17
views
0
recommends
+1 Recommend
2 collections
    0
    shares

      Publish your biodiversity research with us!

      Submit your article here.

      • Record: found
      • Abstract: found
      • Article: found
      Is Open Access

      Seven new species of Notiospathius (Hymenoptera, Braconidae, Doryctinae) from Northwest Venezuela

      Read this article at

      ScienceOpenPublisher
      Bookmark
          There is no author summary for this article yet. Authors can add summaries to their articles on ScienceOpen to make them more accessible to a non-specialist audience.

          Abstract

          We describe seven new species of the doryctine wasp genus Notiospathius Matthews & Marsh from Northwest Venezuela: N. araguae sp. n., N. bolivari sp. n., N. curvilineatus sp. n., N. dantei sp. n., N. estradae sp. n., N. larensis sp. n., and N. venezuelae sp. n. These represent the first described species of the genus reported for this country. Species boundaries for the above taxa were confirmed using the General mixed Yule-coalescent method with 441 DNA barcode sequences from specimens assigned to Notiospathius and other closely related genera collected in different countries along the Neotropics.

          Related collections

          Most cited references9

          • Record: found
          • Abstract: found
          • Article: not found

          Integrative taxonomy: a multisource approach to exploring biodiversity.

          Good alpha taxonomy is central to biology. On the basis of a survey of arthropod studies that used multiple disciplines for species delimitation, we evaluated the performance of single disciplines. All included disciplines had a considerable failure rate. Rigor in species delimitation can thus be increased when several disciplines chosen for complementarity are used. We present a flexible procedure and stopping rule for integrative taxonomy that uses the information from different disciplines separately. Disagreement among disciplines over the number and demarcation of species is resolved by elucidating and invoking evolutionary explanations for disagreement. With the identification of further promising study organisms and of new questions for in-depth analysis, evolutionary biology should profit from integrative taxonomy. An important rationale is clarity in researcher bias in the decision-making process. The success of integrative taxonomy will further increase through methodological progress, taxonomic training of evolutionary biologists, and balanced resource allocation.
            Bookmark
            • Record: found
            • Abstract: not found
            • Article: not found

            A new method for calculating evolutionary substitution rates

              Bookmark
              • Record: found
              • Abstract: found
              • Article: not found

              Independently Evolving Species in Asexual Bdelloid Rotifers

              Introduction Species are fundamental units of biology, but there remains uncertainty on both the pattern and processes of species existence. Are species real evolutionary entities or convenient figments of taxonomists' imagination [1–3]? If they exist, what are the main processes causing organisms to diversify [1,4]? Despite considerable debate, surprisingly few studies have formally tested the evolutionary status of species [1,5,6]. One central question concerning the nature of species has been whether asexual organisms diversify into species [1]. The traditional view is that species in sexual clades arise mainly because interbreeding maintains cohesion within species, whereas reproductive isolation causes divergence between species [7]. If so, asexuals might not diversify into distinct species, because there is no interbreeding to maintain cohesive units above the level of the individual. However, if other processes were more important for maintaining cohesion and causing divergence, for example, specialization into distinct niches, then asexuals should diversify in a manner similar to sexuals, although the rate and magnitude of divergence might differ [8–11]. Empirical evidence to test these ideas has been rare. Most asexual animal and plant lineages are of recent origin [9,12]. The diffuse patterns of variation typical of such taxa [13] could simply reflect their failure to survive long enough for speciation to occur or the effects of ongoing gene flow from their sexual ancestors [9,12]. Distinct genetic and phenotypic clusters have been demonstrated in bacteria [14–17] and discussed as possible evidence for clonal speciation [1]. However, all the study clades engage in rare or even frequent recombination as well as clonal reproduction [14,18,19]. Although horizontal gene transfer can occur between distantly related bacteria, homologous recombination occurs only at appreciable frequency between closely related strains [20,21]. Therefore, clusters in these bacteria could arise from similar processes to interbreeding and reproductive isolation in sexual eukaryotes [20]. Aside from issues of sexuality, previous studies looking for distinct clusters have been descriptive, relying on visual interpretation of plots of genetic or phenotypic variation rather than on formal tests of predictions under null and alternative evolutionary scenarios [1]. Here, we demonstrate that a classic asexual clade, the bdelloid rotifers, has diversified into independently evolving and distinct entities arguably equivalent to species. Bdelloids are abundant animals in aquatic or occasionally wet terrestrial habitats and represent one of the best-supported clades of ancient asexuals [22–24]. They reproduce solely via parthenogenetic eggs, and no males or traces of meiosis have ever been observed. Molecular evidence that bdelloid genomes contain only divergent copies of nuclear genes present as two similar copies (alleles) in diploid sexual organisms rules out anything but extremely rare recombination [25–27]. Yet, bdelloids have survived for more than 100 million y and comprise more than 380 morphologically recognizable species and 20 genera [28]. The diversity of the strictly asexual bdelloids poses a challenge to the idea that sex is essential for long-term survival and diversification [29]. However, taxonomy does not constitute strong evidence for evolutionary species: the species could simply be arbitrary labels summarizing morphological variation among a swarm of clones [7]. We adopt a general evolutionary species concept, namely, that species are independently evolving and distinct entities, and then break the species problem into a series of testable hypotheses derived from population genetic predictions [3]. We use the word “entity” to refer to a set of individuals comprising a unit of diversity according to a given criterion or test: the question of whether to call those entities “species” will be returned to below. Focusing on the genus Rotaria (Figure 1), one of the best-characterized genera of bdelloids, we use combined molecular and morphological analyses to distinguish alternative scenarios for bdelloid diversification (Figure 2). First, the entire clade might represent a single species, that is, a swarm of clones with no diversification into independently evolving subsets of individuals. Second, the clade may have diversified into a series of independently evolving entities. By “independently evolving,” we mean that the evolutionary processes of selection and drift operate separately in different entities [8,9], such that genotypes can only spread within a single entity. Possible causes of independence include geographical isolation or adaptation to different ecological niches [10,17]. The expected outcome is cohesion within entities but genetic and phenotypic divergence between them [9–11]. Figure 1 SEM Pictures of Some Species of the Genus Rotaria (A) R. neptunia, lateral view; (B) R. macrura, ventral view; (C) R. tardigrada, dorsal view; (D) R. sordida, lateral view; and (E) trophi of R. tardigrada with open circles showing the location of landmarks used for the shape analysis. Scale bars: 100 μm for animals, 10 μm for trophi. Figure 2 Scheme Showing the Predicted Patterns of Genetic and Morphological Variation Underlying Our Tests of Alternative Scenarios of Diversification (A) Hypothetical trees showing expected genetic relationships among a sample of individuals under the null model that the sample is drawn from a single asexual population (H0) and under the alternative model that the clade has diversified into a set of independently evolving entities (H1). (B) Expected variation in two ecomorphological traits evolving either neutrally (H0) or by adaptive divergence (H1) in a genus that has diversified into six genetic clusters. Note that a mixed pattern is possible: Some genetic clusters may have experienced divergent selection on morphology, whereas others have not. We first test for the presence of independently evolving entities. Under the null scenario of no diversification, genetic relationships should conform to those expected for a sample of individuals from a single asexual population (H0, Figure 2A). Under the alternative scenario that independently evolving entities are present, we expect to observe distinct clusters of closely related individuals separated by long branches from other such clusters (H1, Figure 2A; and [9]). Coalescent models can be used to distinguish the two scenarios [30]. Failure to reject the null model would indicate a lack of evidence for the existence of independently evolving entities. Next, to investigate the role of adaptation to different niches in generating and maintaining diversity within the clade, we extend classic methods from population genetics to test directly for adaptive divergence of ecomorphological traits. If trait diversity evolves solely by neutral divergence in geographic isolation, we expect morphological variation within and between entities to be proportional to levels of neutral genetic variation (H0, Figure 2B, Materials and Methods). If, instead, different entities experience divergent selection on their morphology, we expect greater morphological variation between clusters than within them, relative to neutral expectations (H1, Figure 2B; and [31]). Past work has often discussed sympatry of clusters as evidence for niche divergence [1], but, in theory, coexistence can occur without niche differences [32]; hence, we introduce an alternative, more direct approach. Our results demonstrate that bdelloids have diversified not only into distinct genetic clusters, indicative of independent evolution, but also into entities experiencing divergent selection on feeding morphology, indicative of niche divergence. In common with findings of cryptic species in sexual organisms [33,34], the morphologically coherent groups experiencing divergent selection often include several genetic clusters: this introduces difficulties in deciding which units to call species, but this problem is shared with sexual organisms [3,33]. In short, bdelloids have diversified into entities equivalent to sexual species in all respects except that individuals do not interbreed. The results demonstrate the benefits of statistical analyses of combined molecular and morphological data for exploring the evolutionary nature of species. Results/Discussion We collected all individuals of Rotaria encountered during 3 y searching rivers, standing water, dry mosses, and lichens, centered on Italy and the United Kingdom but also globally [35]. Individuals were identified to belong to nine taxonomic species (Tables S1 and S2). Most of the described species of Rotaria missing from our sample are known from only one record or are very rarely encountered (Protocol S1). Bayesian and maximum parsimony analyses of mitochondrial cytochrome oxidase I (cox1) and nuclear 28S ribosomal DNA sequences provide strong support for the monophyly of taxonomic species (Figures 3, S1, S2, and S3 and Text S1), with the sole exception of R. rotatoria, which was already suspected to comprise a species complex based on disagreements among authors [36,37]. Figure 3 Phylogenetic Relationships in the Genus Rotaria The consensus of 80,000 sampled trees from Bayesian analysis of the combined cox1 and 28S rDNA data sets is shown, displaying all compatible groupings and with average branch lengths proportional to numbers of substitutions per site under a separate GTR + invgamma substitution model for the cox1 and 28S partitions. Posterior probabilities above 0.5 and bootstrap support above 50% from a maximum parsimony bootstrap analysis are shown above and below each branch, respectively. Support values for within-species relationships are not shown for very short branches but are shown in Figures S1 through S3. Closed circles indicate clusters identified by the clustering analysis. Colors represent traditional species memberships. Diamonds indicate taxonomic species and monophyletic groups of Rotaria. Names refer to the species, the country, the number of site within that country for that species, and the number of individual from that site if several were isolated; for example, R.macr.IT.1.1 refers to the first individual from site 1 in Italy for R. macrura. Pictures of trophi from one individual from each cluster are shown to scale: Representatives of all sampled populations are shown in Figure S4. A full list of names and localities of samples is available in Table S1. Morphometric analyses further support the distinctness of taxonomic species. Bdelloid morphology is hard to measure because of their shape-changing abilities; hence, we used geometric morphometrics [38] to measure the only suitable trait, their hard jaws, called trophi [39] (Figures 1 and S4). Trophi size and shape are not characters that have been used in the traditional taxonomy of the genus (Table S2). Trophi scale weakly with rough measures of body size of each species (mean trophi size against log body length from [37]: r = 0.55, p = 0.2, Spearman's rank test), and both the size and shape of trophi likely reflect different types or sizes of particulate food consumed, although the details of how food is processed remain unclear [28]. Discriminant analysis of the first five principal components (PCs) describing trophi shape (cumulative explained variance, 97.1%; Materials and Methods) produced a correct classification with respect to traditional taxonomy of most specimens of R. macrura, R. neptunia, R. sordida, and R. tardigrada (Table S3). The remaining species overlapped in shape but could be discriminated by size (Figures 4 and S5). Related species on the DNA trees tend to have similar morphology: for example, R. magnacalcarata, R. socialis, and R. rotatoria FR.2.1 and IT.5 overlap in shape, but are more distant from R. rotatoria UK.2.2. Only two of the traditional species found to be monophyletic in the DNA tree displayed significant variation in size or shape among populations: R. sordida and R. tardigrada. In both cases, the populations that differed were deeply divergent in the DNA tree as well. Figure 4 Plot of the Size and Shape (the First Two PCs, PC1 and PC2, from the Generalized Procrustes Analysis) of Trophi across Species The directions of shape variation along each axis are shown for PC1 and PC2, respectively, using ×2 and ×4 magnification of the observed variation for emphasis. PC1 represents a continuum from oval to rounder trophi and from parallel to converging major teeth. PC2 represents a trend in the distance of the major teeth from the attachment point between the two halves of the trophi. Congruence between molecules and morphology confirms that most traditional Rotaria species are monophyletic clades but does not rule out the possibility that taxa reflect variation within a single asexual species or swarm of evolutionarily interacting clones. Under the alternative scenario that independently evolving entities are present, we expect to observe clusters of closely related individuals separated from other such clusters by longer internal branches on a DNA tree [9,30,40]. We therefore tested for significant clustering by comparing two models describing the likelihood of the branching pattern of the DNA trees: first, a null model that the entire sample derives from a single population following a neutral coalescent [41], and, second, a model assuming a set of independently evolving populations joined by branching that reflects the timing of divergence events between them, that is, cladogenesis [9,30,42]. The models allow departures from strict assumptions of constant population size and rates of cladogenesis (see Materials and Methods). The results indicate significant clustering within Rotaria, as expected if several independently evolving entities are present and consistent with patterns of mtDNA diversity from a broad sample of bdelloids [43]. The maximum likelihood solution for the independent evolution model on the combined tree infers 13 isolated clusters, with the remaining individuals inferred to be singletons (Figure 3; Table S4). Two monophyletic taxonomic species contained two separate clusters: R. magnacalcarata has two clusters corresponding to the U.K. and Italian samples, whereas R. macrura has two clusters not matching sampling locality. Uncorrected pairwise distances of cox1 within clusters ranged from 0% to 3.3% (mean, 1.5%), and those between clusters ranged from 4.1% to 23.1% (mean, 16.0%). The null model that the entire lineage represents a single cluster can be rejected (log likelihood ratio test, 2 × ratio = 30.8, χ2 test, three degrees of freedom, p 1) population size [30]. Under the alternative model that the sample derives from a set of independently evolving populations, each one evolving similarly to the null case, we calculated the likelihood of waiting times as Equation 6 from Pons et al. [30]. The alternative model optimizes a threshold age, T, such that nodes before the threshold are considered to be diversification events with branching rate λ D and scaling parameter p D. Branches crossing the threshold define k clusters each obeying a separate coalescent process but with branching rate, λ C, and scaling parameter, p C, assumed to be constant across clusters. The alternative model thus has three additional parameters. Models were fitted using an R script available from T.G.B. to an ultrametric tree obtained by rate smoothing the combined analysis DNA tree using penalized likelihood in r8s (http://ginger.ucdavis.edu/r8s) and cross-validation to choose the optimal smoothing parameter for each tree [60]. Test for divergent selection. In an asexual clade, all genes have the same underlying genealogy: the entire genome is inherited as a single unit. Assuming that silent substitutions are neutral, the expected number of silent mtDNA substitutions on a branch of the genealogy is μt, where t is the branch length in units of time and μ is the mutation rate of the gene. Assuming a neutral morphological trait evolving by Brownian motion, the expected squared change (variance) along a branch is , where is the mutational rate of increase of variance [61]. The expectations are the same for branches within populations or between them. Therefore, the average rate of change of a neutral trait expressed as variance per silent substitution should be the same within populations as between them, that is, This prediction holds even if mutation rates vary across the tree, providing they do so without a systematic bias between the branch classes being compared, a reasonable assumption shared with widely used molecular versions of the test [31]. We reconstructed evolutionary changes in trophi size and shape (PC1 and PC2) onto the DNA tree using the Brownian motion model by Schluter et al. [62] implemented in the Ape library for R [63]. Branch lengths were optimized as the proportion of silent substitutions per codon using PAML software [64]. The null model assumes a constant rate of morphological change across the entire tree. The alternative model labels branches as between taxonomic species, within species and within clusters, and estimates different rates for each class. Under a three rate-class model, the likelihood of the reconstruction, Equation 3 of [62] becomes the product of the equivalent likelihood for each class of branches. where k indicates the branch classes from 1 to 3, βk is the rate parameter for each class of branches, Nk is the number of nodes ancestral to each class of branch, and Q(ũk ) is the sum of the scaled variance of changes across branches [62] of class k. Optimization was implemented in a modified version of the “ace” function of Ape, available from T. G. B. Divergent selection between taxonomic species, for example, would be indicated by a significantly lower rate within cluster and within species branches (classes 1 and 2) than between species branches (class 3). Assumptions and robustness of the test are discussed further in Protocol S2. Supporting Information Figure S1 Phylogenetic Relationships from Bayesian Analysis of the Combined Data Posterior probabilities from the Bayesian analysis are indicated next to each node. Below the branch are bootstrap percentages from a maximum parsimony search with 100 bootstrap replicates each using a heuristic search with 100 random addition replicates, TBR branch swapping, and saving only one tree per addition replicate. (21 KB PDF) Click here for additional data file. Figure S2 Phylogenetic Relationships from Bayesian Analysis of cox1 Posterior probabilities are indicated above each branch; parsimony bootstrap values are indicated below each branch. (20 KB PDF) Click here for additional data file. Figure S3 Phylogenetic Relationships from Bayesian Analysis of 28S rDNA Posterior probabilities are indicated above each branch; parsimony bootstrap values are indicated below each branch. (14 KB PDF) Click here for additional data file. Figure S4 SEM Pictures of Trophi from Each Study Population (A) R. macrura macrIT2; (B) R. macrura macrIT1; (C) R. macrura macrIT3; (D) R. macrura R.macr.UK.1; (E) R. magnacalcarata magnIT1; (F) R. magnacalcarata magnIT3; (G) R. magnacalcarata magnIT2; (H) R. magnacalcarata R.magn.UK.2.1; (I) R. socialis sociIT1; (J) R. socialis sociIT2; (K) R. socialis sociIT3; (L) R. socialis R.soci.UK, (M) R. rotatoria R.rota.IT.5; (N) R. rotatoria R.rota.FR.2.1; (O) R. rotatoria R.rota.UK.2.2; (P) R. sordida sordIT1; (Q) R. sordida sordIT2; (R) R. sordida sordAU; (S) R. neptunoida noidIT; (T) R. neptunia R.nept.IT, (U) R. tardigrada tardIT1; (V) R. tardigrada tardIT3; (W) R. tardigrada R.tard.US; and (X) landmarks and links used for shape analysis. (197 KB PDF) Click here for additional data file. Figure S5 Box Plot of the Size of Trophi for Each Study Population Analysis of variance test, ln CS: F 22,303 = 684.17, p < 0.0001. (56 KB PDF) Click here for additional data file. Protocol S1 Sampling, Molecular Analyses, and Morphometrics (87 KB PDF) Click here for additional data file. Protocol S2 Test for Divergent Selection on Morphology (72 KB PDF) Click here for additional data file. Table S1 Locality Records for DNA Sequences and Morphometric Measurements (78 KB PDF) Click here for additional data file. Table S2 Traditional Taxonomy of Rotaria Species (60 KB PDF) Click here for additional data file. Table S3 Discriminant Analysis of Trophi Shape (29 KB PDF) Click here for additional data file. Table S4 Comparison of Models of Single versus Multiple Independently Evolving Entities (47 KB PDF) Click here for additional data file. Table S5 Comparison of Alternative Models for Rates of Changes in Trophi Size and Shape within and between Clusters and Species (37 KB PDF) Click here for additional data file. Table S6 The Effects of Sampling Type (Clonal versus Population Sample) on the Mean and Variance of Size and Shape of Trophi (39 KB PDF) Click here for additional data file. Text S1 Comparison of cox1 and 28S rDNA Results (57 KB PDF) Click here for additional data file. Accession Numbers DNA sequences have been deposited at GenBank (http://www.ncbi.nlm.nih.gov/Genbank) under accession numbers DQ656756 to DQ656882.
                Bookmark

                Author and article information

                Journal
                Journal of Hymenoptera Research
                JHR
                Pensoft Publishers
                1314-2607
                1070-9428
                October 15 2012
                October 15 2012
                : 29
                : 37-61
                Article
                10.3897/jhr.29.3555
                39e18a3e-9929-4ae1-81d5-b7824ed73e96
                © 2012

                http://creativecommons.org/licenses/by/3.0/

                History

                Comments

                Comment on this article