Next Article in Journal
Effect of Aquatic Vegetation Restoration after Removal of Culture Purse Seine on Phytoplankton Community Structure in Caizi Lakes
Next Article in Special Issue
Molecular and Phytochemical Variability of Endemic Juniperus sabina var. balkanensis from Its Natural Range
Previous Article in Journal
Early Development of the Endemic Delminichthys krbavensis (Leuciscidae, Cypriniformes) from a Karstic Field in Croatia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Molecular Insights into the Centaurea Calocephala Complex (Compositae) from the Balkans—Does Phylogeny Match Systematics?

1
Institute of Botany and Botanical Garden “Jevremovac”, Faculty of Biology, University of Belgrade, Studentski trg 16, 11000 Belgrade, Serbia
2
Institut Botànic de Barcelona (IBB, CSIC-Ajuntament de Barcelona), Passeig del Migdia s. n., Parc de Montjuïc, 08038 Barcelona, Spain
3
Ministry of Economy and Sustainable Development, Radnička cesta 80/7, 10000 Zagreb, Croatia
4
Nature Science Department, National Museum of Banat, Huniade Square 1, 300002 Timisoara, Romania
5
Department of Biology and Ecology, Faculty of Sciences and Mathematics, University of Niš, Višegradska 33, 18000 Niš, Serbia
*
Author to whom correspondence should be addressed.
Submission received: 8 April 2022 / Revised: 12 May 2022 / Accepted: 13 May 2022 / Published: 16 May 2022
(This article belongs to the Special Issue Conservation Genetics and Biogeography of Seed Plant Species II)

Abstract

:
Groups of recent speciation are characterized by high levels of introgression and gene flow, which often confounds delimitation of species on a DNA basis. We analyzed nuclear DNA sequences (ETS spacer and the AGT1 gene) obtained from a large sample of the C. calocephala complex from the Balkan clade of Centaurea sect. Acrocentron (Compositae, Cardueae-Centaureinae) together with a wide representation of other species from the section. Our main goals were to verify the monophyly of the complex as currently defined and to examine the possible presence of introgression and gene flow. Within the complex, species are well-delimited from a morphological point of view and probably originated by allopatric speciation in the Balkan Peninsula. Our results confirm that the Balkan–Eurasian complex is a natural group, but the Centaurea calocephala complex shows a very complicated pattern and its phylogeny is not resolved. Our hypothesis suggests that altitudinal shifts in the transits from glacial to interglacial periods caused successive hybridization events, which are very evident from the DNA networks, between taxa not currently sympatric. As a result, confirmation of interspecific boundaries using molecular markers is extremely complicated.

Graphical Abstract

1. Introduction

The Balkan Peninsula is located in the Southeast of Europe and harbors a very rich flora with more than 7500 species of native plants [1,2], approximately one-third of them endemic [3]. The biodiversity of this region is the result of a complex geological history and interactions between populations, species, and ecosystems [4]. The Balkans were one of the major glacial refugia for plants of the temperate zone, whence a post-glacial expansion to the north took place [5,6,7,8,9,10]. Many molecular and phylogeographic studies have been conducted to clarify genetic relationships between taxa from this area, e.g., Campanula [11,12,13,14,15,16,17], Tanacetum [18], Arundo [19], Cardamine [20,21,22], Cerastium [23,24], Edraianthus [25,26,27,28], Cyanus [29], Salvia [29,30], Goniolimon [31], Sesleria [32,33], and Veronica [34].
After Quaternary glaciations, rapid radiations occurred within genera with generally herbaceous representatives [35], and a good example is the genus Centaurea L. Centaurea belongs to subtribe Centaureinae (Cass.) Dumort., tribe Cardueae Cass. from the family Compositae Giseke, and comprises more than 250 species, divided into 40 sections [36]. Centaurea sect. Acrocentron (Cass.) DC., one of the largest sections of the genus, is widely distributed, mainly in the Mediterranean region, and includes around 100 species [37,38,39]. This section is well-defined in terms of the following characters: very large capitula; peripheral florets sterile, without staminodes but provided with achenioids; spiny, long decurrent appendages; and presence of the “Centaurea scabiosa” pollen type [40,41]. The basic chromosome numbers in sect. Acrocentron are x = 10 (more frequent) and x = 11 [42,43,44,45]. According to the phylogeny by [46], species of sect. Acrocentron constitute a natural group and they are classified in several well-supported, monophyletic clades: Balkan–Eurasian, Iberian, North African, Aegean, and Anatolian–Iranian clade.
Within sect. Acrocentron, there are some groups of taxa with complex taxonomy. One of them is the Centaurea calocephala complex from the Balkans. Species from the Centaurea calocephala complex are characterized by large, showy capitula, and coriaceous, ovate to oblong involucral bracts with large, shortly decurrent appendages usually covering the bracts; mucronulate to shortly spinose at apex; yellow, brown, or black in the central part; and fimbriate at margins [47,48]. The typical species of this complex are Centaurea calocephala Willd. (syn. Centaurea atropurpurea Waldst. and Kit.).; C. chrysolepis Vis.; C. crnogorica Rohlena; C. gjurasinii Bošnjak; C. grbavacensis (Rohlena) Stoj. and Acht.; C. immanuelis-loewii Degen; C. kotschyana Heuffel ex Koch; C. melanocephala Pančić; C. murbeckii Hayek; C. orientalis L.; and C. zlatiborensis Zlatković, Novaković and Janaćković. All species are morphologically well defined (Figure 1 and Figure 2), and although their geographical distribution is partly overlapping in the big picture, populations generally have an allopatric distribution. It is extremely rare for different species of this complex to coexist (syntopia).
Many species of this complex are distributed in different parts of the Balkan Peninsula, and some of them have a wider distribution than the rest: C. calocephala is a Balkan-Carpathian subendemic, C. kotschyana is a Balkan–Carpathian and Central Europe mountain plant, and C. orientalis has a Pontic–Submediterranean distribution (Table S1). The ecological optimum of all species is found in different types of grass communities, mostly on carbonated soils on limestone, dolomite, marble, or loess, less often on ultramafic or siliceous rocks. They usually grow in dry rocky grasslands, rarely on screes, vertical cliffs, loess steppic grasslands, or mesophilous mountain and subalpine meadows and pastures (Table S1).
One of the main drivers of speciation in Centaurea is hybridization [42,49,50,51,52]. Hybridization is often associated with the emergence of new species, with or without changes in chromosome number [53,54]. Homoploid hybrid species arise when there is no change in ploidy. Homoploid speciation presents additional challenges in relation to allopolyploid speciation, because chromosome doubling prevents backcrossing with parent species, and the chromosome set of the new species contains complete parental genomes [55]. In the absence of polyploidy, incompatibility between parental genomes may cause unsustainability of the homoploid hybrid, except when hybridization occurs between closely related taxa [56,57]. In addition, the lack of reproductive, ecological, and spatial barriers between new homoploid hybrids and their ancestors often leads to introgression and genetic intertwining with their parents, preventing hybrid persistence and their subsequent stabilization and speciation [55,58]. Therefore, known homoploid hybrids are rarer than polyploids. However, recent molecular analyses have revealed a number of homoploid hybrids, and surely the frequency of homoploid hybridization has been miscalculated because it is not as obvious as allopolyploidy [59,60]. Homoploid hybridization is common between related Centaurea species [60,61]. The Balkan Centaurea calocephala complex is probably not an exception, as intermediate individuals between some species have been detected [47] and hybridization is usually homoploid because most of species are diploid (Table S2). Dispersal of pollen and achenes in Centaurea is restricted and hybridization is possible only in the same geographical area where two different lineages meet [62,63]. Therefore, historical changes in the area of distribution of the species are usually invoked to explain gene flow between taxa not sympatric at present [51].
Molecular analyses can detect hybridization as well as identify parental species [64], and they may reveal events that are not evident from morphology at present, for example, old hybridization events between species [51]. Low-copy genes (e.g., AGT1) are excellent for detecting the parental species of hybrids because they have high rates of evolution and are present in the genome with one to few copies [64,65]. Nuclear ribosomal ETS can also be used in cases of introgression and reticulation because concerted evolution is generally incomplete in Centaurea sect. Acrocentron and different copies often persist [46,51,64]. Moreover, there are other examples of several copies in nuclear-ribosomal DNA in diploids: Picris [66], Rheum (ITS, cf. [67]), or Leucanthemum (ETS, cf. [68]). We gathered a large sample of the C. calocephala complex and sequenced ETS region and the AGT1 gene with the aims to (a) test the monophyly of the Balkan clade of sect. Acrocentron on a much wider sampling; (b) check the monophyly of the C. calocephala complex as currently defined; and (c) examine the possible presence of introgression and gene flow in the complex.

2. Materials and Methods

2.1. Plant Material

We included all species of Centaurea sect. Acrocentron from the Balkans, following the classification by [47]. Sampling focused on C. calocephala, with 10 populations (Figure 3), together with 12 populations of 9 species from the C. calocephala complex. From each population, we included one individual, with the exception of three mixed populations from Đerdap, Lika, and Suva Gora, for which we used from two to three individuals. To verify whether the Balkan group of species is monophyletic, we included a very wide representation of species of Acrocentron from the rest of the area of the section: Iran, Armenia, Turkey, Greece and the Aegean Islands, the Italian Peninsula and Sicily, North Africa, and the Iberian Peninsula (46 species). The ETS sequences of most of the non-Balkan species were taken from [46]. In total, the study comprised 72 populations of 61 species. The localities, populations, and GenBank accession numbers are detailed in Table S2.

2.2. DNA Extraction, Amplification, Cloning, and Sequencing

Genomic DNA was extracted from silica-dried leaves collected in the field using the CTAB protocol described by [69] as modified by [70,71].
The ETS region was amplified with the forward primer ETS1F [72] and the reverse primer 18S-ETS [73]. The profile and reactions used for PCR amplification were the same as described in [74]. The AGT1 gene was amplified using the specific forward primer Agt1F-podos previously designed by [75] and the universal reverse primer Agt1R [76]. The PCR reaction and amplification profiles were as described in [75].
All PCR products of the ETS region and the AGT1 gene from most of Balkan populations were cloned (Table S2) using a TOPO TA Cloning® Kit (Invitrogen, Carlsbad, CA, USA). Whenever possible, 8 to 16 positive colonies from each PCR reaction were screened with direct PCR using universal primers T7 and M13 following the protocol of [77]. Eight to ten PCR products were selected for sequencing using the same universal primers previously mentioned.
The PCR products were purified using a QIAquick PCR Purification Kit (Qiagen Inc., Valencia, CA, USA). Sequencing was performed on an ABI 3730xl (Applied Biosystems) following the manufacturer’s protocol at Macrogen Inc., Korea.

2.3. Phylogenetic Analyses

Sequences were aligned visually using BioEdit 7.0.5.3 [77]. Single substitutions in clones from a single accession were excluded. Consensus sequences were generated for some accessions and regions, condensing single base-pair differences between clones. This reduced the size of the matrices as well as the impact of PCR artifacts (chimeric sequences and Taq errors; cf. [78,79]). For verifying the presence of possible recombinant sequences, datasets were checked using RDP4.97β [80].
We used three datasets. Dataset 1 comprised 88 sequences of the ETS region from a representative sampling of sect. Acrocentron from all its distribution area following [46]. Dataset 2 included 66 cloned sequences of the ETS region of populations from the Calocephala complex only. Dataset 3 included 41 sequences of populations of the Calocephala complex of cloned AGT1 gene. On dataset 1, we carried out a Bayesian inference analysis using the evolutionary model determined by jModeltest v.2.1.10 [81]. The model GTR + G was selected as the best-fit model of nucleotide substitution using Akaike information criteria (AIC). Bayesian inference analysis was carried out using Mr. Bayes v. 3.2 [82] using default prior settings. Analysis was initiated with random starting trees, and four Markov chains were run simultaneously for 30 × 106 generations. We saved one out every 1000 generations, and the first 7500 generations were discarded as the “burn-in” period after confirming that log-likelihood values had stabilized. Posterior probabilities ≥0.95 were considered statistically significant.
On datasets 2 and 3, we carried out distance network analyses (split graphs) in order to represent groupings in the data and evolutionary distance between pairs of taxa simultaneously. We used the Neighbor-Net (NN) algorithm [83] as implemented in SplitsTree4 v4.13.1 software [84] with the criterion set to uncorrected pairwise (p) distances and including gaps.

2.4. Chromosome Number

Chromosome numbers were determined in 18 populations of 10 taxa (Table S2). For the examination of mitotic chromosomes, we excised root tips from germinating seeds and pre-treated them with 0.002 M 8-hydroxyquinoline for 4 h at 8 °C. Pretreated tips were fixed in cold 3/1 (v/v) absolute ethanol/glacial acetic acid for 48 h and stored in 70% ethanol at 4 °C for further use. They were hydrolyzed in 1N HCl for 11 min at 60 °C and stained in Schiff’s reagent [85] for at least 2 h and squashed in a drop of acetic carmine. Metaphasic plates were examined and photographed using a Leica DMLS light microscope equipped with a Leica DCF 295 digital camera. Chromosome numbers were verified at least on five germinated seeds from five individuals and at least 10 cells per root tip.

3. Results

3.1. Phylogenetic Analysis

The majority-rule consensus resulting from the Bayesian analysis of dataset 1 (ETS region) is illustrated in Figure 4. The outline of the main geographical clades suggested by [46] is confirmed: there is a Balkan and Eurosiberian clade; an Iberian clade, including one species (C. xaveri) from Africa; a North African clade including some Iberian species, sister to an Aegean clade; and an Anatolian–Iranian clade sister to the rest of clades (Figure 4). The monophyly of the Balkan clade is also well supported. This group of taxa is more related to the Eurasian species than to the Aegean and Anatolian taxa because the representatives of the Eurasian species, namely, C. cephalariifolia, C. glehnii, C. legionis-septimae, and C. scabiosa, are nested in the clade. This inclusion, together with the addition of Balkan C. jankae, makes the C. calocephala complex sensu stricto not monophyletic.
Within the Balkan clade (incl. the C. calocephala complex), the sister species to the rest are successively the clade of C. salonitana and C. rupestris, two widely distributed species in the region and also in Greece (C. salonitana has a very wide distribution reaching from Turkey to the Balkans); then, C. ragusina, a narrow endemic from Croatia. The tree also shows that C. calocephala has several different ribotypes.

3.2. Network Analyses

After checking the presence of recombinant sequences using RDP software [80] and discarding only one sequence for the AGT1 dataset, we built two networks. The high number of ties detected in the analyses of the two regions agrees with previous evidence of the intense hybridization and introgression in sect. Acrocentron, a fact that confounds any interpretation of networks. In fact, almost none of the species with more than one population in the analysis was distributed in a single group in the networks (Figure 5 and Figure 6). Despite the fact that the results are difficult to interpret, two groups could be defined in the ETS network. The first one is the Calocephala 1 group (Figure 5), including clones from several populations of C. calocephala (CA1, CA2, CA5, CA6, CA24, CA25, CA26, CA28, and CA29), then C. calocephala var. mixta (CAF11), C. chrysolepis (CH), C. crnogorica (CR), C. grbavacensis f. lutea (GRF8), C. kotschyana (KO18 and KO19), C. melanocephala (ME), and C. orientalis (OR). The Calocephala 2 group includes clones of C. calocephala (CA1, CA5, CA6, CA7, CA24, CA26, and CA29), C. calocephala var. flava (CAF10 and CAF31), C. calocephala var. mixta (CAF11), C. chrysolepis (CH), C. crnogorica (CR), C. gjurasinii (GJ13 and GJ23), C. grbavacensis (GR3), C. grbavacensis f. lutea (GRF8), C. immanuelis-loewii (IM), C. melanocephala (ME), C. orientalis (OR), and C. zlatiborensis (ZL); see Table S2 for population codes.
Regarding the AGT1 network, it also shows intense reticulation. As was the case in the ETS, most populations have sequences that are intermixed within the network (Figure 6).

3.3. Chromosome Numbers

All investigated populations were diploid with x = 10 or x = 11. Chromosome numbers of C. gjurasinii with 2n = 20 and C. melanocephala with 2n = 22 were determined for the first time. Chromosome numbers in eight populations of C. calocephala from different parts of the investigated area (Croatia, Romania, and Serbia) were in all cases 2n = 20 (Table S2). For C. kotschyana, the chromosome number was 2n = 22 in the investigated population from Serbia compared to 2n = 20 and 2n = 22 available in the literature (Table S2). For C. murbeckii, we counted 2n = 22 in the population from Bosnia and Herzegovina differing from 2n = 20 published by [86].

4. Discussion

4.1. Utility of Nuclear-Ribosomal ETS and Low-Copy AGT1 in Acrocentron

The backbone of the ETS tree supports the main geographical clades of sect. Acrocentron defined by [46] and confirms the monophyly of the Balkan + Eurasian clade (Figure 4). However, support is low within the major clades and (especially in the Balkan clade) some species show multiple copies placed in different subclades (Figure 4). Two reasons could explain the low resolution: recent speciation and introgression. Not many changes accumulate in sequences of fast-evolving spacers (ETS) in cases of fast speciations or when the species are very recently diverged; in our case, for such closely related species, the ETS gene tree had not yet attained reciprocal monophyly for the constituent species [87]. As for introgression, the ETS network is very illustrative, and ties are especially evident in Group 2 (Figure 5). However, given the multi-copy character of the rRNA-cistron, this conclusion should be taken with caution.
Low-copy AGT1 showed fewer changes than the ETS, and the resulting phylogram was not informative (tree not shown). Ties were found to be especially intense in the network (Figure 6), to the extreme of blurring differences between groups. However, AGT1 would help in suggesting the parentals of C. zlatiborensis better than the ETS, as discussed below.

4.2. Phylogeny of the Balkan Clade: Introgression and Gene Flow

Focusing on the results in the Balkan clade, Centaurea salonitana and C. rupestris are widely distributed in the south of the Balkans and are at the origin of the clade. In addition, at the base of the Balkan clade, we also find C. ragusina, an endemic Croatian species that inhabits crevices of limestone cliffs near the coast and islands of the Adriatic Sea. The rest of the Balkan clade is a polytomy with some well-supported subclades, which, however, do not correlate with current systematics. In fact, only C. kotschyana forms a well-supported subclade, while other subclades represent a mixture of morphologically and taxonomically well-defined species. For example, subclade B1 in Figure 4 is composed of cloned ribotypes from C. calocephala, C. zlatiborensis, C. gjurasinii, and C. crnogorica (Figure 1), while subclade B2 in Figure 4 is composed of cloned rybotypes of C. calocephala, C. grbavacensis, C. melanocephala, and C. orientalis from the Centaurea calocephala complex (Figure 1 and Figure 2), and Iberian C. cephalariifolia and C. legionis-septimae, as well as an Italian population of C. scabiosa. Furthermore, most of the species of our study are placed in different subclades because they have more than one clone. In the ETS phylogeny, C. calocephala is placed in four different subclades; C. orientalis in three; and C. cephalariifolia, C. crnogorica, C. gjurasinii, C. grbavacensis, C. melanocephala, C. scabiosa, and C. zlatiborensis in two subclades (Figure 4). There are only two exceptions: individual CAF31 of C. calocephala var. flava shows only one clone close to C. grbavacensis clones, and population CA2 of C. calocephala shows only one clone (Figure 5). All this indicates that the morphologically well-delimited species from the Centaurea calocephala complex cannot be defined on a molecular basis, nor does the phylogeny of this group of species match the accepted systematics. This is not an isolated case in which molecular investigation thus far did not allow to a better separation of species. There are some other examples in the Balkans, such as Knautia [88], Heliosperma [89], and Sempervivum [90], where hybridization and reticulate evolution complicate matters when it comes to the field of systematics.
The fact that species of the Balkan clade have more than one clone is probably the result of hybridization events, even though other explanations are possible. If our hypothesis of the multiple clones were acquired by hybridization is true, this implies the presence of rampant ancient homoploid hybridization between species not currently sympatric (they are diploid with 2n = 20 or 2n = 22, Table S2). According to [51], altitude shifts caused by glaciations in the Iberian Peninsula allowed mutual contacts of taxa from the sect. Acrocentron currently located in the mountains, and this could also be the case in the C. calocephala complex. Glaciations could have had a major impact on the evolution and phylogeny of the group, because according to [91,92], C. sect. Acrocentron originated in the Pliocene. This must be the case in the Balkans, which were a refuge for many species during the Ice Age, as evidenced by the great species diversity. It is a well-known case for many plant groups that during glaciations, populations migrated into refugia and became temporarily sympatric [93]. For example, isolated populations of Cardamine maritima came into contact during the glaciation along the Balkan coast and in mountain massifs, which caused gene flow between species [94]. There are also similar hypotheses of hybridization during glaciations in species of Fraxinus [95], Edraianthus [28], or Sesleria [33].
The question of whether molecular methods may clarify the origin of hybrid taxa is exemplified in C. zlatiborensis. Morphology suggests that C. zlatiborensis is closely related to C. calocephala [47,48]. Both networks place clones of C. zlatiborensis in different positions: one clone is always placed close to C. grbavacensis and the other among the C. calocephala clones (Figure 5 and Figure 6). The AGT1 gene is very precise in pointing to C. kotschyana as one of the parental lineages of C. zlatiborensis. The other parental lineage would be C. grbavacensis. Present distributions of C. kotchyana and C. grbavacensis do not overlap, but they probably were in contact in a refugium during glaciations and likely gave rise to hybridogenic C. zlatiborensis. After glaciations, the range of the hybrid changed in relation to the parent species, and C. zlatiborensis probably colonized the Dinaric Alps in the western part of Serbia during the postglacial period.
During the field research, we found some “mixed” populations of C. calocephala with individuals with red (var. atropurpurea), yellow (var. flava), and intermediate red-yellow/yellow-red (var. mixta) capitula, such as the populations from Đerdap (CAF10, CAF11, and CA6) and Lika (CAF31 and CA7), as well as the populations of C. grbavacensis from Northern Macedonia (Suva Gora, GRF8, and GR3). These taxa are well-differentiated on a morphological basis, and some are intermediate (specifically var. mixta), with differences in size, shape, and position of involucral bracts, probably all of them formed by homoploid hybridization involving C. calocephala and an unknow species with yellow florets. Supporting this hypothesis, all these populations have a clone close to C. grbavacensis clones in the AGT1 network (Figure 6).
A frequently asked question is where to place the boundary between hybrids and hybridogenic species. F1 generation hybrids can be defined as the offspring created by crossing individuals that belong to different populations, have different adaptive abilities, and their existence depends completely on the parent taxa [96]. According to the biological concept of species, hybrids should be considered “good” species if they are reproductively or geographically isolated [97]. We can find hybrid zones formed by F1 generation hybrids and individuals with different recombinations that are active sources of new evolutionary types and species [98]. The problem occurs in hybrid zones when hybrids stabilize and produce fertile offspring, but they are not completely reproductively isolated from their parents, i.e., they can interbreed with them. This is an obvious issue when dealing with homoploid hybrids. The main problem is the historical concept of species, which are treated as static entities [99] and not as a dynamic relationship between ancestors and descendants, i.e., an evolutionary line [100,101]. The question is whether hybrids should be considered hybrids only when they are completely reproductively isolated from their parents and produce fertile offspring. Hybrid zones can be very significant and provide a wealth of information about the status and levels of populations that are “on the way” to becoming new species [98].

5. Conclusions

(a)
The Balkan–Eurasian clade of sect. Acrocentron is a natural group.
(b)
The C. calocephala complex as currently defined is not a monophyletic group, given that C. cephalariifolia, C. glehnii, C. legionis-septimae, and C. scabiosa are nested between other species of the complex. All these species should be included in any future study of the complex. Species are well-delimited from a morphological point of view, and they are the result of allopatric speciation in the Balkan Peninsula. In these morphologically well-defined species, successive hybridization events accompanied by introgression and gene flow caused by latitudinal and altitudinal shifts in the transits from glacial to interglacial periods were superimposed.
(c)
As a result of this introgression, any reconstruction of the species’ boundaries using sequence data is problematic within the group, and the use of other, more resolving markers should be considered.

Supplementary Materials

The following supporting information can be downloaded at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/d14050394/s1, Table S1: General chorological and ecological characteristics of analyzed representatives of Centaurea calocephala complex from the Balkans. Table S2: List of investigated species, origin of the plant materials, with voucher numbers of herbarium specimens, chromosome numbers, and GenBank accession numbers [42,43,48,49,51,86,102,103,104,105,106,107,108,109,110,111,112,113].

Author Contributions

Conceptualization, N.G.-J., A.S. and P.J.; methodology, N.G.-J.; investigation, J.N., P.J., M.L., D.L., P.D.M., I.B., S.M., B.Z. and N.G.-J.; writing—original draft preparation, J.N.; writing—review and editing, A.S., N.G.-J., P.J., M.L., D.L., P.D.M., I.B., S.M. and B.Z.; funding acquisition, P.J., P.D.M., N.G.-J. and A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Serbian Ministry of Education, Science and Technological Development, grant no. 451-03-68/2020-14/200178, and the Catalan Government (“Ajuts a grups consolidats” 2017-SGR1116).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data other than DNA sequences are available from the authors by direct request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Turrill, W.B. The Plant-Life of the Balkan Peninsula: A Phytogeographical Study; Clarendon: Oxford, UK, 1929. [Google Scholar]
  2. Stevanović, V. Exploration of Balkan Flora after Turrill’s time—the current situation and future challenges. In Proceedings of the 5th Balkan Botanical Congress: Book of Abstracts; Faculty of Biology, University of Belgrade & Serbian Academy of Sciences and Arts: Belgrade, Serbia, 2009; p. 9. [Google Scholar]
  3. Stevanović, V. Analysis of the central European and Mediterranean orophytic element on the mountains of the w. and central Balkan Peninsula, with special reference to endemics. Bocconea 1996, 5, 77–97. [Google Scholar]
  4. Savić, I.R. Diversification of the Balkan fauna: Its origin, historical development and present status. In Advances in Arachnology and Developmental Biology; Makarov, E., Dimitrijević, R.E., Eds.; SASA Belgrade UNESCO MAB Serbia: Belgrade, Serbia, 2008; pp. 57–78. [Google Scholar]
  5. Hewitt, G.M. Post-glacial re-colonization of European biota. Biol. J. Linn. Soc. 1999, 68, 87–112. [Google Scholar] [CrossRef]
  6. Hewitt, G. The genetic legacy of the Quaternary Ice Ages. Nature 2000, 405, 907–913. [Google Scholar] [CrossRef] [PubMed]
  7. Petit, R.J.; Aguinagalde, I.; de Beaulieu, J.-L.; Bittkau, C.; Brewer, S.; Cheddadi, R.; Ennos, R.; Fineschi, S.; Grivet, D.; Lascoux, M. Glacial Refugia: Hotspots but not melting pots of genetic diversity. Science 2003, 300, 1563–1565. [Google Scholar] [CrossRef] [Green Version]
  8. Tzedakis, P.C. The Balkans as prime glacial refugial territory of European temperate trees. In Balkan Biodiversity; Griffiths, H.I., Kryštufek, B., Reed, J.M., Eds.; Springer: Dordrecht, The Netherlands, 2004; pp. 49–68. [Google Scholar] [CrossRef]
  9. Médail, F.; Diadema, K. Glacial refugia influence plant diversity patterns in the Mediterranean Basin. J. Biogeogr. 2009, 36, 1333–1345. [Google Scholar] [CrossRef]
  10. Surina, B.; Schönswetter, P.; Schneeweiss, G.M. Quaternary range dynamics of ecologically divergent species (Edraianthus serpyllifolius and E. tenuifolius, Campanulaceae) within the Balkan refugium. J. Biogeogr. 2011, 38, 1381–1393. [Google Scholar] [CrossRef]
  11. Lakušić, D.; Liber, Z.; Nikolić, T.; Surina, B.; Kovačić, S.; Bogdanović, S.; Stefanović, S. Molecular phylogeny of the Campanula pyramidalis species complex (Campanulaceae) inferred from chloroplast and nuclear non-coding sequences and its taxonomic implications. Taxon 2013, 62, 505–524. [Google Scholar] [CrossRef] [Green Version]
  12. Janković, I.; Šatović, Z.; Liber, Z.; Kuzmanović, N.; Radosavljević, I.; Lakušić, D. Genetic diversity and morphological variability in the Balkan endemic Campanula secundiflora s.l. (Campanulaceae). Bot. J. Linn. Soc. 2016, 180, 64–88. [Google Scholar] [CrossRef] [Green Version]
  13. Janković, I.; Satovic, Z.; Liber, Z.; Kuzmanović, N.; Di Pietro, R.; Radosavljević, I.; Nikolov, Z.; Lakušić, D. Genetic and morphological data reveal new insights into the taxonomy of Campanula versicolor s.l. (Campanulaceae). Taxon 2019, 68, 340–369. [Google Scholar] [CrossRef]
  14. Ronikier, M.; Zalewska-Gałosz, J. Independent evolutionary history between the Balkan Ranges and more northerly mountains in Campanula alpina s.l. (Campanulaceae): Genetic divergence and morphological segregation of taxa. Taxon 2014, 63, 116–131. [Google Scholar] [CrossRef]
  15. Bogdanović, S.; Rešetnik, I.; Brullo, S.; Shuka, L. Campanula aureliana (Campanulaceae), a new species from Albania. Plant Syst. Evol. 2015, 301, 1555–1567. [Google Scholar] [CrossRef]
  16. Aleksić, J.M.; Škondrić, S.; Lakušić, D. Comparative phylogeography of capitulate Campanula species from the Balkans, with description of a new species, C. daucoides. Plant Syst. Evol. 2018, 304, 549–575. [Google Scholar] [CrossRef]
  17. Bogdanović, S.; Brullo, S.; Rešetnik, I.; Lakušić, D.; Satovic, Z.; Liber, Z. Campanula skanderbegii: Molecular and morphological evidence of a new Campanula species (Campanulaceae) endemic to Albania. Syst. Bot. 2014, 39, 1250–1260. [Google Scholar] [CrossRef]
  18. Grdiša, M.; Liber, Z.; Radosavljević, I.; Carović-Stanko, K.; Kolak, I.; Satovic, Z. Genetic diversity and structure of Dalmatian Pyrethrum (Tanacetum cinerariifolium Trevir./Sch./Bip., Asteraceae) within the Balkan refugium. PLoS ONE 2014, 9, e105265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Hardion, L.; Baumel, A.; Verlaque, R.; Vila, B. Distinct evolutionary histories of lowland biota on Italian and Balkan Peninsulas revealed by the phylogeography of Arundo plinii (Poaceae). J. Biogeogr. 2014, 41, 2150–2161. [Google Scholar] [CrossRef]
  20. Perný, M.; Tribsch, A.; Anchev, M.E. Infraspecific differentiation in the Balkan diploid Cardamine acris (Brassicaceae): Molecular and morphological evidence. Folia Geobot. 2004, 39, 405–429. [Google Scholar] [CrossRef]
  21. Lakušić, D.; Novčić, R.; Kučera, J.; Marhold, K. Cardamine pancicii Hayek (Brassicaceae), a neglected species of the Balkan Peninsula: Morphological and molecular evidence. Willdenowia 2006, 36, 177–191. [Google Scholar] [CrossRef]
  22. Kučera, J.; Tremetsberger, K.; Vojta, J.; Marhold, K. Molecular study of the Cardamine maritima group (Brassicaceae) from the Balkan and Apennine Peninsulas based on amplified fragment length polymorphism. Plant Syst. Evol. 2008, 275, 193–207. [Google Scholar] [CrossRef]
  23. Kutnjak, D.; Kuttner, M.; Niketić, M.; Dullinger, S.; Schönswetter, P.; Frajman, B. Escaping to the summits: Phylogeography and predicted range dynamics of Cerastium dinaricum, an endangered high mountain plant endemic to the western Balkan Peninsula. Mol. Phylogenetics Evol. 2014, 78, 365–374. [Google Scholar] [CrossRef]
  24. Đurović, S.Z.; Temunović, M.; Niketić, M.; Tomović, G.; Schönswetter, P.; Frajman, B.; Lavergne, S. Impact of Quaternary climatic oscillations on phylogeographic patterns of three habitat-segregated Cerastium taxa endemic to the Dinaric Alps. J. Biogeogr. 2021, 48, 2022–2036. [Google Scholar] [CrossRef]
  25. Stefanović, S.; Lakušić, D.; Kuzmina, M.; Međedović, S.; Tan, K.; Stevanović, V. Molecular phylogeny of Edraianthus (Grassy Bells; Campanulaceae) based on non-coding plastid DNA sequences. Taxon 2008, 57, 452–475. [Google Scholar]
  26. Lakušić, D.; Rakić, T.; Stefanović, S.; Surina, B.; Stevanović, V. Edraianthus × lakusicii (Campanulaceae) a new intersectional natural hybrid: Morphological and molecular evidence. Plant Syst. Evol. 2009, 280, 77–88. [Google Scholar] [CrossRef]
  27. Lakušić, D.; Stefanović, S.; Siljak-Yakovlev, S.; Rakić, T.; Kuzmanović, N.; Surina, B. Edraianthus stankovicii (Campanulaceae), an overlooked taxon from the Balkan Peninsula—Evidence from morphometric, molecular and genome size studies. Phytotaxa 2016, 269, 69. [Google Scholar] [CrossRef]
  28. Surina, B.; Rakić, T.; Stefanović, S.; Stevanović, V.; Lakušić, D. One new species of the genus Edraianthus, and a change in taxonomic status for Edraianthus serpyllifolius f. pilosulus (Campanulaceae) from the Balkan Peninsula. Syst. Bot. 2009, 34, 602–608. [Google Scholar] [CrossRef]
  29. Jug-Dujaković, M.; Ninčević, T.; Liber, Z.; Grdiša, M.; Šatović, Z. Salvia officinalis survived in situ Pleistocene glaciation in ‘refugia within refugia’ as inferred from AFLP markers. Plant Syst. Evol. 2020, 306, 38. [Google Scholar] [CrossRef]
  30. Stojanović, D.; Aleksić, J.M.; Jančić, I.; Jančić, R. A Mediterranean medicinal plant in the Continental Balkans: A plastid DNA-based phylogeographic survey of Salvia officinalis (Lamiaceae) and its conservation implications. Willdenowia 2015, 45, 103–118. [Google Scholar] [CrossRef]
  31. Buzurović, U.; Tomović, G.; Niketić, M.; Bogdanović, S.; Aleksić, J.M. Phylogeographic and taxonomic considerations on Goniolimon tataricum (Plumbaginaceae) and its relatives from south-eastern Europe and the Apennine Peninsula. Plant Syst. Evol. 2020, 306, 29. [Google Scholar] [CrossRef]
  32. Kuzmanović, N.; Comanescu, P.; Frajman, B.; Lazarević, M.; Paun, O.; Schönswetter, P.; Lakušić, D. Genetic, cytological and morphological differentiation within the Balkan-Carpathian Sesleria rigida sensu Fl. Eur. (Poaceae): A taxonomically intricate tetraploid-octoploid complex. Taxon 2013, 62, 458–472. [Google Scholar] [CrossRef] [Green Version]
  33. Kuzmanović, N.; Lakušić, D.; Frajman, B.; Alegro, A.; Schönswetter, P. Phylogenetic relationships in Seslerieae (Poaceae) including resurrection of Psilathera and Sesleriella, two monotypic genera endemic to the Alps. Taxon 2017, 66, 1349–1370. [Google Scholar] [CrossRef]
  34. López-González, N.; Bobo-Pinilla, J.; Padilla-García, N.; Loureiro, J.; Castro, S.; Rojas-Andrés, B.M.; Martínez-Ortega, M.M. Genetic similarities versus morphological resemblance: Unraveling a polyploid complex in a Mediterranean biodiversity hotspot. Mol. Phylogenetics Evol. 2021, 155, 107006. [Google Scholar] [CrossRef]
  35. Rundel, P.W.; Arroyo, M.T.K.; Cowling, R.M.; Keeley, J.E.; Lamont, B.B.; Vargas, P. Mediterranean biomes: Evolution of their vegetation, floras, and climate. Annu. Rev. Ecol. Evol. Syst. 2016, 47, 383–407. [Google Scholar] [CrossRef]
  36. Susanna, A.; Garcia-Jacas, N. Tribe Cardueae Cass. (1819). In Flowering Plants. Eudicots: Asterales; Kadereit, J.W., Jeffrey, C., Eds.; Springer: New York, NY, USA, 2007; Volume 8, pp. 123–146. ISBN 3-540-31051-7. [Google Scholar]
  37. Gardou, C. Quelques Vues Synthétiques sur les Centaurées de la Section Acrocentron (Cass.) O. Hoffm. Dans la Flore Méditerranéenne; La flore du bassin méditerranéen; Centre National de la Recherche Scientifique: Paris, France, 1975; pp. 537–547.
  38. Wagenitz, G. Centaurea L. In Flora of Turkey and the East Aegean Islands; Davis, P.H., Ed.; Edinburgh University Press: Edinburgh, UK, 1975; Volume 5, pp. 465–585. [Google Scholar]
  39. Font, M.; Garnatje, T.; Garcia-Jacas, N.; Susanna, A. Delineation and phylogeny of Centaurea sect. Acrocentron based on DNA sequences: A restoration of the genus Crocodylium and indirect evidence of introgression. Plant Syst. Evol. 2002, 234, 15–26. [Google Scholar] [CrossRef]
  40. Cassini, H. Dictionaire de Sciences Naturelles, Paris, 1819. In Cassini on Compositae; King, R., Dawson, H.W., Eds.; Oriole Editions: New York, NY, USA, 1975. [Google Scholar]
  41. Wagenitz, G. Pollenmorphologie und Systematik in der Gattung Centaurea L. s. 1. Flora Allg. Bot. 1955, 142, 213–279. [Google Scholar] [CrossRef]
  42. Garcia-Jacas, N.; Susanna, A. Karyological notes on Centaurea sect. Acrocentron (Asteraceae). Plant Syst. Evol. 1992, 179, 1–18. [Google Scholar] [CrossRef]
  43. Garcia-Jacas, N.; Susanna, A.; Mozaffarian, V. New chromosome counts in the subtribe Centaureinae (Asteraceae, Cardueae) from West Asia, III. Bot. J. Linn. Soc. 1998, 128, 413–422. [Google Scholar] [CrossRef]
  44. Ghaffari, S.M. Chromosome studies of some species of Centaurea section Acrocentron (Asteraceae) from Iran. Pak. J. Bot. 1999, 31, 301–305. [Google Scholar]
  45. Martin, E.; Dinc, M.; Duran, A. Karyomorphological study of eight Centaurea L. taxa (Asteraceae) from Turkey. Turk. J. Bot. 2009, 33, 97–104. [Google Scholar]
  46. Font, M.; Garcia-Jacas, N.; Vilatersana, R.; Roquet, C.; Susanna, A. Evolution and biogeography of Centaurea section Acrocentron inferred from nuclear and plastid DNA sequence analyses. Ann. Bot. 2009, 103, 985–997. [Google Scholar] [CrossRef]
  47. Novaković, J. Morphological, Phytochemical and Molecular Studies of Carpathian-Balkan Complex Centaurea atropurpurea (Asteraceae)—Phylogenetic and Taxonomic Implications. Ph.D. Thesis, Faculty of Biology, University of Belgrade, Belgrade, Serbia, 2019. [Google Scholar]
  48. Novaković, J.; Zlatković, B.; Lazarević, M.; Garcia-Jacas, N.; Susanna, A.; Marin, P.D.; Lakušić, D.; Janaćković, P. Centaurea zlatiborensis (Asteraceae, Cardueae−Centaureinae), a new endemic species from Zlatibor mountain range, Serbia. Nord. J. Bot. 2018, 36, 1–8. [Google Scholar] [CrossRef]
  49. Garcia-Jacas, N. Centaurea kunkelii (Asteraceae, Cardueae), a new hybridogenic endecaploid species of sect. Acrocentron from Spain. Ann. Bot. Fenn. 1998, 159–167. [Google Scholar]
  50. Garcia-Jacas, N.; Susanna, A. Centaurea × polymorpha Lagasca: Los problemas de un híbrido. Fontqueria 1993, 36, 65–66. [Google Scholar]
  51. Garcia-Jacas, N.; Soltis, P.S.; Font, M.; Soltis, D.E.; Vilatersana, R.; Susanna, A. The polyploid series of Centaurea toletana: Glacial migrations and introgression revealed by nrDNA and cpDNA sequence analyzes. Mol. Phylogenetics Evol. 2009, 52, 377–394. [Google Scholar] [CrossRef] [PubMed]
  52. Hilpold, A.; Vilatersana, R.; Susanna, A.; Meseguer, A.S.; Boršić, I.; Constantinidis, T.; Filigheddu, R.; Romaschenko, K.; Suárez-Santiago, V.N.; Tugay, O.; et al. Phylogeny of the Centaurea group (Centaurea, Compositae)—Geography is a better predictor than morphology. Mol. Phylogenetics Evol. 2014, 77, 195–215. [Google Scholar] [CrossRef] [PubMed]
  53. Rieseberg, L.H. Hybrid origins of plant species. Annu. Rev. Ecol. Syst. 1997, 28, 359–389. [Google Scholar] [CrossRef] [Green Version]
  54. Mallet, J. Hybrid speciation. Nature 2007, 446, 279–283. [Google Scholar] [CrossRef] [PubMed]
  55. Soltis, P.S.; Soltis, D.E. The role of hybridization in plant speciation. Annu. Rev. Plant Biol. 2009, 60, 561–588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Chapman, M.A.; Burke, J.M. Genetic divergence and hybrid speciation. Evolution 2007, 61, 1773–1780. [Google Scholar] [CrossRef]
  57. Paun, O.; Forest, F.; Fay, M.F.; Chase, M.W. Hybrid speciation in angiosperms: Parental divergence drives ploidy. New Phytol. 2009, 182, 507–518. [Google Scholar] [CrossRef] [Green Version]
  58. Brennan, A.C.; Barker, D.; Hiscock, S.J.; Abbott, R.J. Molecular genetic and quantitative trait divergence associated with recent homoploid hybrid speciation: A study of Senecio squalidus (Asteraceae). Heredity 2012, 108, 87–95. [Google Scholar] [CrossRef] [Green Version]
  59. Koutecký, P.; Badurová, T.; Štech, M.; Košnar, J.; Karásek, J. Hybridization between diploid Centaurea pseudophrygia and tetraploid C. jacea (Asteraceae): The role of mixed pollination, unreduced gametes, and mentor effects. Biol. J. Linn. Soc. 2011, 104, 93–106. [Google Scholar] [CrossRef] [Green Version]
  60. Mameli, G.; López-Alvarado, J.; Farris, E.; Susanna, A.; Filigheddu, R.; Garcia-Jacas, N. The role of parental and hybrid species in multiple introgression events: Evidence of homoploid hybrid speciation in Centaurea (Cardueae, Asteraceae). Bot. J. Linn. Soc. 2014, 175, 453–467. [Google Scholar] [CrossRef] [Green Version]
  61. Pisanu, S.; Mameli, G.; Farris, E.; Binelli, G.; Filigheddu, R. A natural homoploid hybrid between Centaurea horrida and Centaurea filiformis (Asteraceae) as revealed by morphological and genetic traits. Folia Geobot. 2011, 46, 69–86. [Google Scholar] [CrossRef]
  62. Bancheva, S.; Geraci, A.; Raimondo, F.M. Genetic diversity in the Centaurea cineraria group (Compositae) in Sicily using isozymes. Plant Biosyst. 2006, 140, 10–16. [Google Scholar] [CrossRef]
  63. Albrecht, M.; Duelli, P.; Obrist, M.K.; Kleijn, D.; Schmid, B. Effective long-distance pollen dispersal in Centaurea jacea. PLoS ONE 2009, 4, e6751. [Google Scholar] [CrossRef] [PubMed]
  64. Moreyra, L.D.; Márquez, F.; Susanna, A.; Garcia-Jacas, N.; Vázquez, F.M.; López-Pujol, J. Genesis, evolution, and genetic diversity of the hexaploid, narrow endemic Centaurea tentudaica. Diversity 2021, 13, 72. [Google Scholar] [CrossRef]
  65. Small, R.L.; Cronn, R.C.; Wendel, J.F. Use of nuclear genes for phylogeny reconstruction in plants. Aust. Syst. Bot. 2004, 17, 145. [Google Scholar] [CrossRef]
  66. Slovák, M.; Kučera, J.; Záveská, E.; Vd’ačný, P. Dealing with discordant genetic signal caused by hybridisation, incomplete lineage sorting and paucity of primary nucleotide homologies: A case study of closely related members of the genus Picris Subsection Hieracioides (Compositae). PLoS ONE 2014, 9, e104929. [Google Scholar] [CrossRef] [Green Version]
  67. Wan, D.; Sun, Y.; Zhang, X.; Bai, X.; Wang, J.; Wang, A.; Milne, R. Multiple ITS copies reveal extensive hybridization within Rheum (Polygonaceae), a genus that has undergone rapid radiation. PLoS ONE 2014, 9, e89769. [Google Scholar] [CrossRef] [Green Version]
  68. Oberprieler, C.; Greiner, R.; Konowalik, K.; Vogt, R. The reticulate evolutionary history of the polyploid NW Iberian Leucanthemum pluriflorum clan (Compositae, Anthemideae) as inferred from NrDNA ETS sequence diversity and eco-climatological niche-modelling. Mol. Phylogenetics Evol. 2014, 70, 478–491. [Google Scholar] [CrossRef]
  69. Doyle, J.J.; Dickson, E.E. Preservation of plant samples for DNA restriction endonuclease analysis. Taxon 1987, 36, 715–722. [Google Scholar] [CrossRef]
  70. Cullings, K.W. Design and testing of a plant-specific PCR primer for ecological and evolutionary studies. Mol. Ecol. 1992, 1, 233–240. [Google Scholar] [CrossRef]
  71. Tel-Zur, N.; Abbo, S.; Myslabodski, D.; Mizrahi, Y. Modified CTAB procedure for DNA isolation from epiphytic cacti of the genera Hylocereus and Selenicereus (Cactaceae). Plant Mol. Biol. Report. 1999, 17, 249–254. [Google Scholar] [CrossRef]
  72. Linder, C.R.; Goertzen, L.R.; Heuvel, B.V.; Francisco-Ortega, J.; Jansen, R.K. The complete External Transcribed Spacer of 18S–26S RDNA: Amplification and phylogenetic utility at low taxonomic levels in Asteraceae and closely allied families. Mol. Phylogenetics Evol. 2000, 14, 285–303. [Google Scholar] [CrossRef] [PubMed]
  73. Baldwin, B.G.; Markos, S. Phylogenetic utility of the External Transcribed Spacer (ETS) of 18S–26S RDNA: Congruence of ETS and ITS Trees of Calycadenia (Compositae). Mol. Phylogenetics Evol. 1998, 10, 449–463. [Google Scholar] [CrossRef]
  74. Susanna, A.; Galbany-Casals, M.; Romaschenko, K.; Barres, L.; Martin, J.; Garcia-Jacas, N. Lessons from Plectocephalus (Compositae, Cardueae-Centaureinae): ITS Disorientation in Annuals and Beringian dispersal as revealed by molecular analyses. Ann. Bot. 2011, 108, 263–277. [Google Scholar] [CrossRef]
  75. López-Pujol, J.; Garcia-Jacas, N.; Susanna, A.; Vilatersana, R. Should we conserve pure species or hybrid species? Delimiting hybridization and introgression in the Iberian endemic Centaurea podospermifolia. Biol. Conserv. 2012, 152, 271–279. [Google Scholar] [CrossRef]
  76. Li, M.; Wunder, J.; Bissoli, G.; Scarponi, E.; Gazzani, S.; Barbaro, E.; Saedler, H.; Varotto, C. Development of COS genes as universally amplifiable markers for phylogenetic reconstructions of closely related plant species. Cladistics 2008, 24, 727–745. [Google Scholar] [CrossRef] [Green Version]
  77. Hall, T.A. BioEdit: A user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symp. Ser. 1999, 41, 95–98. [Google Scholar]
  78. Cline, J.; Braman, J.C.; Hogrefe, H.H. PCR Fidelity of pfu DNA polymerase and other thermostable DNA polymerases. Nucleic Acids Res. 1996, 24, 3546–3551. [Google Scholar] [CrossRef] [Green Version]
  79. Popp, M.; Oxelman, B. Inferring the history of the polyploid Silene aegaea (Caryophyllaceae) using plastid and homoeologous nuclear DNA sequences. Mol. Phylogenetics Evol. 2001, 20, 474–481. [Google Scholar] [CrossRef]
  80. Martin, D.P.; Murrell, B.; Golden, M.; Khoosal, A.; Muhire, B. RDP4: Detection and analysis of recombination patterns in virus genomes. Virus Evol. 2015, 1, vev003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Posada, D.J. ModelTest: Phylogenetic Model Averaging. Mol. Biol. Evol. 2008, 25, 1253–1256. [Google Scholar] [CrossRef] [PubMed]
  82. Ronquist, F.; Teslenko, M.; Van Der Mark, P.; Ayres, D.L.; Darling, A.; Höhna, S.; Larget, B.; Liu, L.; Suchard, M.A.; Huelsenbeck, J.P. MrBayes 3.2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Syst. Biol. 2012, 61, 539–542. [Google Scholar] [PubMed] [Green Version]
  83. Bryant, D.; Moulton, V. Neighbor-Net: An agglomerative method for the construction of phylogenetic networks. Mol. Biol. Evol. 2004, 21, 255–265. [Google Scholar] [CrossRef] [PubMed]
  84. Huson, D.H.; Bryant, D. Application of phylogenetic networks in evolutionary studies. Mol. Biol. Evol. 2006, 23, 254–267. [Google Scholar] [CrossRef] [PubMed]
  85. Feulgen, R.; Rossenbeck, H. Mikroskopisch-chemischer Nachweis einer Nucleinsäure vom Typus der Thymonucleinsäure und die-darauf beruhende elektive Färbung von Zellkernen in mikroskopischen Präparaten. Biol. Chem. 1924, 135, 203–248. [Google Scholar] [CrossRef]
  86. Siljak-Yakovlev, S. Etude Cytogénétique et Palynologique de Compositae Endémiques ou Reliques de la Flore Yougoslave. Ph.D. Thesis, Centre d’Orsay, Université Paris-Sud, Orsay, France, 1986. [Google Scholar]
  87. Kay, K.M.; Reeves, P.A.; Olmstead, R.G.; Schemske, D.W. Rapid speciation and the evolution of hummingbird pollination in neotropical Costus subgenus Costus (Costaceae): Evidence from NrDNA ITS and ETS sequences. Am. J. Bot. 2005, 92, 1899–1910. [Google Scholar] [CrossRef] [Green Version]
  88. Frajman, B.; Rešetnik, I.; Niketić, M.; Ehrendorfer, F.; Schönswetter, P. Patterns of rapid diversification in heteroploid Knautia sect. Trichera (Caprifoliaceae, Dipsacoideae), one of the most intricate taxa of the European flora. BMC Evol. Biol. 2016, 16, 204. [Google Scholar] [CrossRef] [Green Version]
  89. Frajman, B.; Eggens, F.; Oxelman, B. Hybrid origins and homoploid reticulate evolution within Heliosperma (Sileneae, Caryophyllaceae)—A multigene phylogenetic approach with relative dating. Syst. Biol. 2009, 58, 328–345. [Google Scholar] [CrossRef] [Green Version]
  90. Klein, J.T.; Kadereit, J.W. Allopatric hybrids as evidence for past range dynamics in Sempervivum (Crassulaceae), a western Eurasian high mountain oreophyte. Alp. Bot. 2016, 126, 119–133. [Google Scholar] [CrossRef]
  91. Barres, L.; Sanmartín, I.; Anderson, C.L.; Susanna, A.; Buerki, S.; Galbany-Casals, M.; Vilatersana, R. Reconstructing the evolution and biogeographic history of tribe Cardueae (Compositae). Am. J. Bot. 2013, 100, 867–882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Herrando-Moraira, S.; Calleja, J.A.; Galbany-Casals, M.; Garcia-Jacas, N.; Liu, J.-Q.; López-Alvarado, J.; López-Pujol, J.; Mandel, J.R.; Massó, S.; Montes-Moreno, N.; et al. Nuclear and plastid DNA phylogeny of tribe Cardueae (Compositae) with Hyb-Seq data: A new subtribal classification and a temporal diversification framework. Mol. Phylogenetics Evol. 2019, 137, 313–332. [Google Scholar] [CrossRef] [PubMed]
  93. Folk, R.A.; Soltis, P.S.; Soltis, D.E.; Guralnick, R. New prospects in the detection and comparative analysis of hybridization in the Tree of Life. Am. J. Bot. 2018, 105, 364–375. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Kučera, J.; Marhold, K.; Lihová, J. Cardamine maritima group (Brassicaceae) in the Amphi-Adriatic area: A hotspot of species diversity revealed by DNA sequences and morphological variation. Taxon 2010, 59, 148–164. [Google Scholar] [CrossRef]
  95. Heuertz, M.; Carnevale, S.; Fineschi, S.; Sebastiani, F.; Hausman, J.F.; Paule, L.; Vendramin, G.G. Chloroplast DNA phylogeography of European ashes, Fraxinus sp. (Oleaceae): Roles of hybridization and life history traits. Mol. Ecol. 2006, 15, 2131–2140. [Google Scholar] [CrossRef] [PubMed]
  96. Stebbins, G.L. The role of hybridization in evolution. Proc. Am. Philos. Soc. 1959, 103, 231–251. [Google Scholar]
  97. Mayr, E. Systematics and the Origin of Species; Columbia University Press: New York, NY, USA, 1942. [Google Scholar]
  98. Harrison, R.G. Hybrids and hybrid zones: Historical perspective. In Hybrid Zones and the Evolutionary Process; Oxford University Press: London, UK, 1993; pp. 3–12. [Google Scholar]
  99. Wilkins, J.S. Species: A History of the Idea; University of California Press: Berkeley, CA, USA, 2009; Volume 1, ISBN 0-520-94507-7. [Google Scholar]
  100. De Queiroz, K. The general lineage concept of species, species chteria, and the process of speciation. In Endless Forms: Species and Speciation; Oxford University Press: London, UK, 1998; pp. 57–75. [Google Scholar]
  101. De Queiroz, K. Species concepts and species delimitation. Syst. Biol. 2007, 56, 879–886. [Google Scholar] [CrossRef] [Green Version]
  102. Fernández Casas, F.J.; Susanna, A. Monografía de la sección Chamaecyanus Willk. del género Centaurea L. Treb. Inst. Bot. Barc. 1985, 10, 5–174. [Google Scholar]
  103. Tonian, T.R. Relation between chromosome number and some morphological features of Centaureinae Less representatives. Rev. Biol. 1980, 33, 552–554. [Google Scholar]
  104. Garcia-Jacas, N.; Susanna, A.; Vilatersana, R.; Guara, M. New chromosome counts in the subtribe Centaureinae (Asteraceae, Cardueae) from West Asia, II. Bot. J. Linn. Soc. 1998, 128, 403–412. [Google Scholar] [CrossRef]
  105. Stefureac, T.; Tacina, A. Cariological and chorological investigations on two endemic taxa in the Romanian flora. Acta Bot. Hort. Bucur. 1982, 111–116. [Google Scholar]
  106. Bancheva, S.; Greilhuber, J. Genome size in Bulgarian Centaurea s.l. (Asteraceae). Plant Syst. Evol. 2006, 257, 95–117. [Google Scholar] [CrossRef]
  107. Rice, A.; Glick, L.; Abadi, S.; Einhorn, M.; Kopelman, N.M.; Salman-Minkov, A.; Mayzel, J.; Chay, O.; Mayrose, I. The Chromosome Counts Database (CCDB)—A community resource of plant chromosome numbers. New Phytol. 2015, 206, 19–26. [Google Scholar] [CrossRef]
  108. Garcia-Jacas, N.; Susanna, A.; Ilarslan, R.; Ilarslan, H. New chromosome counts in the subtribe Centaureinae (Asteraceae, Cardueae) from West Asia. Bot. J. Linn. Soc. 1997, 125, 343–349. [Google Scholar] [CrossRef]
  109. Uysal, T.; Ertuğrul, K.; Susanna, A.; Garcia-Jacas, N. New chromosome counts in the genus Centaurea (Asteraceae) from Turkey. Bot. J. Linn. Soc. 2009, 159, 280–286. [Google Scholar] [CrossRef]
  110. Lovrić, A. IOPB Chromosome Number Reports 77. Taxon 1982, 31, 761–777. [Google Scholar]
  111. Routsi, E.; Georgiadis, T. Cytogeographical study of Centaurea L. sect. Acrocentron (Cass.) DC. (Asteraceae) in Greece. Bot. Helv. 1999, 109, 139–151. [Google Scholar]
  112. Devesa, J.A.; Valdés, B.; Ottonello, D. IOPB chromosome number reports 100. Taxon 1988, 37, 920. [Google Scholar]
  113. Garcia-Jacas, N. Estudi Taxonòmic i Biosistemàtic de les Espècies Ibèriques i Nord-Africanes del Gènere Centaurea sect. Acrocentron. Ph.D. Thesis, Universitat de Barcelona, Barcelona, Spain, 1992. [Google Scholar]
Figure 1. Comparison of representatives of the Centaurea calocephala complex from the Balkans (A,B) Centaurea calocephala; (C,D) Centaurea zlatiborensis; (E,F) Centaurea gjurasinii; (G,H) Centaurea crnogorica. Photos: (A,B) Bogdan Hurdu; (C,D,G,H) Pedja Janaćković; (E,F) Dmitar Lakušić.
Figure 1. Comparison of representatives of the Centaurea calocephala complex from the Balkans (A,B) Centaurea calocephala; (C,D) Centaurea zlatiborensis; (E,F) Centaurea gjurasinii; (G,H) Centaurea crnogorica. Photos: (A,B) Bogdan Hurdu; (C,D,G,H) Pedja Janaćković; (E,F) Dmitar Lakušić.
Diversity 14 00394 g001
Figure 2. Comparison of representatives of the Centaurea calocephala complex from the Balkans (continued). (A,B) Centaurea grbavacensis; (C,D) Centaurea melanocephala; (E,F) Centaurea orientalis. Photos: (A,B) Jelica Novaković; (C,D) Bojan Zlatković; (E,F) Pedja Janaćković.
Figure 2. Comparison of representatives of the Centaurea calocephala complex from the Balkans (continued). (A,B) Centaurea grbavacensis; (C,D) Centaurea melanocephala; (E,F) Centaurea orientalis. Photos: (A,B) Jelica Novaković; (C,D) Bojan Zlatković; (E,F) Pedja Janaćković.
Diversity 14 00394 g002
Figure 3. Distribution map of the studied populations of the C. calocephala complex. Country acronyms: ALB—Albania, AUT—Austria, BIH—Bosnia and Hercegovina, BGR—Bulgaria, HRV—Croatia, GRC—Greece, HUN—Hungary, ITA—Italy, MKD—North Macedonia, MNE—Montenegro, ROU—Romania, SRB—Serbia, SVN—Slovenia. Population codes are detailed in Table S2.
Figure 3. Distribution map of the studied populations of the C. calocephala complex. Country acronyms: ALB—Albania, AUT—Austria, BIH—Bosnia and Hercegovina, BGR—Bulgaria, HRV—Croatia, GRC—Greece, HUN—Hungary, ITA—Italy, MKD—North Macedonia, MNE—Montenegro, ROU—Romania, SRB—Serbia, SVN—Slovenia. Population codes are detailed in Table S2.
Diversity 14 00394 g003
Figure 4. Consensus phylogram obtained from 9000 Bayesian trees with higher posterior probability (PP) from ETS data with the Anatolian–Iranian clade as an outgroup. Population codes are detailed in Table S2. Abbreviations: cl = clon; Pop = population; con = consensus sequence.
Figure 4. Consensus phylogram obtained from 9000 Bayesian trees with higher posterior probability (PP) from ETS data with the Anatolian–Iranian clade as an outgroup. Population codes are detailed in Table S2. Abbreviations: cl = clon; Pop = population; con = consensus sequence.
Diversity 14 00394 g004
Figure 5. Neighbor-Net for the ETS region of the Centaurea calocephala complex. Population codes are detailed in Table S2. Abbreviations: cl = clon; con = consensus sequence. A dotted red line indicates the boundaries between the two groups.
Figure 5. Neighbor-Net for the ETS region of the Centaurea calocephala complex. Population codes are detailed in Table S2. Abbreviations: cl = clon; con = consensus sequence. A dotted red line indicates the boundaries between the two groups.
Diversity 14 00394 g005
Figure 6. Neighbor-Net for the AGT1 region of the Centaurea calocephala complex. Population codes are detailed in Table S2. Abbreviations: cl = clon; con = consensus sequence.
Figure 6. Neighbor-Net for the AGT1 region of the Centaurea calocephala complex. Population codes are detailed in Table S2. Abbreviations: cl = clon; con = consensus sequence.
Diversity 14 00394 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Novaković, J.; Janaćković, P.; Susanna, A.; Lazarević, M.; Boršić, I.; Milanovici, S.; Lakušić, D.; Zlatković, B.; Marin, P.D.; Garcia-Jacas, N. Molecular Insights into the Centaurea Calocephala Complex (Compositae) from the Balkans—Does Phylogeny Match Systematics? Diversity 2022, 14, 394. https://0-doi-org.brum.beds.ac.uk/10.3390/d14050394

AMA Style

Novaković J, Janaćković P, Susanna A, Lazarević M, Boršić I, Milanovici S, Lakušić D, Zlatković B, Marin PD, Garcia-Jacas N. Molecular Insights into the Centaurea Calocephala Complex (Compositae) from the Balkans—Does Phylogeny Match Systematics? Diversity. 2022; 14(5):394. https://0-doi-org.brum.beds.ac.uk/10.3390/d14050394

Chicago/Turabian Style

Novaković, Jelica, Pedja Janaćković, Alfonso Susanna, Maja Lazarević, Igor Boršić, Sretco Milanovici, Dmitar Lakušić, Bojan Zlatković, Petar D. Marin, and Núria Garcia-Jacas. 2022. "Molecular Insights into the Centaurea Calocephala Complex (Compositae) from the Balkans—Does Phylogeny Match Systematics?" Diversity 14, no. 5: 394. https://0-doi-org.brum.beds.ac.uk/10.3390/d14050394

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop