnostic X-linked loci with known chromosomal positions using a maximum likelihood model. Interlocus comparisons clearly identify one locus with reduced introgression across the center of the hybrid zone, pinpointing a candidate region for reproductive isolation. Results also reveal one locus with high frequencies of M. domesticus alleles in populations on the M. musculus side of the zone, suggesting the possibility that positive selection may act to drive the spread of alleles from one species on to the genomic background of the other species. Finally, cline width and cline center are strongly positively correlated across the X chromosome, indicating that gene flow of the X chromosome may be asymmetrical. This study highlights the utility of natural populations of hybrids for mapping speciation genes and suggests that the middle of the X chromosome may be important for reproductive isolation between species of house mice.
Article
The divergence of Drosophila pseudoobscura and close relatives D. persimilis and D. pseudoobscura bogotana has been studied using comparative DNA sequence data from multiple nuclear loci. New data from the Hsp82 and Adh regions, in conjunction with existing data from Adh and the Period locus, are examined in the light of various models of speciation. The principal finding is that the three loci present very different histories, with Adh indicating large amounts of recent gene flow among the taxa, while little or no gene flow is apparent in the data from the other loci. The data were compared with predictions from several isolation models of divergence. These models include no gene flow, and they were found to be incompatible with the data. Instead the DNA data, taken together with other evidence, seem consistent with divergence models in which natural selection acts against gene flow at some loci more than at others. This family of models includes some sympatric and parapatric speciation models, as well as models of secondary contact and subsequent reinforcement of sexual isolation.
Article
The homeodomain is a DNA binding motif that is usually conserved among diverse taxa. Rapidly evolving homeodomains are thus of interest because their divergence may be associated with speciation. The exact site of the Odysseus (Ods) locus of hybrid male sterility inDrosophila contains such a homeobox gene. In the past half million years, this homeodomain has experienced more amino acid substitutions than it did in the preceding 700 million years; during this period, it has also evolved faster than other parts of the protein or even the introns. Such rapid sequence divergence is driven by positive selection and may contribute to reproductive isolation.