Plan
Comptes Rendus

Evolution/Évolution
DNA barcoding of African fruit bats (Mammalia, Pteropodidae). The mitochondrial genome does not provide a reliable discrimination between Epomophorus gambianus and Micropteropus pusillus
[« Codes-barres ADN » des chauves-souris frugivores africaines (Mammalia, Pteropodidae). Le génome mitochondrial ne permet pas une discrimination fiable entre Epomophorus gambianus et Micropteropus pusillus]
Comptes Rendus. Biologies, Volume 334 (2011) no. 7, pp. 544-554.

Résumés

Sequences of the mitochondrial cytochrome c oxidase subunit I (COI) gene have been shown to be useful for species identification in various groups of animals. However, the DNA barcoding approach has never been tested on African fruit bats of the family Pteropodidae (Mammalia, Chiroptera). In this study, the COI gene was sequenced from 120 bats collected in the Central African Republic and belonging to either Epomophorus gambianus or Micropteropus pusillus, two species easily diagnosed on the basis of morphological characters, such as body size, skull shape and palatal ridges. Two additional molecular markers were used for comparisons: the complete mitochondrial cytochrome b gene and the intron 7 of the nuclear β-fibrinogen (FGB) gene. Our results reveal an unexpected discordance between mitochondrial and nuclear genes. The nuclear FGB signal agrees with our morphological identifications, as the three alleles detected for E. gambianus are divergent from the fourteen alleles found for M. pusillus. By contrast, this taxonomic distinction is not recovered with the analyses of mitochondrial genes, which support rather a polyphyletic pattern for both species. The conflict between molecular markers is explained by multiple mtDNA introgression events from M. pusillus into E. gambianus or, alternatively, by incomplete lineage sorting of mtDNA haplotypes associated with positive selection on FGB alleles of M. pusillus. Our work shows the failure of DNA barcoding to discriminate between two morphologically distinct fruit bat species and highlights the importance of using both mitochondrial and nuclear markers for taxonomic identification.

Supplementary Materials:
Supplementary material for this article is supplied as a separate file:

Les séquences du gène mitochondrial de la première sous-unité de la cytochrome c oxydase (COI) sont de plus en plus utilisées comme « codes-barres ADN » pour identifier les espèces animales. Toutefois, cette approche n’a jamais été testée sur les chauves-souris frugivores d’Afrique. Lors de cette étude, le gène COI a été séquencé à partir de 120 chauves-souris collectées en République centrafricaine et appartenant à Epomophorus gambianus et Micropteropus pusillus, deux espèces faciles à différencier sur la base de caractères morphologiques, tels que la taille corporelle, la forme du crâne ou les plis palataux. Deux autres marqueurs moléculaires ont été utilisés : le gène mitochondrial du cytochrome b et l’intron 7 du gène nucléaire de la chaîne bêta du fibrinogène (FGB). Nos résultats révèlent une discordance inattendue entre les gènes mitochondriaux et le gène nucléaire. Les données nucléaires confirment les identifications morphologiques puisque les trois allèles détectés chez E. gambianus sont divergents des quatorze allèles découverts chez M. pusillus. En revanche, les analyses reposant sur les gènes mitochondriaux ne soutiennent pas cette distinction taxonomique, puisque les deux espèces apparaissent polyphylétiques. Un tel conflit peut être expliqué par de multiple évènements d’introgression du génome mitochondrial de M. pusillus vers E. gambianus, ou alternativement par la persistance d’haplotypes mitochondriaux ancestraux chez les deux espèces, associée à une sélection positive sur les allèles du gène FGB chez M. pusillus. Notre cas d’étude montre que l’approche des « codes-barres ADN » ne permet pas de distinguer deux espèces de chauves-souris morphologiquement très différentes. Ainsi, nous préconisons de réaliser les identifications moléculaires à l’aide d’une approche combinant le gène mitochondrial COI avec un ou plusieurs marqueurs nucléaires.

Compléments :
Des compléments sont fournis pour cet article dans le fichier séparé :

Métadonnées
Reçu le :
Accepté le :
Publié le :
DOI : 10.1016/j.crvi.2011.05.003
Keywords: Taxonomy, Phylogeny, Megachiroptera, Mitochondrial genome, Nuclear DNA, Topological incongruence, Interspecific hybridization, Ancestral polymorphism
Mot clés : Taxinomie, Phylogénie, Megachiroptera, Génome mitochondrial, ADN nucléaire, Incohérence topologique, Hybridation interspécifique, Polymorphisme ancestral
Nicolas Nesi 1, 2 ; Emmanuel Nakouné 3 ; Corinne Cruaud 4 ; Alexandre Hassanin 1, 2

1 UMR 7205, origine, structure et évolution de la biodiversité, département systématique et évolution, Muséum national d’histoire naturelle, case postale no 51, 55, rue Buffon, 75005 Paris, France
2 Service de systématique moléculaire, département systématique et évolution, Muséum national d’histoire naturelle, 43, rue Cuvier, 75005 Paris, France
3 Laboratoire des arbovirus, virus des fièvres hémorragiques, virus de la grippe et rage, institut Pasteur de Bangui, BP 923 Bangui, Central African Republic
4 Genoscope, Centre national de séquençage, 2, rue Gaston-Crémieux, CP5706, 91057 Evry cedex, France
@article{CRBIOL_2011__334_7_544_0,
     author = {Nicolas Nesi and Emmanuel Nakoun\'e and Corinne Cruaud and Alexandre Hassanin},
     title = {DNA barcoding of {African} fruit bats {(Mammalia,} {Pteropodidae).} {The} mitochondrial genome does not provide a reliable discrimination between {\protect\emph{Epomophorus~gambianus}} and {\protect\emph{Micropteropus~pusillus}}},
     journal = {Comptes Rendus. Biologies},
     pages = {544--554},
     publisher = {Elsevier},
     volume = {334},
     number = {7},
     year = {2011},
     doi = {10.1016/j.crvi.2011.05.003},
     language = {en},
}
TY  - JOUR
AU  - Nicolas Nesi
AU  - Emmanuel Nakouné
AU  - Corinne Cruaud
AU  - Alexandre Hassanin
TI  - DNA barcoding of African fruit bats (Mammalia, Pteropodidae). The mitochondrial genome does not provide a reliable discrimination between Epomophorus gambianus and Micropteropus pusillus
JO  - Comptes Rendus. Biologies
PY  - 2011
SP  - 544
EP  - 554
VL  - 334
IS  - 7
PB  - Elsevier
DO  - 10.1016/j.crvi.2011.05.003
LA  - en
ID  - CRBIOL_2011__334_7_544_0
ER  - 
%0 Journal Article
%A Nicolas Nesi
%A Emmanuel Nakouné
%A Corinne Cruaud
%A Alexandre Hassanin
%T DNA barcoding of African fruit bats (Mammalia, Pteropodidae). The mitochondrial genome does not provide a reliable discrimination between Epomophorus gambianus and Micropteropus pusillus
%J Comptes Rendus. Biologies
%D 2011
%P 544-554
%V 334
%N 7
%I Elsevier
%R 10.1016/j.crvi.2011.05.003
%G en
%F CRBIOL_2011__334_7_544_0
Nicolas Nesi; Emmanuel Nakouné; Corinne Cruaud; Alexandre Hassanin. DNA barcoding of African fruit bats (Mammalia, Pteropodidae). The mitochondrial genome does not provide a reliable discrimination between Epomophorus gambianus and Micropteropus pusillus. Comptes Rendus. Biologies, Volume 334 (2011) no. 7, pp. 544-554. doi : 10.1016/j.crvi.2011.05.003. https://comptes-rendus.academie-sciences.fr/biologies/articles/10.1016/j.crvi.2011.05.003/

Version originale du texte intégral

1 Introduction

The family Pteropodidae (Old World fruit bats) represents one of the most diversified taxa of the order Chiroptera with 189 species placed in 43 genera [1–3]. Its geographic distribution covers most tropical and subtropical zones of the Old World, from Africa (38 species) to Australia and Pacific islands (68 species), through India (13 species), East Asia (14 species) and Southeast Asia (84 species).

Many African fruit bat species have been recognised as potential reservoirs for viruses, including Eidolon helvum, Epomophorus gambianus, Epomophorus wahlbergi, Epomops buettikoferi, Micropteropus pusillus and Rousettus aegyptiacus for the Lagos bat virus [4,5], Eidolon helvum, Epomops franqueti, Hypsignathus monstrosus and Myonycteris torquata for the Ebola virus [6,7] and Rousettus aegyptiacus for the Marburg virus [8]. Whereas Lagos bat virus infection affects only cats and dogs, Ebola and Marburg viruses cause hemorrhagic fever, a severe, often-fatal disease in both human and great ape populations. Fruit bats appear to play a major role in disseminating zoonoses. To assess future threats posed by bat-associated zoonoses to humans, there is a need for accurate knowledge of the factors underlying the emergence of disease, as well as the specific relationships between the host and virus. One fundamental step for these studies is correct and accurate identification of the suspected reservoir species.

Species have traditionally been identified using a taxonomic key based on morphological characters. In the case of African fruit bats, genera can be distinguished by their skull shape, dental formula and unique palatal ridge pattern [9]. At the specific level, only a few characters can be used and most of them consist of adult body and skull dimensions. This lack of information makes identification tricky, notably when the specimens are juvenile and subadult, or when body measurements overlap greatly between closely related species. To deal with these difficulties, the DNA Barcode project was conceived as a standard system for identifying animal species. DNA barcoding is based on the subunit I of the cytochrome c oxidase (COI) gene from the mitochondrial DNA (mtDNA) genome, which acts as a ‘barcode’ to identify animal species [10]. Although the efficiency of DNA Barcoding has gained increasing validation, COI has not been used extensively on mammals and it has never been applied to African fruit bats.

In this study, the DNA Barcoding approach was tested on African fruit bats collected in the Central African Republic (CAR) in 2008. Our first COI data revealed very low genetic distance (0.14%) between two specimens belonging to two different species, Epomophorus gambianus Ogilby, 1835 and Micropteropus pusillus Peters, 1867. This result was highly unexpected, given that they are two morphologically distinct species. E. gambianus is a large fruit bat (mean weight of 105.5 g), whereas M. pusillus is a small-sized bat (mean weight of 27.5 g) (Fig. 1A). The skull of E. gambianus is long and narrow, whereas that of M. pusillus is short and broad (Fig. 1B). Finally, their six palatal ridges show distinct patterns: in E. gambianus, all are undivided, except for the last two postdental ridges, which are notched or narrowly divided, whereas, in M. pusillus, all are divided by a deep, continuous median groove, except for the first ridge, which is prominent and undivided (Fig. 1B).

Fig. 1

Morphological differences between Micropteropus pusillus (on the left) and Epomophorus gambianus (on the right). A. Forearm length (FA, mm) and body weight (W, g) of males (above the abscissa) and females (below the abscissa). Data were extracted from Bergmans [39,47]. B. Skull, ventral view (modified after [54,55]).

M. pusillus occurs in West and Central Africa (Fig. 2) in a variety of habitats, like savannah woodlands and vegetation belts on the edge of the forest [11]. E. gambianus is also a savannah and woodland species. It is sympatric with M. pusillus from Senegal to southern Ethiopia, but is also found in South-East Africa (Fig. 2).

Fig. 2

Distribution map of Epomophorus gambianus (red) and Micropteropus pusillus (green) (adapted from Bergmans [39,47] and [2]). The five localities sampled for this study are detailed in the box. Two subspecies have been described for E. gambianus: E. g. gambianus in the two northern areas (red) and E. g. crypturus in South-East Africa (orange) [39]. The latter subspecies has been elevated to species rank, i.e. Epomophorus crypturus, in several recent classifications [1,2].

Here we conduct a specific study on E. gambianus and M. pusillus, to understand better why COI sequences of these two species show such a low genetic distance. A total of 120 specimens were sequenced for two mitochondrial genes, i.e., COI (685 nt) and the complete cytochrome b gene (Cytb; 1140 nt), and the nuclear β-fibrinogen intron 7 (FGB; 700 nt). Our analyses reveal that mitochondrial markers do not provide a reliable signal for identifying E. gambianus and M. pusillus, whereas the nuclear marker supports the taxonomic distinction. Two hypotheses are retained to explain the misleading signal of mitochondrial genes: mtDNA introgression and incomplete lineage sorting of mtDNA haplotypes.

2 Materials and methods

2.1 Taxonomic sample

All samples were collected in CAR in 2008. Fruit bats were determined using the species identification key published by Bergmans [9]. A total of 23 individuals of E. gambianus from four localities and 97 individuals of M. pusillus from five geographic localities were included in this study. The localities are depicted in Fig. 2 and detailed in Table S1 (see Supplementary Material). In agreement with previous phylogenetic results [12], three outgroup species were used to root the tree: Epomops franqueti, which is another species of the tribe Epomophorini; Myonycteris torquata, which is a member of the tribe Myonycterini; and Rousettus aegyptiacus, which is the most distant outgroup, as it belongs to Rousettinae, a different subfamily from Epomophorinae.

2.2 DNA extraction, amplification and sequencing

Total genomic DNA was extracted from muscle or patagium samples using QIAGEN DNeasy Tissue Kit (Qiagen, Hilden, Germany) according to the manufacturer's protocol.

Three genes were sequenced for this study: two belonging to the mitochondrial genome (COI and Cytb) and one from the nuclear genome (a non-coding fragment: FGB). Primers used for polymerase chain reaction (PCR) were: UTyrLA and C1L705 for COI, UGluMA and LThrCH for Cytb, and 5′-CCA-CAA-CRG-CAT-GTT-CTT-CAG-CAC-3′ and 5′-GTA-TCT-GCC-ATT-TGG-ATT-GGC-TGC-3′ for FGB [13,14]. Amplifications were done in 20 μl using 2 μl of Buffer 10X with MgCl2, 0.8 μl of dNTP (6.6 mM), 0.12 μl of Taq DNA polymerase (2.5 U, Qiagen, Hilden, Germany) and 0.32 μl of the two primers at 10 μM. The standard PCR conditions were as follows: 4 min at 94 °C; 35 cycles of denaturation/annealing/extension with 30 s at 94 °C for denaturation, 45 s at 50 °C for annealing and 90 s at 72 °C for extension; and 10 min at 72 °C.

PCR products were purified using ExoSAP Kit (GE Healthcare, Buckinghamshire, UK) and then sequenced in both directions using an automated DNA Sequencer (Applied Biosystems 3100). These two last steps were performed by GENOSCOPE (Ivry-sur-Seine, France). Nuclear fragments that contained multiple heterozygous single-nucleotide polymorphisms (characterized by double peaks in the chromatogram) were cloned using a QIAGEN PCR Cloningplus Kit in order to isolate the exact sequence of each allele in heterozygous individuals. Sequences were edited and assembled using Sequencher 4.7 (Gene Codes Corporation). Sequences generated for this study were deposited in the EMBL/DDBJ/Genbank database (accession numbers JF728368-JF728760; Table S1 for details).

2.3 Sequence data analysis

Sequences were aligned by eye using BioEdit version 7.0.9 [15]. No gaps were included in the alignments of COI and Cytb. For FGB, two unambiguous gaps were inferred to align the most divergent outgroup species, R. aegyptiacus, with other taxa.

Phylogenetic analyses were performed using Neighbor Joining (NJ) and Bayesian methods. The three genes (COI: 685 nt; Cytb: 1140 nt; FGB: 700 nt) were analyzed separately to evaluate their own signal and to detect any incongruence. The mitochondrial markers were then combined (COI + Cytb). MODELTEST version 3.7 [16] was run to determine the best-fitting evolutionary model. According to AIC criterion, the selected model was HKY + G for COI and Cytb, TVM + G for COI + Cytb and HKY for FGB. PAUP* 4b10 [17] was used for the NJ analyses with the selected model. Nodal support was assessed by the bootstrap analysis (1000 replicates). Bayesian inferences were conducted with MrBayes v3.1.2 [18]. Posterior probabilities (PP) were calculated using four independent Markov Chains run for 10,000,000 Metropolis-coupled MCMC generations with tree sampling every 1000 generations. TRACER [19] was used to calculate the effective sample size statistic to make sure that a sufficient number of generations were run and to decide on the length of the burn-in period. Analyses were run twice independently to check for convergence of the results.

Heterozygosity of nuclear genes complicates phylogenetic analyses because the two alleles from a heterozygous individual are not necessarily most closely related to each other. In practice, for many phylogenies based on diploid gene sequences, allelic sequences are not separated and heterozygous sites are treated as polymorphic. This procedure can make phylogenetic inference ambiguous, especially for closely related taxa [20]. However, the pairing of different alleles can provide information on interbreeding and may be useful in delimitating species. Therefore, the nuclear sequences of FGB were analyzed using the median-joining algorithm available in Network version 4.5.1 [21].

The pairwise genetic distances were calculated with PAUP* 4b10 [17] using the Kimura 2-parameter (K2P) correction. For each species, we calculated the number of polymorphic sites, number of haplotypes, nucleotide diversity (π) and haplotype diversity (h) [22,23] using DnaSP version 5.10.01 [24].

An “Isolation with Migration” (IM) model was used to test introgression rate between the two fruit bat species. We used a Bayesian MCMC method in the IMa version 2.0 coalescent program [25], which attempts to fit the data to a null model where an ancestral population bifurcates into two allopatric populations 1 and 2. This MCMC simulation estimates the joint posterior probability of six demographic parameters: population size of species 1 and 2 (Ne1, Ne2) and of the ancestral species (NeA), migration (m1, m2) and divergence time (t). In our study, subscript 1 refers to the M. pusillus species and subscript 2 refers to the E. gambianus species. Although migration typically refers to movements of individuals between populations, the m parameter is better interpreted as the rate of mtDNA introgression for the purposes of this study. In this way, the m1 parameter refers to introgression from E. gambianus into M. pusillus, whereas the m2 parameter refers to introgression from M. pusillus into E. gambianus. As recommended in Hey & Nielsen [26], the inheritance scalars were set to 1 for the nuclear locus (FGB) and 0.25 for the mitochondrial sequences (COI + Cytb) and the HKY model was used. To assess convergence, we checked effective sample sizes throughout the run and compared results between three independent runs. Following the approach used by Won and Hey [27], we first ran IMa using large, flat priors for each parameter. Based on those runs, we defined narrower upper bounds that encompassed the full posterior distribution of each parameter. In the final runs, priors were set to 5 for t, 2 for m, 1000 for NeA and 100 for Ne1 and Ne2. The input files were executed in IMa using 10 Metropolis-coupled chains with 100,000 steps for burn-in followed by 10 million steps for parameter estimation.

3 Results

3.1 Analyses of mitochondrial genes

The 5′ region of COI (685 nt) and the complete Cytb gene (1140 nt) were sequenced for all samples available for E. gambianus (23 individuals) and M. pusillus (97 individuals). Mitochondrial genes showed high nucleotide and haplotype diversities for both species (Table 1), suggesting large effective size and stable populations (i.e., no recent bottleneck). Among COI sequences, a total of 60 haplotypes were identified: seven specific to E. gambianus, 49 specific to M. pusillus and four shared between the two species (Table S1). Among Cytb sequences, thirteen haplotypes were found for E. gambianus, 71 for M. pusillus and none were shared between the two taxa (Table S1). Using the alignment combining COI and Cytb sequences, we identified thirteen haplotypes for E. gambianus, 83 for M. pusillus and none were shared between the two taxa (Table S1).

Table 1

Genetic variability of molecular markers sequenced for Epomophorus gambianus and Micropteropus pusillus.

Locus Species N S Hap h ± SD π ± SD Het
Cytb M. pusillus 97 134 71 0.984 ± 0.007 0.00859 ± 0.00057
E. gambianus 23 48 13 0.933 ± 0.001 0.00987 ± 0.00062
COI M. pusillus 97 70 53 0.956 ± 0.014 0.00935 ± 0.00068
E. gambianus 23 24 11 0.909 ± 0.034 0.01042 ± 0.0007
COI + Cytb M. pusillus 97 204 83 0.995 ± 0.003 0.00888 ± 0.00056
E. gambianus 23 72 13 0.933 ± 0.033 0.01008 ± 0.00056
FGB M. pusillus 97 16 14 0.333 ± 0.058 0.00118 ± 0.00031 17
E. gambianus 23 2 3 0.360 ± 0.010 0.00055 ± 0.00017 6

The Bayesian tree reconstructed from the alignment combining COI and Cytb genes (123 taxa, 1825 nt) is shown in Fig. 3. All individuals of E. gambianus and M. pusillus fall into a robust clade (PP = 1), but neither species was found to be monophyletic. The trees obtained from separate analyses of COI and Cytb genes (Figs. S3 and S1) show very similar relationships, but the COI tree is less resolved than the Cytb tree, because COI has much less phylogenetic information content (165 and 89 parsimony informative sites for Cytb and COI, respectively). Moreover, the analysis including one complete Cytb sequence of Epomophorus wahlbergi (accession numbers DQ445706) revealed that the genus Epomophorus is paraphyletic, as E. wahlbergi occupies a more basal position with respect to the clade uniting E. gambianus and M. pusillus sequences (Fig. S2).

Fig. 3

Bayesian tree obtained from the analysis of mitochondrial genes. The values on the branches indicate posterior probabilities (> 0.75) calculated from left to right with three alignments: the combination of COI and Cytb genes, Cytb and COI.

In Fig. 4, the comparisons between pairwise COI genetic distances show that DNA barcodes cannot be used for discriminating between E. gambianus and M. pusillus. Indeed, interspecific distances, which are comprised between 0% and 2.5%, are not higher than intraspecific distances found for E. gambianus (0–2%) and M. pusillus (0–2.5%). As a consequence, a query sequence cannot be correctly assigned to either E. gambianus or M. pusillus.

Fig. 4

Intra- and interspecific COI distances. Pairwise distances were calculated with the Kimura 2-parameter correction for comparisons within Micropteropus pusillus (green histograms), within Epomophorus gambianus (red histograms) and for interspecific comparisons (grey histograms).

3.2 Analyses of nuclear alleles

The nuclear FGB gene (700 nt) was also sequenced for all samples available for E. gambianus and M. Pusillus. The nuclear gene showed low nucleotide and haplotype diversities for both species (Table 1). We found three alleles for E. gambianus and fourteen alleles for M. pusillus (Table S1). None of them was shared between the two species. We identified six heterozygous individuals for E. gambianus (26%) and seventeen for M. pusillus (17.5%) (Table 1). In contrast with mtDNA data, the comparisons of FGB sequences indicate a higher haplotype diversity for E. gambianus and a higher nucleotide diversity for M. pusillus (Table 1).

The allelic network of FGB is presented in Fig. 5. The three alleles of E. gambianus are closely related and separated by only one mutation. The fourteen alleles of M. pusillus can be divided into two linked groups: the first one forms a star-like topology with a main central allele (Mp1), from which ten alleles are derived by one mutational step (Mp2 to Mp9; Mp12; Mp13) and two alleles by two mutational steps (Mp10 and Mp11); the second group corresponds to allele Mp14. Interestingly, the allele Mp14 is found in only five heterozygous individuals of M. pusillus and it is always associated with the most frequent allele, Mp1. In addition, Mp14 is six mutational steps distant from other alleles of M. pusillus, but it differs by only two mutations from Eg1, the most common allele of E. gambianus.

Fig. 5

Median-joining network of FGB alleles. The three alleles found for Epomophorus gambianus are indicated by red circles. The fourteen alleles found for Micropteropus pusillus are indicated by green circles. The circle size is proportional to the number of alleles found in the populations with the exception of alleles Mp1 and Eg1. The number of substitutions (greater than one) between alleles are indicated on the branches. We detailed the nature and position of all substitutions separating the alleles of E. gambianus and M. pusillus.

3.3 Introgression rates

The introgression (or migration) parameter curves obtained from the isolation with migration program (IMa version 2.0) are presented in Fig. 6. In the case of introgression of E. gambianus into M. pusillus, the curve has a peak near zero and the probability of the peak is very near to what the probability is at zero. In this case, we cannot reject an introgression rate of zero. However, in the opposite direction, i.e., introgression from M. pusillus into E. gambianus, the peak is far from zero and the estimated probability of zero introgression is zero; we can therefore reject a model of no introgression from M. pusillus into E. gambianus.

Fig. 6

The marginal posterior probability distributions of the introgression parameter, as estimated from the isolation with migration (IMa) program. The grey curve corresponds to the hypothesis of mtDNA introgression from E. gambianus into M. pusillus. The black curve corresponds to the hypothesis of mtDNA introgression from M. pusillus into E. gambianus. Statistical significance as assessed by the likelihood ratio test is indicated by * (< 0.001).

4 Discussion

4.1 Discordance between mitochondrial and nuclear data sets

According to Bergmans [9], Epomophorus and Micropteropus can be ranged in the tribe Epomophorini with the three other genera Epomops, Hypsignathus and Nanonycteris. By analyzing mtDNA sequences of Cytb and 16S rRNA genes, Juste et al. [12] have concluded on a sister-group relationship between Epomophorus wahlbergi and Micropteropus pusillus, whereas Epomops franqueti was found to be more divergent. Other species of Epomophorini were however not included in the molecular analyses. Although our mtDNA results confirm that Epomophorus and Micropteropus are closely-related genera, it was highly unexpected to find a paraphyletic pattern for Epomophorus (Fig. S2) and a polyphyletic pattern for both E. gambianus and M. pusillus species (Fig. 3). Indeed, M. pusillus can be easily distinguished from E. gambianus and E. wahlbergi on the basis of morphological characteristics, such as body size (Fig. 1; weight: 20–35 g versus 56–155 g for Egambianus and 54–125 g for E. wahlbergi; forearm length: 48–56 mm versus 75–100 mm for E. gambianus and 68–95 mm for E. wahlbergi) and the pattern of palatal ridges (Fig. 1; only five bridges for Ewahlbergi, see in Bergmans [9]). Since our analyses of the nuclear gene FGB agree with morphology, as both M. pusillus and E. gambianus possess distinct alleles (Fig. 5), we can conclude that mtDNA data are problematic. Three hypotheses can be advanced to explain the incongruence of mtDNA with nuclear and morphological data: the amplification of Numts, mtDNA introgression and incomplete lineage sorting of mtDNA haplotypes.

4.2 Can Numts explain the mitonuclear discordance?

Nuclear mitochondrial pseudogenes or ‘Numts’ are segments of mtDNA translocated to the nuclear genome. These paralogous sequences are commonly found in animal genomes and can be accidentally amplified instead of the targeted mitochondrial genes [28]. The undetected presence of Numts in the analyses can result in erroneous interpretations of phylogenetic relationships, because Numts evolve under constraints that are different than those of mtDNA [29–31]. The nuclear and mtDNA genomes have different rates and patterns of mutations and, because Numts are not functional, they do not evolve under purifying selection, explaining why they readily accumulate stop codon and frameshift mutations [32].

Four major arguments suggest that the COI and Cytb sequences produced in the present study cannot be considered as being Numts: (1) as the primers used here for PCR and sequencing were specially designed to amplify mtDNA genes, all our chromatograms were perfectly readable, without double peaks; (2) there is a high identity (99.8%) between our Cytb sequences of M. pusillus and the partial sequence previously obtained by Juste et al. [12] using a different set of primers; (3) there is no stop codon or indel (insertion or deletion); and (4) the tree reconstructed from COI sequences is very similar to that obtained from Cytb sequences (Figs. S3 and S1), with several identical nodes robustly supported in both analyses (Fig. 3). Therefore, there is no doubt about the mitochondrial origin of our COI and Cytb sequences.

4.3 Rampant mitochondrial introgression into Epomophorus gambianus?

The occasional mating between distinct species can result in genetic introgression of the mtDNA genome or nuclear alleles of one species into the gene pool of another species. This is a cause of gene/species tree discordance. Mitochondrial introgression following hybridization has been widely inferred, especially when dealing with differences in genetic signal between mtDNA and nuclear markers [13,33–36], but documented cases in bats are very rare [37]. The mtDNA from one species may completely replace that of another species, without leaving any trace of nuclear introgression or morphological change [38]. In such cases, the mtDNA tree generally shows a typical topology, in which one of the two species is found to be monophyletic and the other paraphyletic. In our case, the mtDNA tree shows a more complex pattern, in which both species are polyphyletic. Assuming the hypothesis of mtDNA introgression implies therefore that interspecific breeding occurred frequently and at different periods of time. In addition, IMa analyses suggest that mtDNA introgression events occurred from M. pusillus into E. gambianus. According to this scenario, females of M. pusillus have been integrated several times into E. gambianus populations and all thirteen mtDNA haplotypes detected here for E. gambianus were transferred from M. pusillus.

Three main arguments concerning DNA variation, ecology and sexual behaviour render this scenario plausible.

Firstly, mtDNA variations between E. gambianus and M. pusillus are between 0 and 2.5% for COI and between 0 and 2.4% for Cytb. These low levels of variation are comparable with intraspecific divergence calculated for COI and Cytb sequences in several species of the family Pteropodidae (Tables S2 and S3). By contrast, the complete Cytb sequence available for E. wahlbergi is divergent from those of E. gambianus and M. pusillus (4.8–5.7%). These comparisons support therefore the fact that all mtDNA sequences produced here for E. gambianus and M. pusillus originated from only one species, namely M. pusillus.

Secondly, both species are found in sympatry from Senegal to CAR (Fig. 2), where they occupy the same type of habitat, i.e. the forest-savannah mosaic [9,39], which is a transition zone, or ecotone, between the tropical moist broadleaf forests and the drier savannahs. As a consequence, they were often collected in the same mist nets (personal observations) and were observed roosting in the same tree [40]. The geographic distribution of M. pusillus (Fig. 2) suggests, however, that it is more dependent on rainforests than E. gambianus. Hybrid zones often coincide with ecotones, as previously shown in African elephants [33] or mouse lemurs [41]. Astonishingly, the situation in African fruit bats is comparable with that of African elephants, as multiple events of mtDNA introgression have been inferred from the smallest forest species (M. pusillus/L. cyclotis) into the largest savannah species (E. gambianus/L. africana).

Thirdly, introgressive hybridization is known to occur more frequently between sister-species [42], for which prezygotic isolation (behavioural and mechanical) and postzygotic isolation (zygote mortality and hybrid inviability and sterility) have not been fully accomplished. Here, both species share very similar reproductive patterns. Females have two parturition periods during rainy seasons around April and September/October in West Africa. Furthermore, adult males exhibit similar sexual characteristics (white tufts of erectile hair grow within circular pocket-like folds of skin on the shoulders) and have the same typical courtship behaviour (with courting calls, the display of the white shoulder patches and wing flapping) [40,43].

However, three arguments dispute an introgression from M. pusillus into E. gambianus. Firstly, the male courting call is at a different frequency: 1750 and 2800 Hz for Epomophorus and M. pusillus, respectively [44]. Secondly, the two species have different male karyotypes: M. pusillus is XY1Y2, with 2n = 35 and a fundamental number (FN) of 64 [45], whereas E. gambianus is XO, with 2n = 35 and a fundamental number (FN) of 68 [46]. The karyotype available for E. gambianus was established from two specimens from Kenya and Zimbabwe, i.e. from two populations of E. gambianus crypturus. Since this subspecies has been elevated to full species status in recent classifications, i.e. Epomophorus crypturus [1,2], we can expect a different karyotype for specimens from CAR. Thirdly, and probably most importantly, their body size is very different, as adult males of E. gambianus are four times heavier than females of M. pusillus (Fig. 1; [39,47]). Such morphological differences may preclude copulation and cause mortality in pregnant M. pusillus females carrying an overly large hybrid foetus.

4.4 Incomplete lineage sorting of mtDNA haplotypes and positive selection in the FGB gene?

As an alternative to the hypothesis of mtDNA introgression, the phenomenon of incomplete lineage sorting, due to the retention of ancient polymorphisms, can be invoked to explain the polyphyletic pattern observed here for mtDNA haplotypes of E. gambianus and M. pusillus. If the mtDNA genome has not reached coalescence at the species level, some lineages from one species may be more closely related to a lineage in the other species than to other lineages within the same species. Incomplete lineage sorting can take place between sister-species, when the recent speciation event occurs before sorting is completed [48,49]. A recent speciation event between E. gambianus and M. pusillus is here supported by the fact that both mtDNA and nuDNA markers show very low interspecies distances (< 2.5% for mtDNA and < 1% for nuDNA).

The shorter the coalescence time, the less likely the gene will suffer from the ‘incomplete lineage sorting’ problem. In this regard, mtDNA has an advantage, because it coalesces in one-fourth the time of nuclear genes [48,49]. Therefore, if incomplete lineage sorting of mtDNA haplotypes is possibly the cause of the polyphyly, we should assume a faster time of coalescence for the FGB nuclear gene than for the mitochondrial genome.

A strong positive selection in the FGB gene may have led to selecting one or several particular allelic forms rapidly in one or both species before the coalescence of mtDNA. Actually, positive selection can be advanced for M. pusillus, because most FGB alleles detected in this species are highly derived with respect to those found in E. gambianus. Indeed, Mp1 to Mp13 alleles differ by five mutations from the ancestral allelic form shared by E. gambianus and M. pusillus (Fig. 5). In accordance with the hypothesis of positive selection, the primitive allele Mp14 is very rare in M. pusillus (2.6%) and it is always found in heterozygosity with a derived Mp1 allele, suggesting that at least one derived allele is necessary to ensure the function of FGB in M. pusillus.

The FGB gene codes for the beta polypeptide chains of the fibrinogen, a glycoprotein involved in blood coagulation. In humans, several polymorphisms in the FGB gene were found to affect the concentration of clotting factor and therefore the plasma viscosity [50,51] or to protect against myocardial infarction [52]. In addition, previous physiological studies on fruit bats have concluded that there is a linear correlation between body mass and heart pulse rate [53]. In other words, the smallest bats, which have the highest wing beat frequency, have the highest heart rate. Therefore, we can expect to have a higher selective pressure for FGB alleles in M. pusillus than in E. gambianus.

5 Conclusion

Our study has revealed that mtDNA barcodes are not useful for assigning an unknown specimen to either E. gambianus or M. pusillus. By chance, these two species can be easily identified in the field. Within Epomophorus, several sympatric species are however difficult to distinguish using a morphological taxonomic key. For example, the distinction between E. anselli, E. gambianus, E. labiatus and E. minimus, may be tricky in East Africa, as diagnosing morphological characters are only measurements, which show important overlapping variations between species and which cannot be used on juvenile and subadult specimens.

Additional nuclear markers and other individuals from allopatric populations of E. gambianus (i.e., from South-East Africa) and M. pusillus (i.e., from Equatorial Africa), as well as from other species of Epomophorus and Micropteropus, need to be included in further molecular studies in order to decipher between introgression and incomplete lineage sorting and to provide taxonomic conclusions at genus and species levels.

Disclosure of interest

The authors declare that they have no conflicts of interest concerning this article.

Acknowledgements

We are very grateful to Carine Ngoagouni who helped in collecting tissue samples. We thank Pr. Alain Le Faou, from the Institut Pasteur of Bangui, for permission and support to conduct research in CAR. We thank Céline Bonillo for help in DNA cloning. This work was supported by the Muséum National d’Histoire Naturelle (MNHN), CNRS, Institut Pasteur of Bangui, PPF “État et structure phylogénétique de la biodiversité actuelle et fossile,” and “Consortium National de Recherche en Génomique.” It is part of agreement No. 2005/67 between the Genoscope and the MNHN on the project “Macrophylogeny of Life”.


Bibliographie

[1] N.B. Simmons Order Chiroptera (D.E. Wilson; D.M. Reeder, eds.), Mammals species of the world: a taxonomic and geographic reference, 1, Johns Hopkins University Press, 2005, pp. 312-529

[2] IUCN Red List of Threatened Species. Version 2010.4. www.iucnredlist.org.

[3] K.M. Helgen; L.E. Helgen; D.E. Wilson Pacific flying foxes (Mammalia: Chiroptera): two new species of Pteropus from Samoa, probably extinct, Am Museum Novitates, Volume 3646 (2009), pp. 1-37

[4] D.T.S. Hayman; A.R. Fooks; D. Horton; R. Suu-Ire; A.C. Breed; A.A. Cunningham; J.L.N. Wood Antibodies against Lagos bat virus in Megachiroptera from West Africa, Emerg. Infect. Dis., Volume 14 (2008), pp. 926-928

[5] W. Markotter; I. Kuzmin; C.E. Rupprecht; L.H. Nel Phylogeny of Lagos bat virus: challenges for lyssavirus taxonomy, Virus Res., Volume 135 (2008), pp. 10-21

[6] E.M. Leroy; B. Kumulungui; X. Pourrut; P. Rouquet; A. Hassanin; P. Yaba; A. Delicat; J.T. Paweska; J.P. Gonzalez; R. Swanepoel Fruit bats as reservoirs of Ebola virus, Nature, Volume 438 (2005), pp. 575-576

[7] D.T.S. Hayman; P. Emmerich; L.F. Yu; L.F. Wang; R. Suu-Ire; A.R. Fooks; A.A. Cunningham; J.L.N. Wood Long-term survival of an urban fruit bat seropositive for Ebola and Lagos viruses, Plos One, Volume 5 (2010), p. e11978

[8] J.S. Towner; X. Pourrut; C.G. Albarino; C.N. Nkogue; B.H. Bird; G. Grard; T.G. Ksiazek; J.P. Gonzalez; S.T. Nichol; E.M. Leroy Marburg virus infection detected in a common African bat, Plos One, Volume 2 (2007) no. 8, p. e764

[9] W. Bergmans Taxonomy and biogeography of African fruit bats (Mammalia, Megachiroptera). 5. The genera Lissonycteris Andersen, 1912, Myonycteris Matschie, 1899 and Megaloglossus Pagenstecher, 1885; general remarks and conclusions; annex: key to all species, Beaufortia, Volume 47 (1997), pp. 11-90

[10] P.D.N. Hebert; A. Cywinska; S.L. Ball; J.R. DeWaard Biological identification through DNA barcodes, Proc R Soc Lond B, Volume 270 (2003), pp. 313-321

[11] R.W. Hayman; J.E. Hill Order Chiroptera, The Mammals of Africa. An identification manual, Part 2, Smithsonian Institution Press, City of Washington, 1971

[12] J. Juste; Y. Alvarez; E. Tabares; A. Garrido-Pertierra; C. Ibanez; J.M. Bautista Phylogeography of African fruitbats (Megachiroptera), Mol. Phylogenet. Evol., Volume 13 (1999), pp. 596-604

[13] A. Hassanin; A. Ropiquet Resolving a zoological mystery: the Kouprey is a real species, Proc. R. Soc. Lond. B., Volume 274 (2007), pp. 2849-2855

[14] A. Hassanin, A. Ropiquet, F. Delsuc, C. Hammer, B. Jansen van Vuuren, C. Matthee, M. Ruiz-Garcia, F. Catzeflis, V. Areskoug, T.T. Nguyen, A. Couloux. Pattern and timing of diversification of Cetartiodactyla (Mammalia, Laurasiatheria), as revealed by a comprehensive analysis of mitochondrial genomes, submitted.

[15] T.A. Hall BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT, Nucleic Acids Symposium Series, Volume 41 (1999), pp. 95-98

[16] D. Posada; K.A. Crandall MODELTEST: testing the model of DNA substitution, Bioinformatics, Volume 14 (1998), pp. 817-818

[17] D.L. Swofford PAUP*: phylogenetic analysis using parsimony (*and other methods). Version 4. 0b10, Sinauer Associates, Sunderland, MA, 2003

[18] J.P. Huelsenbeck; F. Ronquist MRBAYES: Bayesian inference of phylogenetic trees, Bioinformatics, Volume 17 (2001), pp. 754-755

[19] A. Rambaut, A.J. Drummond, Tracer Version 1.4 software available at http://tree.bio.ed.ac.uk/software/tracer.

[20] T. Sota; A.P. Vogler Reconstructing species phylogeny of the carabid beetles Ohompterus using multiple nuclear DNA sequences: heterogeneous information content and the performance of simultaneous analyses, Mol. Phylogenet. Evol, Volume 26 (2003), pp. 139-154

[21] H.J. Bandelt; P. Forster; A. Rohl Median-joining networks for inferring intraspecific phylogenies, Mol. Biol. Evol., Volume 16 (1999), pp. 37-48

[22] M. Nei Molecular evolutionary genetics, Colombia University Press, 1987

[23] F. Tajima Measurement of DNA polymorphism, Mechanisms of molecular evolution, Sinauer Associates, 1993 (p. 37–59)

[24] P. Librado; J. Rozas DnaSP v5: a software for comprehensive analysis of DNA polymorphism data, Bioinformatics, Volume 25 (2009), pp. 1451-1452

[25] J. Hey Isolation with migration models for more than two populations, Mol. Biol. Evol., Volume 27 (2010), pp. 905-920

[26] J. Hey; R. Nielsen Multilocus methods for estimating population sizes, migration rates and divergence time, with applications to the divergence of Drosophila pseudoobscura and D. Persimilis, Genetics, Volume 167 (2004), pp. 747-760

[27] Y.J. Won; J. Hey Divergence population genetics of chimpanzees, Mol. Biol. Evol., Volume 22 (2005), pp. 297-307

[28] D. Bensasson; D.X. Zhang; D.L. Hartl; G.M. Hewitt Mitochondrial pseudogenes: evolution's misplaced witnesses, Trends Ecol. Evol., Volume 16 (2001), pp. 314-321

[29] D.A. Triant; J.A. Dewoody The occurrence, detection, and avoidance of mitochondrial DNA translocations in mammalian systematic and phylogeography, J. Mammalogy, Volume 88 (2007), pp. 908-920

[30] H. Song; J.E. Buhay; M.F. Whiting; K.A. Crandall Many species in one: DNA barcoding overestimates in the number of species when nuclear mitochondrial pseudogenes are coamplified, PNAS, Volume 105 (2008), pp. 13486-13491

[31] A. Hassanin; A. Ropiquet; A. Couloux; C. Cruaud Evolution of the Mitochondrial Genome in mammals living at high altitude: new insights from a study of the Tribe Caprini (Bovidae, Antilopinae), J. Mol. Evol., Volume 68 (2009), pp. 293-310

[32] A. Hassanin; C. Bonillo; X.N. Bui; C. Cruaud Comparisons between mitochondrial genomes of domestic goat (Capra hircus) reveal the presence of Numts and multiple sequencing errors, Mitochondrial DNA, Volume 21 (2010), pp. 68-76

[33] A.L. Roca; N. Georgiadis; S.J. O’Brien Cytonuclear genomic dissociation in African elephant species, Nature Genetics, Volume 37 (2005), pp. 96-100

[34] A. Ropiquet; A. Hassanin Hybrid origin of the Pliocene ancestor of wild goats, Mol. Phylogenet. Evol., Volume 41 (2006), pp. 395-404

[35] P.C. Alves; J. Melo-Ferreira; H. Freitas; P. Boursot The ubiquitous mountain hare mitochondria: multiple introgressive hybridization in hares, genus Lepus, Philo, Transactions of the Royal Society B: Biological Sciences, Volume 363 (2008), pp. 2831-2839

[36] R.J. Petit; L. Excoffier Gene flow and species delimitation, Trends Ecol. Evol., Volume 24 (2009), pp. 386-393

[37] P. Berthier; L. Excoffier; M. Ruedi Recurrent replacement of mtDNA and cryptic hybridization between two sibling bat species Myotis myotis and Myotis blythii, Pro. R. Soc. Lond. B., Volume 273 (2006), pp. 3101-3109

[38] J.W.O. Ballard; M.C. Whitlock The incomplete natural history of mitochondria, Mol. Ecol., Volume 13 (2004), pp. 729-744

[39] W. Bergmans Taxonomy and biogeography of African fruit bats (Mammalia, Megachiroptera). 2. The genera Micropteropus Matschie 1899 Epomops Gray 1870 Hypsignathus H. Allen 1861 Nanonycteris Matschie 1899 and Plerotes Andersen 1910, Beaufortia, Volume 39 (1989), pp. 89-153

[40] M.C. Boulay; C.B. Robbins Epomophorus gambianus in mammalian species, Mammalian Species, 344, American Society of Mammalogists, 1989 (p. 1–5)

[41] M. Gligor; J.U. Ganzhorn; D. Rakotondravony; O.R. Ramilijaona; E. Razafimahatratra; H. Zischler; A. Hapke Hybridization between mouse lemurs in an ecological transition zone in southern Madagascar, Mol. Ecol., Volume 18 (2009), pp. 520-533

[42] L. Gay; G. Neubauer; M. Zagalska-Neubauer; C. Debain; J.M. Pons; P. David; P.A. Crochet Molecular and morphological patterns of introgression between two large white-headed gull species in a zone of recent secondary contact, Mol. Ecol., Volume 16 (2007), pp. 3215-3227

[43] N.T. Owen-Ashley; D.E. Wilson Micropteropus pusillus , Mammalian Species, 577, American Society of Mammalogists, 1998 (p. 1–5)

[44] J. Haft Beobachtungen zu Balzverhalten und Jugentwicklung beim Zwerg-Epualettenflughund, Micropteropus pusillus Peters, 1868 (Mammalia: Megachiroptera: Epomophorini), Zoologische Garten, Volume 72 (2002), pp. 61-67

[45] M.W. Haiduk; R.J. Baker; L.W. Robbins; D.A. Schlitter Chromosomal evolution in African megachiroptera: G- and C-band assessment of the magnitude of change in similar standard karyotypes, Cytogenet. Cell Genet., Volume 29 (1981), pp. 221-232

[46] R.L. Peterson; D.W. Nagorsen Chromosomes of fifteen species of bats (Chiroptera) from Kenya and Rhodesia, Life Sci. Occas. Pap., Volume 27 (1975), pp. 1-16

[47] W. Bergmans W Taxonomy and biogeography of African fruit bats (Mammalia, Megachiroptera). 1. General introduction Material and Methods Results the genus Epomophorus Bennett 1836, Beaufortia, Volume 38 (1988), pp. 75-146

[48] Phylogeography: the history and formation of species (J.C. Avise, ed.), Harvard Univ. Press, Cambridge, 2000

[49] D.J. Funk; K.E. Omland Species-level paraphyly and polyphyly: frequency, causes, and consequences, with insights from animal mitochondrial DNA, Ann. Rev. Ecol. Evol. Syst., Volume 34 (2003), pp. 397-423

[50] G.D.O. Lowe; F.G.R. Fowkes; J. Dawes; P.T. Donnan; S.E. Lennie; E. Housley Blood-viscosity, fibrinogen, and activation of coagulation and leucocytes in peripheral arterial-disease and the normal population in the Edinburg artery study, Circulation, Volume 87 (1993), pp. 1915-1920

[51] F.M. Van’t Hooft; S.J.F. von Bahr; A. Silveira; A. Iliadou; P. Eriksson; A. Hamsten Two common, functional polymorphisms in the promoter region of the beta-fibrinogen gene contribute to regulation of plasma fibrinogen concentration, Arterioscl. Thromb. Vasc. Biol., Volume 19 (1999), pp. 3063-3070

[52] X.F. Lu; H.J. Yu; X.Y. Zhou; L.Y. Wang; J.F. Huang; D.F. Gu Influence of fibrinogen beta-chain gene variations on risk of myocardial infarction in a Chinese Han population, Chinese Med. J., Volume 121 (2008), pp. 1549-1553

[53] R.E. Carpenter Flight physiology of intermediate sized fruit-bats (Family: Pteropodidae), J. Exp. Biol., Volume 120 (1986), pp. 619-747

[54] K. Andersen Catalogue of the Chiroptera in the Collection of the British Museum, Megachiroptera, 1, Trustees British Museum (Natural History), London, 1912

[55] D.R. Rosevear The bats of West Africa, British Museum (Natural History), London, 1965


Commentaires - Politique


Ces articles pourraient vous intéresser

Comparative phylogeography of African fruit bats (Chiroptera, Pteropodidae) provide new insights into the outbreak of Ebola virus disease in West Africa, 2014–2016

Alexandre Hassanin; Nicolas Nesi; Julie Marin; ...

C. R. Biol (2016)


The comparative phylogeography of fruit bats of the tribe Scotonycterini (Chiroptera, Pteropodidae) reveals cryptic species diversity related to African Pleistocene forest refugia

Alexandre Hassanin; Souraya Khouider; Guy-Crispin Gembu; ...

C. R. Biol (2015)


Description of a new bat species of the tribe Scotonycterini (Chiroptera, Pteropodidae) from southwestern Cameroon

Alexandre Hassanin

C. R. Biol (2014)