Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Genetic, Cytogenetic and Morphological Trends in the Evolution of the Rhodnius (Triatominae: Rhodniini) Trans-Andean Group

  • Sebastián Díaz,

    Affiliation Grupo BCEI, Universidad de Antioquia UdeA, Medellin, Antioquia, Colombia

  • Francisco Panzera,

    Affiliation Sección Genética Evolutiva, Facultad de Ciencias, Universidad de la República, Montevideo, Montevideo, Uruguay

  • Nicolás Jaramillo-O,

    Affiliation Grupo BCEI, Universidad de Antioquia UdeA, Medellin, Antioquia, Colombia

  • Ruben Pérez,

    Affiliation Sección Genética Evolutiva, Facultad de Ciencias, Universidad de la República, Montevideo, Montevideo, Uruguay

  • Rosina Fernández,

    Affiliation Sección Genética Evolutiva, Facultad de Ciencias, Universidad de la República, Montevideo, Montevideo, Uruguay

  • Gustavo Vallejo,

    Affiliation Laboratorio de Investigación en Parasitología Tropical, Universidad del Tolima, Ibagué, Tolima, Colombia

  • Azael Saldaña,

    Affiliation Instituto Conmemorativo Gorgas de Estudios de la Salud (ICGES), Ciudad de Panamá, Panamá, Panamá

  • Jose E. Calzada,

    Affiliation Instituto Conmemorativo Gorgas de Estudios de la Salud (ICGES), Ciudad de Panamá, Panamá, Panamá

  • Omar Triana,

    Affiliation Grupo BCEI, Universidad de Antioquia UdeA, Medellin, Antioquia, Colombia

  • Andrés Gómez-Palacio

    amgomezpa@gmail.com

    Affiliation Grupo BCEI, Universidad de Antioquia UdeA, Medellin, Antioquia, Colombia

Abstract

The Rhodnius Pacific group is composed of three species: Rhodnius pallescens, R. colombiensis and R. ecuadoriensis, which are considered important vectors of trypanosomes (Trypanosoma cruzi and T. rangeli) infecting humans. This group is considered as a recent trans-Andean lineage derived from the widespread distributed sister taxa R. pictipes during the later uplift of northern Andes mountain range. The widest spread species R. pallescens may be a complex of two divergent lineages with different chromosomal attributes and a particular biogeographical distribution across Central America and Colombia with several southern populations in Colombia occupying the same sylvatic habitat as its sister species R. colombiensis. Although the taxonomy of Rhodnius Pacific group has been well studied, the unresolved phylogenetic and systematic issues are the target of this paper. Here we explore the molecular phylogeography of this species group analyzing two mitochondrial (ND4 and cyt b) and one nuclear (D2 region of ribosomal 28S gene) gene sequences. The molecular analyses suggest an early divergence of the species R. ecuadoriensis and R. colombiensis, followed by a recent expansion of R. pallescens lineages. The phylogenetic relationship between sympatric R. pallescens Colombian lineage and R. colombiensis was further explored using wing morphometry, DNA genome size measurements, and by analyzing chromosomal behavior of hybrids progeny obtained from experimental crosses. Our results suggest that the diversification of the two R. pallescens lineages was mainly influenced by biogeographical events such as (i) the emergence of the Panama Isthmus, while the origin and divergence of R. colombiensis was associated with (ii) the development of particular genetic and chromosomal features that act as isolation mechanisms from its sister species R. pallescens (Colombian lineage). These findings provide new insights into the evolution of the Rhodnius Pacific group and the underlying biological processes that occurred during its divergence.

Introduction

The subfamily Triatominae (Hemiptera: Reduviidae) includes 144 species of hematophagous insects classified in 5 tribes and 15 genera [1][5]. They are vectors of the protozoan parasite Trypanosoma cruzi, causative agent of American Trypanosomiasis or Chagas disease, one of the most important parasitic diseases in Latin America [6].

One prominent Triatominae tribe, the Rhodniini, is comprised of two genera: Rhodnius (18 species) and Psammolestes (three species) [2][5], a few species of which are targets of several vector control initiatives in Andean and Central American countries [6]. Within this tribe are the pictipes and robustus lineages that are recognized by molecular, biochemical, morphometric and biogeographical attributes [7], [8]. The pictipes lineage gave rise to the trans-Andean (pallescens) and Amazonian (pictipes) species groups, while the robustus lineage diversified within Amazonia and spread to neighbouring ecoregions (Orinoco, Cerrado-Caatinga-Chaco, and Atlantic Forest) [9]. The pallescens group, composed of R. pallescens, R. colombiensis and R. ecuadoriensis, is also named the trans-Andean or Pacific group because its geographic distribution is basically restricted to coastal regions of Ecuador, Colombia, Panamá and Costa Rica. Rhodnius ecuadoriensis is the most southerly-distributed species of the Pacific group and is restricted to southern Ecuador and northern Peru. This species occupies palms trees (Phytelephas aequatorialis) and human dwellings [10]. Rhodnius ecuadoriensis is geographically separated from R. pallescens and R. colombiensis by the Andean mountains and by the pluvial forests of the Colombian Pacific coast [11]. Rhodnius pallescens inhabits palms trees, such as Attalea butyracea and Cocos nucifera, and is widely distributed across Central America and Colombia [11] in different climatic conditions and ecological zones [12], [13]. The most recently described species R. colombiensis [14] is completely a sylvatic species, and it is restricted to the inter Andean valley of Magdalena River in central Colombia. Although both R. colombiensis and R. pallescens species have been found in the same eco-geographical region of central Andean valleys in Colombia as well as inhabiting the same palm tree species (A. butyracea) [15], natural hybrids have not been reported. Due to the lack of knowledge concerning the inter-fertility between these sympatric species, chromosomal analyses of hybrids obtained from experimental crosses could help to identify its reproductive limits.

The genetic diversity within the Pacific group across its geographic distribution is only beginning to be ascertained, and apparent contradictory phylogenetic relationships among its members [16][18] suggest that further systematic studies are still needed.

Intraspecific differences of R. pallescens were accessed by cytogenetic, wing morphometric and molecular analysis of partial cytochrome b (cyt b) gene [19], [20]. Two divergent lineages are differentially distributed in Colombia (termed as R. pallescens I) and Central America (termed as R. pallescens II) as a consequence of both evolutionary diversification and environmental influence [19]. In R. ecuadoriensis, the analysis of mitochondrial cytochrome b gene [10], [21] and in the chromosome location of 45S rDNA cluster [22] suggested that Peruvian and Ecuadorian insects represent discrete populations or even incipient species [10], [21].

Although the Pacific group is unquestionably a monophyletic clade, several genetic incongruences among its species were detected by isoenzyme profiles [7] and nucleotide sequences comparisons of mitochondrial genes [16][18].

Rhodnius colombiensis was initially described as a sylvatic R. prolixus form occupying palms trees in central Andean valleys region of Colombia [7]. Its taxonomic status as a separate species was confirmed by isoenzyme and molecular analyses [14], [23]. The first assessment in phylogenetic reconstruction of Rhodniini tribe including the Pacific group`s species was performed by molecular and morphometric studies [7]. Isoenzymatic analysis of twelve enzyme systems as well as heads and wing measurements was performed in 13 R. ecuadoriensis specimens, 11 R. pallescens and 8 R. colombiensis (thought as “sylvatic R. prolixus” in that time) [7]. Cladograms inferred from isoenzyme alleles synapomorphies as well as in Mahalanobis distances derived from morphological measurements were congruent supporting the Pacific clade, but in this case a basal branch was observed for R. pallescens whereas R. ecuadoriensis and R. colombiensis were grouped in a derivate clade [7]. However, later phylogenetic and phylogeographic analyses using several mitochondrial genes (i.e. cytochrome b – cyt b or the large subunit ribosomal RNA - 16S) showed distinct phylogenetic arrays. A basal R. pallescens branch and derivate R. ecuadoriensis - R. colombiensis clade was observed in phylogenetic analysis based on cyt b gene [16], whereas basal R. ecuadoriensis branch and derivate R. pallescens - R. colombiensis clade was observed in phylogeographic analyses using 16S gene [17], [18]. Besides of the little information about chromosomal and genetic attributes of R. colombiensis, its phylogenetic relationship with R. ecuadoriensis and with recently identified R. pallescens lineages, as well as a credible hypothesis about its origin or evolutionary divergence with other Pacific species remained unresolved. Therefore further genetic analyses as well as new evolutionary hypotheses about Pacific group were still needed to understand the evolutionary trends that shape the phylogeographic and evolutionary landscape of Rhodnius Pacific group.

So far at least two theories about Pacific group origin (including R. colombiensis origin) have been proposed [8], [9]. Although both indicate a monophyletic origin from a widespread generalist species similar to R. pictipes, the first theory [8] suggests northwestern origin of the Pacific group, whereas the second [9] a southeastern origin.

In the northern origin theory is thought a basal R. pallescens clade was dispersed from northern Colombia to Central American, and across of the Caribbean coast and Andean valleys of Colombia, to eastern Ecuador and northern Peru giving thus origin to R. pallescens lineages, R. ecuadoriensis and R. colombiensis [8].

The second, and more recent theory about the evolution of the Rhodnius Pacific group suggests it may have occurred by a combination of adaptive radiation and vicariant processes [9], [10], [18]. This group is considered a lineage that derived from an ancestral population close to R. pictipes that reached the western side of the Andes range from the eastern Orinoco plains during the late Miocene (∼6 Mya) [10]. The rise of the Andean mountains during the Pliocene (∼5 Mya) split that population into two main clades: the “Colombian cluster”, comprising the ancestral forms of R. pallescens, and R. colombiensis. Rhodnius ecuadoriensis originated from an isolated pocket in the south, which adapted to new ecotopes [9].

In order to describe several chromosomal, genetic and morphological attributes of the poorly studied species R. colombiensis as well as clarify the phylogenetic picture and the species limits within the Pacific group, we performed cytogenetic, morphometric and phylogeographic studies based on the analysis of mitochondrial cyt b and ND4 genes and nuclear 28S gene of individuals belonging to R. pallescens lineages, R. ecuadoriensis and R. colombiensis taxa. Our findings provide new insights into the underlying biological processes that shape the evolution of this important group and improve the systematic picture within the genus Rhodnius.

Materials and Methods

Samples, DNA Extraction and PCR-amplification

No specific permissions were required for insect collections performed in this work, and did not involve endangered or protected species. A total of 67 specimens representing 18 locations of R. pallescens from Colombia and Panama, 21 individuals of R. colombiensis from 4 locations, and 8 specimens of R. ecuadoriensis from 3 locations of Ecuador were included in our analyses (Table 1, Figure 1). Individuals were collected between 1997–2013. R. pallescens and R. ecuadoriensis identification was performed according to the morphological keys proposed by Lent and Wygodzinsky [24], and keys proposed for R. colombiensis by Moreno et al., [14].

thumbnail
Figure 1. Collection sites of Rhodnius species of the Pacific group.

Numbers indicated origin sites detailed in Table 1. Color indicates the species/lineages of the Pacific group: R. pallescens I in red; R. pallescens II in blue; R. colombiensis in yellow; and R. ecuadoriensis in green. Outgroup species R. robustus and R. pictipes-like are represented in black triangles. Numbers match with locality numbers in Table 1.

https://doi.org/10.1371/journal.pone.0087493.g001

thumbnail
Table 1. Geographic origins and number of specimens of the Pacific group used in this study.

https://doi.org/10.1371/journal.pone.0087493.t001

Genomic DNA was obtained from leg or thorax muscles [25]. For each specimen a 631-bp fragment of nicotinamide adenine dinucleotide dehydrogenase 4 (ND4) gene was PCR-amplified using primers ND4-F (5′-TCAACATGAGCCCTTGGAAG -3′) and ND4-R (5′-TAATTCGTTGTCATGGTAATG -3′) [26]; and a 682-bp fragment of cytochrome B (cyt b) gene was PCR-amplified using primers CYTB7432 (5′-GGACGWGGWATTTATTATGGATC-3′) and CYTB7433 (5′-GCWCCAATTCARGTTARTAA-3′) [27]. PCR reactions for both mitochondrial genes were conducted in a final volume of 35 µl using 30-ng of DNA templates, 1X PCR buffer (0.1 M Tris–HCl, 0.5 M KCl, and 0.015 M MgCl2, pH 8.3), 250- µM dNTP, 0.016-µM of each primer, 35-mM MgCl2 and 2 U of Taq DNA polymerase. The fragments were amplified with the following thermal cycling conditions: 95°C for 5 min; 35 cycles of 94°C for 30 s, 50°C for 30 s, and 72°C for 60 s; 72°C for 10 min.

For the D2 variable region of the 28S rDNA gene (D2-28S) a fragment of 434-bp was PCR-amplified using primers D2F (5′-GCGAGTCGTGTTGCTTGATAGTGCAG-3′) and D2R, (5′-TTGGTCCGTGTTTCAAGACGGG-3′) [28]. PCR reactions were conducted in a final volume of 35 µl using 30-ng of DNA templates, 1X PCR buffer (0.1 M Tris–HCl, 0.5 M KCl, and 0.015 M MgCl2, pH 8.3), 250- µM dNTP, 0.025-µM of each primer, 3-mM MgCl2 and 2 U of Taq DNA polymerase (Promega®). After an initial denaturation of 95°C for 5 min, PCR reactions were 35 cycles at 95°C for 30 s, 60°C for 30 s, and 72°C for 30 s, followed by a final extension of 72°C for 7 min [29]. Amplicons were sent to Macrogen Inc., Korea to be purified and sequenced in both directions.

Nucleotide Sequence Alignment and Diversity

All sequences used here are available in GenBank (accession numbers KC543506–66, and GQ850481, FJ229357–60, JQ686674–89 for those reported in Gómez-Palacio et al., (2012)). Alignments were made using CLUSTALW algorithm [30] implemented in Bioedit 7.0.5 [31]. ND4 and cyt b nucleotide sequences from specimens R. pictipes-like (form resembling R. pictipes; NJO, unpublished) collected in the Sierra Nevada de Santa Marta (SNSM) mountain in northern Colombia were also included in analyses. Colombian southern form of R. robustus [32] was included as outgroup in phylogenetic analyses (see below).

As a measure of gene diversity within the Pacific group, several parameters such as haplotype number (h), haplotype diversity (Hd), and nucleotide diversity (π) were calculated within species/lineages by using DNAsp 5.10.01 software [33]. Genetic distances among species/lineages were estimated for D2-28S, ND4 and cyt b genes based on the substitution model of K 2-p [34] as reported for Rhodnius sibling species [35], using MEGA 5.05 software [36].

Phylogenetic Analysis

The combined data set of ND4 and cyt b (1313-bp) nucleotide sequences as well as the D2-28S fragment (434-bp) were used to infer the partitioned best-fitting model using the Akaike information criterion [37], as implemented in jMODELTEST 0.1.1 [38]. A Bayesian Metropolis coupling Markov Chain Monte Carlo (MC3) approach was implemented in BEAST 1.7.4 package [39]. Model parameters (base frequencies, transition/transversion ratio, rate variation shape parameter) were derived empirically, and two chains run for 10×109 generations, with a sampling frequency of 10,000 generations. Finally by using as criteria the maximum sum of posterior probabilities (Maximum clade credibility) a final topology was chosen after discarding a burn-in of 10%.

Maximum Parsimony (MP) trees were constructed using TNT 1.1 [40]; shortest trees were found via a traditional heuristic search with tree bisection-reconnection (TBR) branch swapping saving 10 trees per replication, replacing existing trees. Statistical support for clades in the phylogenetic tree was assessed by the standard bootstrap method [41] with 100,000 replicates. Topologies were edited with the software FigTree 1.3.1. [42].

Divergence Time Estimation

Based-tree topology a Maximum Likelihood (ML) test for molecular clock (it means that all tips of the tree are equidistant from the root of the tree) was performed using MEGA 5.05 [36] giving as result the rejection of a strict molecular clock model for the dataset. Thus a relaxed molecular clock model, which allows branch lengths to vary following an uncorrelated log-normal distribution [43] was performed using as prior assumption the independence of the clock models for both ND4 and cyt b data set using BEAST 1.7.4 package [39]. We considered a range of 15±5 Mya (before to the first rise of the Andes, dated about 10±7 Mya; [44]) as a calibration point for separation of trans-Andean Rhodnius species as inferred from formation of the Pebas System, a great Amazonian wetland that occurred in early and middle Miocene [44].

Haplotype Network

Median-Joining haplotype network [45] of D2-28S fragment as well as combined ND4 and cyt b genes was performed to examine inter-haplotype relationships among species/lineages using Network 4.6.0.0 (http://www.fluxus-engineering.com). Haplotype networks were built using default parameters (equal character weight = 10; epsilon value = 10; transversions/transitions weight = 1∶1 and connection cost as a criterion).

Total DNA Content Measured by Flow Cytometry in R. colombiensis

We measured DNA content only for R. colombiensis because R. ecuadoriensis and R. pallescens genome sizes were already known [19], [46]. The total nuclear DNA content of 13 specimens of R. colombiensis came from the type locality (Coyaima-Tolima) was measured from gonadal cells of male specimens as previously reported [47]. The cell DNA content was measured on an EPICS XL-MCL flow cytometer (Coulter Electronics, Hialeah, FL, USA) with an air-cooled argon-ion laser set at 488 nm and 15 mW. Propidium fluorescence (FL3), proportional to DNA content, was collected through a 650-nm DL dichroic filter fitted with a 625-nm BP band-pass filter. The DNA content in single cells was determined from FL3 linear histograms. For each sample, information for a minimum of 10,000 nuclear events was acquired using the System II software program (Beckman Coulter Inc., Brea, CA, USA). To evaluate the DNA content in picograms of DNA, a sample of normal human lymphocytes was fixed in ethanol/acetic acid and used as the standard reference (2C = 7.0 pg of DNA according to the Animal Genome Size Database (http://www.genomesize.com/). The absolute DNA amount was calculated with the ratio of the mean channel of the insect haploid peak to the mean channel of the human lymphocyte diploid G0/G1 peak.

Experimental Crosses between Rhodnius pallescens and R. colombiensis, and Cytogenetic Analyses of F1 Progeny

Experimental crosses were performed between R. colombiensis and R. pallescens I (Colombian lineage). Individuals of R. colombiensis came from the type locality (Coyaima-Tolima) and individuals of R. pallescens lineage I were from Colombia (San Onofre-Sucre).

Crosses were made using three couples in each direction (male R colombiensis X female R. pallescens, or female R. colombiensis X male R. pallescens). All bugs used for the crosses were collected as nymphs; recently emerged (virgin) adults were place together in plastic vials with folded filter paper.

Testes from 6 freshly killed male hybrids (F1) were fixed in an ethanol–acetic acid mixture (3∶1) and stored at −20°C. Chromosome preparations and C-banding technique were performed as previously reported [48]. Females were not studied because meiotic stages are not usually observed in their ovaries.

Geometric Morphometric Analysis

We identified morphological differences between R. pallescens lineages and R. colombiensis by performing two analyses, one based on wing size and the other for wing shape. Right wings were dissected and mounted using standard techniques as reported elsewhere [19] and photographed with a Nikon 990 digital camera fitted to a Nikon SMS 800 stereomicroscope. The wings were placed in the center of the visual field to reduce the risk of optical distortion. Ten landmarks were selected and the geometric coordinates of each landmark were digitalized using tpsDIG 2.16 [49].

To compare wing size among species/lineages, the “centroid-size” was used as isometric estimator of overall size derived from coordinate data [50]. It is defined as the square root of the sum of squared distances between the center of the configuration of landmarks and each individual landmark [50]. Differences in size were tested by one-way ANOVA and a post-hoc pairwise comparison was performed based on Tukey's HSD (honestly significant difference) test.

The Generalized Procrustes Analysis (GPA) superimposition algorithm implemented in the tpsRelw 1.11 [51] was used to obtain the wing shape variables. Shape variation was analyzed using the principal components of shape variables or relative warps. Differences in shape were tested by MANOVA. A post-hoc analysis by a pairwise Hotelling test (with Bonferroni correction) was performed when the MANOVA showed a significant overall difference between groups. A principal component analysis was performed to produce a scatter plot of specimens along the first two component axes, producing maximal and second to maximal separation between all groups. Additionally, a discriminant analysis (DA) reclassification test based on the first two discriminant functions (DF) to each observation was designed and the Mahalanobis distances among the shape’s centroids for each pair species/lineage (see Table 1) were compared. Statistics were performed using the free software PAST 1.94b [52], and the statistical significance of pairwise comparisons was tested by a null model using 1000 permutations.

Results

Sequence Variation and Genetic Distances

In the complete dataset a total of 121, 150 and 12 variable sites (S) were observed for ND4, cyt b and D2-28S genes, respectively, as well as 35, 33 and 8 haplotypes (h). Congruent gene diversity values for both ND4 and cyt b genes were detected within species/lineage Pacific group samples (Table 2). Nucleotide (π) diversity showed the lowest values in R. colombiensis (π = 0.004 for both ND4 and cyt b) and the highest within R. pallescens (π = 0.012 and 0.015), followed by R. ecuadoriensis (π = 0.011 and 0.009 for ND4 and cyt b, respectively). Haplotype diversity (Hd) was marginally similar among samples studied (Table 2). For D2-28S gene, 12 variable sites out of the 434-bp were observed across species/lineages, harboring the eight haplotypes (Table S1). Only one haplotype was species-specific to R. colombiensis, while the most frequent D2-28S haplotype was shared for both R. pallescens lineages and R. ecuadoriensis, but not by R. colombiensis (Table S1).

thumbnail
Table 2. Summary of genetic diversity indices Notations: h: number of haplotypes; Hd (± SD): haplotype diversity (standard deviation); π (± SD): nucleotide diversity (standard deviation).

https://doi.org/10.1371/journal.pone.0087493.t002

K 2-p-based genetic distances (Table 3) were moderately high between R. pallescens lineages (d = 0.041 and 0.051 for ND4 and cyt b genes respectively). Similar values were observed between R. pallescens I vs. R. colombiensis (d = 0.034 and 0.065 for ND4 and cyt b, respectively), and between R. pallescens II vs. R. colombiensis (d = 0.045 and 0.062 for ND4 and cyt b, respectively). The highest genetic distance was estimated between R. pallescens II and R. ecuadoriensis (d = 0.102 and 0.133 for ND4 and cyt b, respectively). For the D2-28S gene, estimated K 2-p-based genetic distances were remarkably low between R. pallescens lineages (d = 0.001), and between those with R. ecuadoriensis (d = 0.001, for both comparisons); but comparatively higher when R. colombiensis was included (R. pallescens I vs. R. colombiensis d = 0.005; R. pallescens II vs. R. colombiensis d = 0.006; and R. ecuadoriensis vs. R. colombiensis d = 0.005).

thumbnail
Table 3. Pairwise K 2-p-based genetic distance (upper right) and cyt b (lower left) gene fragments.

https://doi.org/10.1371/journal.pone.0087493.t003

Phylogenetic Results and Divergence

Due to a low number of variable sites observed in D2-28S gene, no resolute topologies were obtained for this marker (data not shown). For the combined ND4 and cyt b dataset partitioned substitution GTR+G and HKY+I models were used respectively. MP strict consensus tree showed 673 steps, with a consistency index of 0.74 and a retention index of 0.95 (Figure S1). Both Bayesian and MP analyses showed identical topologies; hence only the Bayesian tree is shown (Figure 2). Four well-supported monophyletic clades (posterior probabilities = 1.0, and bootstrap values >0.9) were observed harboring the three species and both R. pallescens lineages (Figure 2). Basal clade was composed of R. ecuadoriensis followed by a clade belonging to R. colombiensis, and both lineages of R. pallescens (Figure 2). Based on a relaxed clock model Bayesian analysis indicated that R. ecuadoriensis clade diverged from the ancestral Pacific group at 11.5 (±3.9 Mya); followed by the divergence of R. colombiensis (10±3.4 Mya); and R. pallescens lineages (6.1±2.1 Mya).

thumbnail
Figure 2. Bayesian consensus tree for the combined ND4 and cyt b genes.

Posterior probabilities/bootstrap support (from MP analysis, Figure S1) for representative nodes is shown. Color of clades indicate the Pacific group species/lineages: R. pallescens I in red; R. pallescens II in blue; R. colombiensis in yellow; and R. ecuadoriensis in green.

https://doi.org/10.1371/journal.pone.0087493.g002

mtDNA and rDNA Haplotype Network

The haplotype network based on 41 haplotypes of the combined ND4 and cyt b dataset (Table S2) shows four clearly separated groups belonging to the three species, plus the two R. pallescens lineages (Figure 3A). Moreover for D2-28S gene a star-shape haplotype network was observed (Figure 3B). The most frequent D2-28S haplotype included all R. pallescens I and R. ecuadoriensis individuals (32 and 6 respectively), and most R. pallescens II specimens (15 individuals) (Figure. 3B and Table S1). From this haplotype derived the remaining haplotypes of R. pallescens II and R. colombiensis (Figure 3B). This last species differs by two mutational steps from the principal haplotype (Figure 3B).

thumbnail
Figure 3. Haplotype network of (A) combined ND4 - cyt b genes, and (B) D2-28S gene of the Pacific group species/lineages.

The size of the circles indicates the haplotype frequency. Mutation number for every node (>5 for mtDNA and >1 for D2-28S gene) are detailed in (A) and (B) in red. Color of the pie indicates the species/lineage of the Pacific group: R. pallescens I in red; R. pallescens II in blue; R. colombiensis in yellow; and R. ecuadoriensis in green. Line in bold in (B) shows two mutation steps separating R. colombiensis from the central haplotype.

https://doi.org/10.1371/journal.pone.0087493.g003

Genome Size of R. colombiensis

Based on DNA flow cytometric profiles obtained (Figure S2), genome size measure in R. colombiensis indicated an amount of haploid DNA content (C value) of 0.58±0.01 pg.

Experimental Crosses between Rhodnius pallescens and R. colombiensis, and Cytogenetic Analyses of F1 Progeny

Crosses between R. pallescens lineage I (males) and R. colombiensis (females) resulted in a F1 progeny that include both adult males and females. In the reciprocal cross R. pallescens lineage I females and R. colombiensis males, the eggs did not eclose and thus no progeny was obtained.

All male hybrids had the same diploid chromosome number as the progenitors (2n = 22) consisting of 20 autosomes and a pair of sex chromosomes (XY). Mitotic chromosomes have C-blocks or were completely euchromatic (Figure 4). Chromosomes were rather similar in size and it was not possible to establish if they came from R. colombiensis or R. pallescens. Early meiotic cells appeared normal and had a heteropycnotic chromocenter constituted by the associated XY sex chromosomes plus autosomal heterochromatin (Figure 4a). Later meiotic stages had several meiotic anomalies that affected spermatogenesis to varying degrees. Chromosomal pairing was greatly altered during first meiotic division. Some autosomes synapsed and formed bivalents, while others remained as univalents (Figures 4b and 4c). The frequencies of bivalents and univalents varied among different individuals and between cells of the same specimen. The number of cells that progress through meiosis was drastically reduced and the resulting cells are expected to have unbalanced genomes as a consequence of missegregation during first meiotic division.

thumbnail
Figure 4. Male meiosis of hybrids (F1 progeny) from experimental crosses by C-banding technique.

The crosses were made between R. pallescens (males) and R. colombiensis (females). (A) Early meiotic prophase (diffuse stage) showing a heteropycnotic chromocenter constituted by XY sex chromosomes plus one autosomal C-heterochromatic dot. Other C-dots dispersed in the nucleus (arrows) are observed. (B) Late diplotene or diakinesis and (C) Metaphase I showing a variable number of chromosomes, instead of twelve (ten autosomal pairs plus 2 sex chromosomes) observed in normal cells. Some of autosomes are bivalents (asterisks) and other are univalents (arrowheads).

https://doi.org/10.1371/journal.pone.0087493.g004

Wing Size and Shape Comparison among R. pallescens Lineages and R. colombiensis

Significant differences in size were detected among all species/lineages (F = 24.70, p<0.001). After a Tukey HSD post-hoc test, no significant differences (p>0.01) in size were identified between the smallest insects R. colombiensis and R. pallescens I, but both of them were significant differentiated (p<0.01) from the largest ones R. pallescens II (Figure 5 and Table 4).

thumbnail
Figure 5. Wing size differences of R. colombiensis and R. pallescens.

There were used specimens of the two Colombian-Central American R. pallescens lineages (termed as R. pallescens I and II respectively). The arithmetic median can be observed as a line that divides the boxes into two. The ends of the boxes correspond to the 25% and 75% quantiles; the vertical lines show the maximum and minimum value of the centroid size distribution. * Indicate significant differences (p<0.001).

https://doi.org/10.1371/journal.pone.0087493.g005

thumbnail
Table 4. ANOVA and Tukey HSD test of pairwise comparisons.

https://doi.org/10.1371/journal.pone.0087493.t004

The first two axes of relative warps explained 40.6% and 12.5% of the variation, respectively, corresponding to 53.1% of the total variation of the specimens’ wing shape, dismissing the use of other axes, and shape differences were detected between R. pallescens lineages and R. colombiensis (Figure 6). The MANOVA test allowed detecting significant differences in wing shape among all species/lineages (F = 18.70, p<0.001) (Table 5). The Hotelling post-hoc analysis showed that significant differences in wing shape were detected between all pairwise species/lineages (Table 5). Reclassification of specimens based on discriminant factors resulted in the correct classification of 96.12% of the individuals to their attributed species/lineage. Reclassification was perfect for the individuals of R. colombiensis (100%), followed by R. pallescens I (97.9%), and R. pallescens II (92.7%) (Table 6).

thumbnail
Figure 6. Morphometric analysis of male wings in R. pallescens lineages and R. colombiensis.

The polygons represent the shape of wings projected on the first (x-axis) and second relative warp (y-axis), which were derived from a relative warp analysis. RW1 explains 40.6% of the variance while RW2 explains 12.5%. For easy visualization of populations the lines connect the most external individuals and filled color represents species/lineage: R. colombiensis in yellow; R. pallescens I in red and R. pallescens II in blue.

https://doi.org/10.1371/journal.pone.0087493.g006

thumbnail
Table 5. MANOVA and Hotelling pairwise comparisons based on shape variation.

https://doi.org/10.1371/journal.pone.0087493.t005

thumbnail
Table 6. Reclassification analysis of individuals based on wing shape.

https://doi.org/10.1371/journal.pone.0087493.t006

Discussion

Phylogenetic Trends in the Pacific Group Species

Phylogeny supports that R. colombiensis is a sister taxa of R. pallescens and that both have diverged from an ancestral clade that includes R. ecuadoriensis. We suggest additional timing for the Pacific group origin according to relaxed clock estimate divergence time by using as calibration point the formation of the Pebas wetland dated at middle Miocene [44], which is thought to have the most important role in cis and trans-Andean higher lineages diversification in Rhodniini. We estimate that R. ecuadoriensis and R. colombiensis divergence occurred in late Miocene (∼11 to 7 Mya), a period when the Northern Andes had reached no more than half of their modern elevation [53]; and the diversification of the R. pallescens lineages occurred during the latest Miocene-early Pliocene, being contemporary to the rapid uplift and extensive emergence of the Central American isthmus [54]. Genetic distances of ND4 and cyt b genes indicates that R. ecuadoriensis is the most divergent species, while similar values are evidenced between R. pallescens lineages, and between them with R. colombiensis, indicating that molecular differentiation between R. pallescens and R. colombiensis species is almost as recent as the arising two lineages with R. pallescens. This assumption was also supported by molecular relaxed clock results shown here.

Several genetic and phenotypic attributes of R. ecuadoriensis, R. colombiensis and R. pallescens explored here suggest that the Pacific group evolution involved at least two macro-evolutionary processes that gave rise to the current phylogeographic distribution: i) a wide geographical dispersion (from south to north) along the northern Andes during uplifting that separated R. ecuadoriensis and R. colombiensis basal populations, and ii) the formation of the Panamian isthmus that extended the spread of R. pallescens. These hypotheses support the southeastern origin theory of Pacific group that is suggested as a combination of adaptive radiation and vicariant processes [9], [10], [18].

Taxonomic Validity of the Pacific Group Species

Specific status of R. colombiensis was initially proposed by isoenzymatic differentiation and by comparison of morphometric and genitalic structures [7], [14]. However, data about its evolutionary history and genetic relationship with its closer conspecific species R. pallescens were lacking. Here, we explored for the first time the biological differentiation of R. colombiensis within the Pacific group using molecular and morphometric analyses. Our results support the taxonomic status of R. colombiensis as a R. pallescens sibling species. Genetic distance estimates of mtDNA among Pacific group species (ranged between d = 0.034 to 0.133) were similar to those observed in cis-Andean species, specifically in cryptic species belonging to robustus lineage (that ranged between d = 0.023 to 0.072) [35]. The chromosomal and genome size measurements suggest relevant genomic arrangements could be involved in R. colombiensis speciation, suggesting distinct evolutionary trends were placed in the Pacific group evolution.

Chromosomal traits and genome size in Triatominae had been broadly used in understanding the taxonomy, diversification and genome evolution in several species of this subfamily [46]. The three species of Pacific group possess similar chromosomal characteristics, such as the same number of autosomes (20), sex mechanism (XY in males and XX in females) and the presence of several autosomal pairs with small terminal C-dots [20], [55]. For these reasons it is very difficult to differentiate these three species by standard cytogenetic analyses but the chromosomal data presented here suggested that striking structural chromosome rearrangements occurred during the divergence of these species. The haploid DNA content for R. pallescens was 0.73±0.04 pg [19], similar to that observed in R. ecuadoriensis (0.72 pg) [47]. Our results revealed that R. colombiensis has the lowest value of DNA content (0.58±0.01 pg) and suggest that DNA loss has taken placed during the evolution of this species. Recent analyses about the chromosomal location of ribosomal genes (45S rDNA clusters) also indicated a striking genome differentiation among these species. Fluorescent in situ hybridization assays shown that ribosomal rDNA clusters are located in both XY sex chromosomes in R. pallescens and in Ecuadorian populations of R. ecuadoriensis, while that they are located only in the X chromosome of R. colombiensis as well as in some specimens of the Peruvian R. ecuadoriensis populations [22], [55].

The mechanisms that limit gene flow between populations and species can be studied by performing experimental crosses and analyzing the progeny [56]. Hybridization studies have been implemented in several species of the Triatoma and Meccus genera (for review see [57][59]. The present study constitutes the first cytogenetic analysis of experimental hybrids among Rhodnius species. The analysis of the crosses between R. colombiensis and R. pallescens reveals the existence of both pre-zygotic and post-zygotic reproductive barriers. The crosses R. pallescens lineage I (females) and R. colombiensis (males) did not produce progeny, indicating that non-fertile eggs were obtained by chance or supporting the possible existence of a pre-zygotic isolation mechanism (e.g. incompatibilities between genitalia of both species). Crosses between R. pallescens lineage I (males) and R. colombiensis (females) overcome pre-zygotic barriers but produce infertile F1 hybrids (post-zygotic reproductive barrier). Sterility was associated with failures in chromosome pairing during meiosis (Figure 4) that led to the production of unbalanced gametes as observed in other interspecific triatomine hybrids [58], [59]. The observed isolation mechanisms explain the lack of natural hybrids although both species are sympatric in the inter Andean valley of Magdalena River and occupy the same sylvatic ecotope (A. butyracea palm trees).

The alteration in the meiotic pairing in the gonad cells of the hybrids between R. pallescens and R. colombiensis also reveals the lack of genetic homology between chromosomes of both species. This result, together with the variation in genome size and rDNA chromosome location between the species [22], indicates that drastic genome rearrangements may have occurred between the species, giving rise to the speciation event in R. colombiensis.

Patterns of morphological variation that involves size or shape dimensions have been often interpreted with regard to their evolutionary importance [60]. Environmental influences (such as elevation and humidity) are well known in several Triatominae species (reviewed in [61]), and while wing size differentiation has been suggested to be strongly influenced (but not exclusively) by ecological attributes of species and populations, wing shape is thought often more affected by genetic and historical attributes of the evolutionary process. Under this consideration, as R. pallescens and R. colombiensis species are sympatric and have ecological similarity, they were expected to have similar wing sizes but large differences in wing shape are evident between them. These results unveil different biological (genetic) and ecological (environmental) pressures influencing morphological diversity within both species.

Conclusions

In summary, we consider that different rates of molecular divergence detected in mtDNA and rDNA sequences; genome size variation and rDNA location in R. colombiensis, the evidence of post-zygotic barriers, and wing shape differentiation between sympatric species R. pallescens and R. colombiensis, indicate that distinct evolutionary trends are drivers of the evolutionary change within the Pacific group. Similar studies should be extended to determine the evolutionary history of the larger pictipes group.

Supporting Information

Figure S1.

Maximum Parsimony tree for the combined ND4 and cyt b genes. Bootstrap support for representative nodes is shown. Color of clades indicate specie/lineage: R. pallescens I in red; R. pallescens II in blue; R. colombiensis in yellow; and R. ecuadoriensis in green.

https://doi.org/10.1371/journal.pone.0087493.s001

(TIF)

Figure S2.

Representative DNA flow cytometric histogram showing the distribution of testis cells from R. colombiensis. Relative intensity (in arbitrary units) of DNA associated PI fluorescence are shown on the x-axis. The corresponding number of cells is displayed on the y-axis. (A) Human polimophonuclear leukocytes and (B) R. colombiensis. Mean of PI fluorescence of M1 peak indicating C-value is shown.

https://doi.org/10.1371/journal.pone.0087493.s002

(TIF)

Table S1.

Multiple alignment of D2-28S gene sequences of Rhodnius Pacific group species. Codes and origins are shown in Table 1. Identical bases are represented by a dot (.). Only variable sites, with sequence positions given above, are shown.

https://doi.org/10.1371/journal.pone.0087493.s003

(DOC)

Table S2.

Haplotype description of combined ND4 and cyt b genes for Rhodnius Pacific group species/lineages.

https://doi.org/10.1371/journal.pone.0087493.s004

(DOC)

Acknowledgments

Special thanks to Juan Carlos Fernandez of Centro de Investigaciones en Microbiología y Parasitología Tropical (CIMPAT), Universidad de los Andes; and Santander’s Secretary of Health (Bucaramanga, Santander, Colombia) for their generous provision of part of specimens used in this study.

Author Contributions

Conceived and designed the experiments: SD AGP FP. Performed the experiments: SD FP OT RP RF NJO AGP. Analyzed the data: SD FP RP OT AGP. Contributed reagents/materials/analysis tools: NJO GV AS JC. Wrote the paper: SD FP AGP. Contributed to the writing of the manuscript: NJO RP RF GV AS JC OT.

References

  1. 1. Ayala J (2009) Una nueva especie de Panstrongylus Berg de Venezuela. (Hemiptera: Reduviidae, Triatominae). Entomotropica 24: 105–109.
  2. 2. Da Rosa J, Rocha C, Gardim S, Pinto M, Mendonca V, et al. (2012) Description of Rhodnius montenegrensis n. sp (Hemiptera: Reduviidae: Triatominae) from the state of Rondonia, Brazil. Zootaxa 3478: 62–76.
  3. 3. Frías-Lasserre D (2010) A new species and karyotype variation in the bordering distribution of Mepraia spinolai (Porter) and Mepraia gajardoi Frías et al (Hemiptera: Reduviidae: Triatominae) in Chile and its parapatric model of speciation. Neotrop Entomol 39: 572–583.
  4. 4. Jurberg J, Rocha DdS, Galvão C (2009) Rhodnius zeledoni sp. nov. afim de Rhodnius paraensis Sherlock, Guitton & Miles, 1977 (Hemiptera, Reduviidae, Triatominae). Biota Neotropica 9: 123–128.
  5. 5. Schofield CJ, Galvão C (2009) Classification, evolution, and species groups within the Triatominae. Acta Trop 110: 88–100.
  6. 6. WHO (2007) Grupo de trabajo científico: Reporte sobre la enfermedad de Chagas.17–20 de abril de 2005. Actualizado julio de 2007. ed Guhl F and Lazdins-Helds.
  7. 7. Dujardin JP, Chavez T, Moreno JM, Machane M, Noireau F, et al. (1999) Comparison of isoenzyme electrophoresis and morphometric analysis for phylogenetic reconstruction of the Rhodniini (Hemiptera: Reduviidae: Triatominae). J Med Entomol 36: 653–659.
  8. 8. Schofield C, Dujardin J-P (1999) Theories on the evolution of Rhodnius. Actualidades Biológicas 21: 183–197.
  9. 9. Abad-Franch F, Monteiro FA (2007) Biogeography and evolution of Amazonian triatomines (Heteroptera: Reduviidae): implications for Chagas disease surveillance in humid forest ecoregions. Mem Inst Oswaldo Cruz 102 Suppl 157–70.
  10. 10. Abad-Franch F, Monteiro FA, Jaramillo O N, Gurgel-Gonçalves R, Dias FB, et al. (2009) Ecology, evolution, and the long-term surveillance of vector-borne Chagas disease: a multi-scale appraisal of the tribe Rhodniini (Triatominae). Acta Trop 110: 159–177.
  11. 11. Galvão C, Carcavallo R, Da Silva D, Jurberg J (2003) A checklist of the current valid species of the subfamily Triatominae Jeannel, 1919 (Hemiptera, Reduviidae) and their geographical distribution, with nomenclatural and taxonomic notes. Zootaxa 202: 1–36.
  12. 12. Gottdenker NL, Calzada JE, Saldaña A, Carroll CR (2011) Association of anthropogenic land use change and increased abundance of the Chagas disease vector Rhodnius pallescens in a rural landscape of Panama. Am J Trop Med Hyg 84: 70–77.
  13. 13. Jaramillo N, Schofield CJ, Gorla DE, Caro-Riaño H, Moreno J, et al. (2000) The role of Rhodnius pallescens as a vector of Chagas disease in Colombia and Panama. Res Rev Parasitol 60: 75–82.
  14. 14. Moreno J, Galvao C, Jurberg J (1999) Rhodnius colombiensis sp. n da Colombia, com quadros comparativos entre estruturas fállicas do gênero Rhodnius stal, 1859 (hemiptera, reduviidae, triatominae). Entomología y Vectores 6: 601–617.
  15. 15. Guhl F, Aguilera G, Pinto N, Vergara D (2007) [Updated geographical distribution and ecoepidemiology of the triatomine fauna (Reduviidae: Triatominae) in Colombia]. Biomedica 27 Suppl 1143–162.
  16. 16. Monteiro FA, Wesson DM, Dotson EM, Schofield CJ, Beard CB (2000) Phylogeny and molecular taxonomy of the Rhodniini derived from mitochondrial and nuclear DNA sequences. Am J Trop Med Hyg 62: 460–465.
  17. 17. Hypsa V, Tietz DF, Zrzavý J, Rego RO, Galvao C, et al. (2002) Phylogeny and biogeography of Triatominae (Hemiptera: Reduviidae): molecular evidence of a New World origin of the Asiatic clade. Mol Phylogenet Evol 23: 447–457.
  18. 18. De Paula AS, Diotaiuti L (2007) Cleber (2007) Systematics and biogeography of Rhodniini (Heteroptera: Reduviidae: Triatominae) based on 16S mitochondrial rDNA sequences. J Biogeogr 34: 699–712.
  19. 19. Gómez-Palacio A, Jaramillo-O N, Caro-Riaño H, Diaz S, Monteiro FA, et al. (2012) Morphometric and molecular evidence of intraspecific biogeographical differentiation of Rhodnius pallescens (Hemiptera: Reduviidae: Rhodniini) from Colombia and Panama. Infect Genet Evol 12: 1975–1983.
  20. 20. Gómez-Palacio A, Jaramillo-Ocampo N, Triana-Chávez O, Saldaña A, Calzada J, et al. (2008) Chromosome variability in the Chagas disease vector Rhodnius pallescens (Hemiptera, Reduviidae, Rhodniini). Mem Inst Oswaldo Cruz 103: 160–164.
  21. 21. Abad-Franch F, Monteiro F (2005) Molecular research and the control of chagas disease vectors. Annals of the Brazilian Academy of Sciences 77: 437–454.
  22. 22. Pita S, Panzera F, Ferrandis I, Galvão C, Gómez-Palacio A, et al. (2013) Chromosomal divergence and evolutionary inferences of Rhodniini based in chromosome location of the ribosomal genes. Mem Inst Oswaldo Cruz 108: 376–382.
  23. 23. Jaramillo C, Montana MF, Castro LR, Vallejo GA, Guhl F (2001) Differentiation and genetic analysis of Rhodnius prolixus and Rhodnius colombiensis by rDNA and RAPD amplification. Mem Inst Oswaldo Cruz 96: 1043–1048.
  24. 24. Lent H, Wygodzinsky P (1979) Revision of the Triatominae (Hemiptera, Reduviidae) and their significance as vectors of Chagas′ disease. Bull Am Mus Nat Hist 163: 123–520.
  25. 25. Collins FH, Mendez MA, Rasmussen MO, Mehaffey PC, Besansky NJ, et al. (1987) A ribosomal RNA gene probe differentiates member species of the Anopheles gambiae complex. Am J Trop Med Hyg 37: 37–41.
  26. 26. Grisales N, Triana O, Angulo V, Jaramillo N, Parra-Henao G, et al. (2010) [Genetic differentiation of three Colombian populations of Triatoma dimidiata (Heteroptera: Reduviidae) by ND4 mitochondrial gene molecular analysis]. Biomedica 30: 207–214.
  27. 27. Lyman DF, Monteiro FA, Escalante AA, Cordon-Rosales C, Wesson DM, et al. (1999) Mitochondrial DNA sequence variation among triatomine vectors of Chagas' disease. Am J Trop Med Hyg 60: 377–386.
  28. 28. Porter CH, Collins FH (1996) Phylogeny of nearctic members of the Anopheles maculipennis species group derived from the D2 variable region of 28S ribosomal RNA. Mol Phylogenet Evol 6: 178–188.
  29. 29. Herrera-Aguilar M, Be-Barragán LA, Ramirez-Sierra MJ, Tripet F, Dorn P, et al. (2009) Identification of a large hybrid zone between sympatric sibling species of Triatoma dimidiata in the Yucatan peninsula, Mexico, and its epidemiological importance. Infect Genet Evol 9: 1345–1351.
  30. 30. Thompson JD, Gibson TJ, Plewniak F, Jeanmougin F, Higgins DG (1997) The ClustalX windows interface: Xexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Res 24: 4876–4882.
  31. 31. Hall TA (1999) BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symp Ser 41: 95–98.
  32. 32. Márquez E, Jaramillo-O N, Gómez-Palacio A, Dujardin JP (2011) Morphometric and molecular differentiation of a Rhodnius robustus-like form from R. robustus Larousse, 1927 and R. prolixus Stal, 1859 (Hemiptera, Reduviidae). Acta Trop 120: 103–109.
  33. 33. Librado P, Rozas J (2009) DnaSP v5: a software for comprehensive analysis of DNA polymorphism data. Bioinformatics 25: 1451–1452.
  34. 34. Kimura M (1980) A simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide sequences. J Mol Evol 16: 111–120.
  35. 35. Monteiro FA, Barrett TV, Fitzpatrick S, Cordon-Rosales C, Feliciangeli D, et al. (2003) Molecular phylogeography of the Amazonian Chagas disease vectors Rhodnius prolixus and R. robustus. Mol Ecol 12: 997–1006.
  36. 36. Tamura K, Peterson D, Peterson N, Stecher G, Nei M, et al. (2011) MEGA5: molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Mol Biol Evol 28: 2731–2739.
  37. 37. Akaike H (1974) A new look at the statistical model identification. Automatic Control, IEEE Transactions on Automatic Control 19: 716–723.
  38. 38. Posada D (2008) jModelTest: phylogenetic model averaging. Mol Biol Evol 25: 1253–1256.
  39. 39. Drummond AJ, Suchard MA, Xie D, Rambaut A (2012) Bayesian phylogenetics with BEAUti and the BEAST 1.7. Mol Biol Evol 29: 1969–1973.
  40. 40. Goloboff PA, Farris JS, Nixon KC (2008) TNT, a free program for phylogenetic analysis. Cladistics 24: 774–786.
  41. 41. Efron B, Halloran E, Holmes S (1996) Bootstrap confidence levels for phylogenetic trees. Proc Natl Acad Sci U S A 93: 7085–7090.
  42. 42. Rambaut A (2010) FigTree 1.3.1. http://tree.bio.ed.ac.uk/software/figtree/
  43. 43. Drummond AJ, Ho SY, Phillips MJ, Rambaut A (2006) Relaxed phylogenetics and dating with confidence. PLoS Biol 4: e88.
  44. 44. Hoorn C, Wesselingh F, Hovikoski J, Guerrero J (2010) The Development of the Amazonian Mega-Wetland (Miocene; Brazil, Colombia, Peru, Bolivia); Wiley-Blackwell, editor: Oxford.
  45. 45. Bandelt HJ, Forster P, Röhl A (1999) Median-joining networks for inferring intraspecific phylogenies. Mol Biol Evol 16: 37–48.
  46. 46. Panzera F, Pérez R, Panzera Y, Ferrandis I, Ferreiro MJ, et al. (2010) Cytogenetics and genome evolution in the subfamily Triatominae (Hemiptera, Reduviidae). Cytogenet Genome Res 128: 77–87.
  47. 47. Panzera F, Ferrandis I, Ramsey J, Salazar-Schettino PM, Cabrera M, et al. (2007) Genome size determination in chagas disease transmitting bugs (Hemiptera-Triatominae) by flow cytometry. Am J Trop Med Hyg 76: 516–521.
  48. 48. Panzera F, Perez R, Panzera Y, Alvarez F, Scvortzoff E, et al. (1995) Karyotype evolution in holocentric chromosomes of three related species of triatomines (Hemiptera-Reduviidae). Chromosome Res 3: 143–150.
  49. 49. Rohlf FJ (2010) tpsDig, digitize landmarks and outlines, version 2.16. Department of Ecology and Evolution, State University of New York at Stony Brook, NY.
  50. 50. Bookstein FL (1991) Morphometric Tools for Landmark Data: Geometry and Biology. New York: Cambridge University Press. 435 p.
  51. 51. Rohlf FJ (1997) tpsRelw, version 1.11.: Department of Ecology and Evolution, State University of New York at Stony Brook, NY.
  52. 52. Hammer Ø, Harper D, Ryan P (2001) PAST: Paleontological Statistics Software Package for Education and Data Analysis. Palaeontologia Electronica 4: 9.
  53. 53. Gregory-Wodzicki KM (2000) Uplift history of the Central and Northern Andes: A review. Geol Soc Am Bull 112: 1091–1105.
  54. 54. Coates AG, Collins LS, Aubry M-P, Berggren WA (2004) The Geology of the Darien, Panama, and the late Miocene-Pliocene collision of the Panama arc with northwestern South America. Geol Soc Am Bull 116: 1327–1344.
  55. 55. Panzera Y, Pita S, Ferreiro MJ, Ferrandis I, Lages C, et al. (2012) High dynamics of rDNA cluster location in kissing bug holocentric chromosomes (Triatominae, Heteroptera). Cytogenet Genome Res 138: 56–67.
  56. 56. Arnold M (1997) Natural Hybridization and Evolution. New York: University Press, New York, Oxford.
  57. 57. Costa J, Peterson AT, Dujardin JP (2009) Morphological evidence suggests homoploid hybridization as a possible mode of speciation in the Triatominae (Hemiptera, Heteroptera, Reduviidae). Infect Genet Evol 9: 263–270.
  58. 58. Pérez R, Hernández M, Quintero O, Scvortzoff E, Canale D, et al. (2005) Cytogenetic analysis of experimental hybrids in species of Triatominae (Hemiptera-Reduviidae). Genetica 125: 261–270.
  59. 59. Martínez-Ibarra JA, Grant-Guillén Y, Ventura-Rodríguez LV, Osorio-Pelayo PD, Macías-Amezcua MD, et al. (2011) Biological and genetic aspects of crosses between species of the genus Meccus (Hemiptera: Reduviidae Triatominae). Mem Inst Oswaldo Cruz 106: 293–300.
  60. 60. Pigliucci M (2005) Evolution of phenotypic plasticity: where are we going now? Trends Ecol Evol 20: 481–486.
  61. 61. Dujardin JP, Costa J, Bustamante D, Jaramillo N, Catalá S (2009) Deciphering morphology in Triatominae: the evolutionary signals. Acta Trop 110: 101–111.