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

A comprehensive DNA barcode inventory of Austria’s fish species

  • Lukas Zangl ,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Project administration, Software, Supervision, Validation, Visualization, Writing – original draft, Writing – review & editing

    lukas.zangl@uni-graz.at

    Affiliations Institute of Biology, University of Graz, Graz, Austria, Universalmuseum Joanneum, Studienzentrum Naturkunde, Graz, Austria

  • Sylvia Schäffer,

    Roles Data curation, Methodology, Validation, Writing – original draft, Writing – review & editing

    Affiliation Institute of Biology, University of Graz, Graz, Austria

  • Daniel Daill,

    Roles Data curation, Methodology, Project administration, Validation, Writing – original draft, Writing – review & editing

    Affiliations Institute of Biology, University of Graz, Graz, Austria, Consultants in Aquatic Ecology and Engineering—blattfisch e.U., Wels, Austria

  • Thomas Friedrich,

    Roles Data curation, Formal analysis, Validation, Writing – original draft, Writing – review & editing

    Affiliation Institute of Hydrobiology and Aquatic Ecosystem Management, University of Natural Resources and Life Sciences, Vienna, Austria

  • Wolfgang Gessl,

    Roles Data curation, Validation, Visualization, Writing – original draft, Writing – review & editing

    Affiliation Institute of Biology, University of Graz, Graz, Austria

  • Marija Mladinić,

    Roles Data curation, Validation, Writing – original draft, Writing – review & editing

    Affiliation Department of Biology, Faculty of Science, University of Zagreb, Zagreb, Croatia

  • Christian Sturmbauer,

    Roles Funding acquisition, Project administration, Resources, Supervision, Validation, Writing – original draft, Writing – review & editing

    Affiliation Institute of Biology, University of Graz, Graz, Austria

  • Josef Wanzenböck,

    Roles Data curation, Investigation, Project administration, Validation, Writing – original draft, Writing – review & editing

    Affiliation Research Department for Limnology, Mondsee, University of Innsbruck, Mondsee, Austria

  • Steven J. Weiss,

    Roles Conceptualization, Data curation, Investigation, Project administration, Supervision, Validation, Writing – original draft, Writing – review & editing

    Affiliation Institute of Biology, University of Graz, Graz, Austria

  • Stephan Koblmüller

    Roles Conceptualization, Data curation, Funding acquisition, Methodology, Project administration, Resources, Software, Supervision, Validation, Visualization, Writing – original draft, Writing – review & editing

    Affiliation Institute of Biology, University of Graz, Graz, Austria

Abstract

Austria is inhabited by more than 80 species of native and non-native freshwater fishes. Despite considerable knowledge about Austrian fish species, the latest Red List of threatened species dates back 15 years and a systematic genetic inventory of Austria’s fish species does not exist. To fulfill this deficit, we employed DNA barcoding to generate an up-to-date and comprehensive genetic reference database for Austrian fish species. In total, 639 newly generated cytochrome c oxidase subunit 1 (COI) sequences were added to the 377 existing records from the BOLD data base, to compile a near complete reference dataset. Standard sequence similarity analyses resulted in 83 distinct clusters almost perfectly reflecting the expected number of species in Austria. Mean intraspecific distances of 0.22% were significantly lower than distances to closest relatives, resulting in a pronounced barcoding gap and unique Barcode Index Numbers (BINs) for most of the species. Four cases of BIN sharing were detected, pointing to hybridization and/or recent divergence, whereas in Phoxinus spp., Gobio spp. and Barbatula barbatula intraspecific splits, multiple BINs and consequently cryptic diversity were observed. The overall high identification success and clear genetic separation of most of the species confirms the applicability and accuracy of genetic methods for bio-surveillance. Furthermore, the new DNA barcoding data pinpoints cases of taxonomic uncertainty, which need to be addressed in further detail, to more precisely assort genetic lineages and their local distribution ranges in a new National Red-List.

Introduction

DNA barcoding was introduced as a suitable method for biological species discrimination in animals in 2003 [1], and since then the method has continued to receive unprecedented attention. For most animal groups, the region near the 5’-end of the cytochrome C oxidase subunit 1 (COI) is established as the standard barcoding marker. Despite certain valid reservations [e.g. 24], an enormous number of studies on various taxonomic groups (e.g., see [5] for plants [6], for insects [7,8], for amphibians and reptiles [9], for fungi, and [10] for fish) have accumulated over the last two decades. One particular upside of DNA barcoding is the breadth of useful applications. When applied to fishes, it can be used to investigate freshwater [10] or marine species [1113], to determine species regardless of their ontogenetic stage [1417] or to identify only residual parts of animals [18]. Furthermore, DNA barcoding data is increasingly used as a means for tracking catch records, food authenticity, mislabeling or fraud [1922]. Moreover, freshwater ecosystems are among the most threatened throughout the world and freshwater species in Europe have experienced an 83% decline in populations over the last 50 years [23,24]. Habitat degradation, water pollution, river channel regulation, hydropower exploitation, invasive species and ultimately climate change entail a range of pressures that threaten freshwater biodiversity worldwide [2426]. Furthermore, the high level of endemism within freshwater ecosystems, coupled with challenges in direct observation, requires tools for sound identification of species and evolutionary significant units to implement conservation efforts [27,28]. Species discrimination is also critical for biological monitoring and conservation purposes, hence DNA barcoding has gained additional importance in the light of recent alerts of biodiversity loss across all terrestrial and aquatic habitats [29,30]. Furthermore, biological surveillance increasingly encourages non-invasive sampling techniques like environmental DNA (eDNA) approaches [31,32], which heavily rely on high-quality genetic reference databases in order to facilitate reliable read identification and species assignment. Tracking biodiversity, however, requires precise species determination and while the identification of most adult (European) fishes can usually be achieved quite easily by experts, some morphologically challenging cases like the whitefishes (Coregonus spp.), minnows (Phoxinus spp.) or alien species like weatherfishes (Misgurnus spp.) [3236] as well as the identification of juvenile fish remain difficult tasks [1417]. In such cases, DNA barcoding might not necessarily replace classical morphology-based approaches as a stand-alone technique, but can aid as a complementary method to increase resolution [16,37,38]. However, in order to yield optimal identification results, DNA barcoding is heavily dependent on high quality, deep coverage reference libraries (e.g. the BOLD database [39]), which profit from the steady augmentation with unambiguously determined reference specimens [10]. Several national barcoding initiatives (such as GBOL, www.bolgermany.de; Barcoding Fauna Bavarica, barcoding-zsm.de/bfb; SWISSBOL, www.swissbol.ch; FINBOL, www.finbol.org; NORBOL, www.norbol.org) contribute their share and ensure continuity and the steady increase in reference data quality [40,41]. The Austrian Barcode of Life initiative (ABOL, www.abol.ac.at) is part of this international network aiming to contribute to this global database and, concomitantly investigate native biodiversity.

Based on the latest Austrian Red List of endangered teleost fish and lamprey species from 2007 [42] as well as other literature on the Austrian fish fauna [44], approximately 85 fish species are present in Austria, 70 of which are considered native. However, these literature sources differ widely concerning some taxa. For example, the genus Coregonus accounts for 12 out of 85 species in [42], but only a single entity in [43], where it was considered to be a “species complex” due to taxonomic uncertainties. As the current Red List was compiled almost 15 years ago (last version from 2007) and new/alien invaders/species/lineages [34,35,4548] have been recently recorded, the current ABOL-project also provides a valuable source of data for an update of the current Red List of Austrian teleost fish and lampreys, and a timely overview of the current freshwater fish diversity of Austria. Comprehensive knowledge on fish diversity is key for designing appropriate conservation action plans and may also support initial assessment of the need for management actions to be taken against invasive species.

Taken together, this study aims to i) add unambiguously determined reference specimens of Austrian fish to the international barcode of life database (BOLD), ii) contribute to the current understanding of the Austrian fish fauna and investigate the extant diversity (loss of species in the wild, new invaders/introductions) and iii) test the discriminating power of DNA barcoding for Austrian fishes.

Material and methods

The cumulative combination of all teleost fish and lamprey species listed in [43,44] as well as the current Red List for Austrian freshwater fishes [42] was used to define the extant freshwater fish diversity in Austria. According to the literature, 70 out of 85 species are listed as native. Additionally, a newly described species of gudgeon [45] and an alien species of weatherfish [48] have been added to the known fish diversity. In order to comprehensively cover the Austrian species assemblage, the present dataset consists of two sources of barcode sequences: i) COI sequences of Austrian fish species already available from BOLD ([32,34,35,4548] including unpublished records (iBOL data release)) and ii) new COI barcode sequences generated in the course of this study. At the time this dataset was compiled, 1,048 COI sequences of Austrian fishes were available on BOLD (22.03.2021). Of those, samples not identified to the species level as well as all samples with sequences less than 500 bp in length were excluded, leaving 377 BOLD sequences. For more in-depth analyses of potentially ambiguous taxa pinpointed by the initial investigation (see below), sequences from other regions of Europe, outside of Austria, were downloaded from BOLD and compiled into separate datasets for Phoxinus spp. ([10,32,34,35,45,4958], DS-EPHO (dx.doi.org/10.5883/DS-EPHO)) and Barbatula barbatula ([10,4951,59,60], DS-EBBAR (dx.doi.org/10.5883/DS-EBBAR)). For the fresh material, all samples were opportunistically obtained in the framework of licensed electrofishing surveys in the years 2014–2021 conducted by a variety of private and public authorities. The rest of the samples were donated by state natural history museums (Natural History Museum Vienna, Oberösterreichisches Landesmuseum Linz). All newly collected specimens are stored permanently at Natural History Museums [see project code ‘BCAFL’ on BOLD (www.boldsystems.org) for sampling and taxonomic information]. Fin clips were taken and stored in pure ethanol at -20°C. Extraction of DNA of all 689 samples from 70 localities (Fig 1) followed a rapid Chelex protocol [61]. PCR, chain termination sequencing and SephadexTM G-50 (Amersham Biosciences) purification of the DNA barcode region (COI) amplicons using the primer combination C_FishF1t1 and C_FishR1t1 [62] and FishF1 and Str_R [63] followed [64,65] with the BioTherm DNA polymerase (GeneCraft Germany) and 50°C annealing temperature being the only alterations. Sequences were visualized on an ABI 3500xl capillary sequencer (Applied Biosystems).

thumbnail
Fig 1. Map of sampling localities.

Map of Austria and surrounding countries indicating the sampling locations of newly sequenced samples (orange spheres).

https://doi.org/10.1371/journal.pone.0268694.g001

All sequences were edited manually using MEGA 6.06 [66] and uploaded to the BOLD database, and are accessible under the project ‘ABOL–Barcoding of the Austrian fish and lampreys (BCAFL)’. The final dataset of both downloaded and newly generated sequences consisted of 1,016 sequences (DS-AFISH dx.doi.org/10.5883/DS-AFISH) for subsequent analyses (see Table 1 for number of sequences per species). Visualization of sequence similarity clustering was conducted using the ‘Taxon ID Tree’ tool implemented on BOLD with the BOLD aligner algorithm. Intra (Imax)- and interspecific genetic distances (distance to nearest neighbor–DNN) were calculated under the K2P model with the ‘Barcode Gap Analysis’ tool also implemented on BOLD (K2P distance model, BOLD aligner, complete deletion for ambiguous base/gap handling). Furthermore, both distance-based, Automatic Barcode Gap Discovery’ (ABGD, [67]) or ‘Assemble Species by Automatic Partitioning’ (ASAP, [68]), and tree-based, the ‘Bayesian Poisson Tree Processes’ model (bPTP, [69]), species delimitation methods were conducted. For ABGD, the alignment containing all sequences was downloaded from BOLD and uploaded to the ABGD webserver (https://bioinfo.mnhn.fr/abi/public/abgd/abgdweb.html). Analyses were run with the Kimura (K2P) TS/TV model with the preset parameters (Pmin: 0.001, Pmax: 0.1, Steps: 10, X (relative gap width): 1.5). The same procedure was conducted for ASAP, also run from a webserver (https://bioinfo.mnhn.fr/abi/public/asap/asapweb.html) with the default parameters. For the bPTP analysis, the phylogenetic input tree was inferred using the IQ-TREE webserver (http://iqtree.cibiv.univie.ac.at/) with the automatic substitution model and 1000 ultrafast bootstrap replicates [70]. The resulting tree was converted to Newick format in FigTree v1.4.4 (http://tree.bio.ed.ac.uk/software/figtree/) and uploaded to the bPTP webserver (https://species.h-its.org/ptp/) where the analysis was run with 100,000 MCMC generations, the thinning set to 100, a burn-in fraction of 0.1 and a random seed [69].

Results

From the 689 samples covering all but one of the extant families (only Anguillidae is missing), 96% of the genera and 95% of all fish species present in Austria (based on [4244]), 639 COI barcodes ranging from 512 to 700 bp in length were generated, representing an overall sequencing success rate of 93%. All sequences are accessible on BOLD (project code ‘BCAFL’) and GenBank (ON097269—ON097906). The overall dataset (1,016 sequences), including downloaded records from Austrian fish samples, covers a total of 94% of all families, 98% of all genera and 96% of all species present in Austria. The sequence similarity clustering resulted in 84 distinct clades largely mirroring morphological species identification and 83 Barcode Index Numbers (BINs, Fig 2).

thumbnail
Fig 2. NJ tree based on DNA barcode sequences of Austrian fish species.

NJ tree of Austrian teleost fish and lamprey species based on K2P distances of 1,016 COI DNA barcode sequences. The topology of the tree was inferred with the “Taxon ID Tree” tool implemented in BOLD and visualized in FigTree v1.4.4 (http://tree.bio.ed.ac.uk/software/figtree/). Black frames mark species that are investigated in a broader geographic context further below.

https://doi.org/10.1371/journal.pone.0268694.g002

One specimen originally identified as Prussian carp (Carassius gibelio) was quite divergent from other alleged C. gibelio samples. A BLAST search in BOLD/GenBank indicated, with 100% sequence similarity, that this divergent haplotype sampled in Schwarzaubach in Styria most likely represents the Ginbuna, Carassius langsdorfii, a species hitherto unknown for Austria. In addition to this new record, discordances between currently accepted species, DNA barcodes and BIN assignment were detected in gudgeons of the genus Gobio (three distinct clades, three individual BINs), minnows of the genus Phoxinus (four distinct clades, 4 individual BINs) and stone loaches (Barbatula barbatula, two distinct clades, two individual BINs). Additionally, two different BINs were detected in the Danube roach (Rutilus virgo, BOLD:AAE3231 and BOLD:ADG8651) including a unique new BIN for Austria. Furthermore, BIN sharing was detected in four cases (Leuciscus leuciscus/L. idus, BOLD:AAD5733; Abramis brama/Blicca bjoerkna, BOLD:AAC8592; Carassius langsdorfii/C. gibelio/C. auratus, BOLD:AAA7176 and Ameiurus nebulosus/A. melas, BOLD:AAA7255). These results were also largely reflected by the analysis of genetic distances (Table 1).

thumbnail
Table 1. K2P distances (in %) of COI sequences within and between Austrian fish species.

https://doi.org/10.1371/journal.pone.0268694.t001

With mean intra- and interspecific distances of 0.22 and 6.49% respectively, the barcode gap (i.e., interspecific distances exceeding intraspecific distances) was well reflected for most of the species (Fig 3). Only Blicca bjoerkna (maximum intraspecific distance (Imax): 4.68 (due to a single morphologically clear B. bjoerkna specimen with introgressed Abramis brama mtDNA), the species/lineages of Phoxinus spp. (Imax: 6.28) and Eudontomyzon mariae (Imax: 5.25) showed higher intraspecific than interspecific distances. Additionally, distances to conspecifics exceeding 1.0% were also detected within Alburnoides bipunctatus, Barbatula barbatula, Chondrostoma nasus, Cottus gobio, Gobio spp., Perca fluviatilis, Romanogobio carpathorossicus, Rutilus virgo, Scardinius erythrophthalmus, Squalius cephalus, Tinca tinca and Thymallus thymallus (Table 1).

thumbnail
Fig 3. Visualization of the ‘Barcode gap’.

Barcode gap plot of the distance to the nearest neighbor (NN) vs. the maximum intraspecific distance of Austrian fish species. Dots above the red line suggest the presence of a barcoding gap. Outliers were detected in Abramis brama, Barbus barbus, Blicca bjoerkna, Carassius auratus, Carassius gibelio, Eudontomyzon marie, Leuciscus idus, Leuciscus leuciscus and Phoxinus spp.

https://doi.org/10.1371/journal.pone.0268694.g003

However, except for Gobio spp. and Barbatula barbatula these cases did not result in additional BINs. Similar results were also obtained from the other species delimitation analyses (see S1 Table). ABGD resulted in 88 species in the initial and 90 species in the recursive partition using a prior maximal distance of P = 0.0129. ASAP on the other hand reported 65–91 partitions/species based on the ten best partitioning schemes regarding the ASAP score. Even though the exact grouping of samples/species varies slightly between the individual priors and partitions, the overall patterns are the same, e.g., Gobio gudgeons are lumped into two groups despite the three lineages found by [47], Phoxinus minnows result in at least three distinct groups and that Ameiurus nebulosus and A. melas result in different groups despite their shared BIN. Finally, the maximum likelihood partitioning of the tree-based bPTP resulted in 88 species. Analysis of available pan-European stone loach data revealed at least five distinct lineages (and BINs) of Barbatula barbatula in Europe (Fig 4). Two of those lineages are solely comprised by samples from Germany or Russia, while the other three lineages contain samples from several countries reflecting a geographical pattern with an eastern (Germany and Austria), Danubian, and Northeastern European clade. Interestingly, the Danubian clade branches off from B. vardarensis native to Greece and North Macedonia, with which it shares a common BIN (BOLD:AAA1243).

thumbnail
Fig 4. NJ tree of European Barbatula species.

Phylogeny of European Barbatula species based on COI barcode sequences available on BOLD and from this study. Species names and BINs are given, countries of origin are indicated by acronyms in parentheses.

https://doi.org/10.1371/journal.pone.0268694.g004

Minnows of the genus Phoxinus, however, revealed a far more complex pattern based on their COI sequences. Besides country- and Balkan-specific MOTUs, four Central European lineages containing samples from several countries including Austria were found (Fig 5, S1 Table). Assignment of species names to these molecular taxonomic units (MOTUs) proved difficult, as each cluster contained specimens of various determinations (e.g. BOLD:ADL2661 contained Phoxinus sp., P. phoxinus and P. marsilii). Nonetheless, our results are wholly congruent with the presence of more than one species of Phoxinus in Austria and consequently also in Europe [35].

thumbnail
Fig 5. NJ tree of European Phoxinus species.

Phylogeny of European Phoxinus species based on COI barcode sequences available on BOLD and from this study. Species names and BINs are given, countries of origin are indicated by acronyms in parentheses.

https://doi.org/10.1371/journal.pone.0268694.g005

Discussion

In this study, we present an almost complete DNA barcode reference inventory for Austrian fishes. From the 639 newly generated COI barcode sequences, only the European eel (Anguilla anguilla) as well as two sturgeon species, namely the Russian sturgeon (Acipenser gueldenstaedtii) and the ship sturgeon (Acipenser nudiventris), which have also been listed for Austria [43], are missing. For the two former species, PCRs (of old museum tissue) were unsuccessful, for the latter species no samples could be obtained. For all species, two or more samples were obtained, except for the racer goby (Babka gymnotrachelus), Balon’s ruffe (Gymnocephalus baloni), the stellate sturgeon (Acipenser stellatus), the blue bream (Ballerus ballerus), the sunbleak (Leucaspius delineatus) and the bighead carp (Hypophthalmichthys nobilis), for which only a single sample was available. Whitefish (Coregonus spp.) were not treated as distinct species in our study as there is no consensus yet on whether the different forms found in the different lakes represent different species or ecotypes and because previous studies have shown that divergence of these species/ecotypes is too recent to be fully resolved by mtDNA data [73,74]. These issues are further complicated by hybridization with closely related introduced species throughout their ranges [73,74]. Similar to previous studies [10], analysis of the DNA barcoding data largely mirrors the known national species inventory. However, we found a few cases of BIN sharing and deep intraspecific divergence, potentially indicating cryptic diversity and/or new species records for Austria, in our new dataset.

Taxa sharing BINs

BIN sharing was detected with two species pairs and one trio of species: i) Leuciscus leuciscus and Leuciscus idus, ii) Ameiurus nebulosus and Ameiurus melas and iii) Carassius auratus, Carassius gibelio and Carassius langsdorfii. For L. leuciscus and L. idus, hybridization and mitochondrial replacement has been reported [75], resulting in a shared common haplotype and consequently the same BIN (BOLD:AAD5733). The black bullhead (Ameiurus melas) and the brown bullhead (A. nebulosus) shared the same BIN (BOLD:AAA7255), even though they are clearly separated in the NJ tree (see Fig 2) and other species delimitation analyses. However, this pattern is not an artefact of the Austrian samples alone, but a general pattern evident on BOLD, as this particular BIN is comprised nearly equally by A. melas and A. nebulosus samples (https://www.boldsystems.org/index.php/Public_BarcodeCluster?clusteruri=BOLD:AAA7255), underscoring the shallow divergence between the two species. The two species can be clearly distinguished by morphological characters [76], but introgressive hybridization has been reported repeatedly [77 and references therein] and could be an additional problem for molecular delimitation. Furthermore, genetic distances (2.75 DNN) among these two taxa, albeit high enough to support two distinct species, are fairly low compared to most species. Thirdly, the Prussian carp (Carassius gibelio) and the goldfish (Carassius auratus) share the same BIN with C. langsdorfii. All three species belong to the C. auratus species complex and have long been considered different sub-species of C. auratus, but molecular genetic analyses indicated their distinctness, despite shallow divergence (e.g., [78,79]), a pattern that we also find in our data (see e.g. NJ tree in Fig 2).

Cases of deep intraspecific divergence

In addition to the few taxa sharing BINs, we found three cases of deep divergence, i.e. in the gudgeons of the genus Gobio, in the stone loach, Barbatula barbatula, and in the minnows of the genus Phoxinus.

Gudgeons of the genus Gobio in Austria comprise three distinct mitochondrial lineages that were also resolved as distinct BINs (BOLD:AAC5607, BOLD:ABY6890 and BOLD:ADH1249), which is in sharp contrast to [42,43] who only list one species, G. gobio, and [44], who suggest the presence of two species, G. gobio and G. obtusirostris, for the Austrian Danube system with a potential hybrid zone in the Upper Danube. A recent detailed study [47] found that the three mitochondrial lineages present in Austria correspond to G. gobio, G. obtusirostris and a third lineage that is closely related to other Gobio species from the Balkans. Patterns of genetic diversity suggest that these originally allopatric lineages/species expanded their distribution recently (probably post-glacially) to come into secondary contact and hybridize in the (Austrian) Danube system, thus forming a large hybrid zone in Austria. Even though there seems to be a cline in the relative frequency of the distinct haplogroups from the upper to the lower parts of Danube system [47,80], the distribution of these lineages/species throughout Austria (and adjacent countries) is currently unresolved, and particularly complicated. Barbatula barbatula poses another ambiguous case, where sequences from the 17 morphologically identified samples can be allocated into two separate clusters in the NJ tree, forming two BINs (BOLD:AAA1239 and BOLD:AAA1243). This result is partly in line with the three clades recovered by [10], who also found high levels of divergence (<7.02% sequence divergence), potentially indicating cryptic species. The two lineages recovered in Austrian samples (4.66% divergence) are part of the eastern as well as the southern (Danubian) lineage [10] (Fig 4). This pattern also becomes evident when looking at the pan-European dataset (Fig 4). In addition to the Central European lineages, two Eastern/Northeastern lineages were recovered. This finding is consistent with previous studies [10,81], which also found pronounced structure based on other markers, but did not include Northern European samples. Furthermore, this pattern is similar to what has been observed in gudgeons of the genus Gobio [47], with separate glacial refugia and post-glacial secondary contact and admixture. Similarly, additional nuclear genetic or genomic data would be required to comprehensively dis-entangle the complex pattern observed in the mitochondrial data.

The most complex pattern was found in the genus Phoxinus (the European minnow species complex). While [44] reported Phoxinus phoxinus and P. lumaireul for Central Europe, [34,35,46] identified four species and three additional lineages of Phoxinus in Austria. These are Phoxinus marsilii and P. lumaireul (represented by three different subclades), P. csikii and P. phoxinus (introduced). Discriminating between Phoxinus species and dis-entangling their respective distribution ranges and geographical origins is impeded by subtle morphological differences as well as small interspecific genetic variation, which cannot be detected by DNA barcoding. Species delimitation is further complicated by a long and irreproducible history of stocking and translocation as well as hybridization [35]; thus, further in-depth morphological and genetic/genomic investigations are needed.

First record of ginbuna, Carassius langsdorfii, for Austria

Two species of Carassius, the Crucian carp (C. carassius) and the Prussian carp (C. gibelio), are native to Europe. Additionally, the goldfish (C. auratus) was introduced in the 17th century as an ornamental fish and has established feral populations throughout Europe (e.g., [44,82,83]), a pattern mirrored by more recent introductions of eastern Asiatic strains of C. gibelio [84,85]. Since 2000, another non-native Carassius species, C. langsdorfii, originally distributed in Japan, has been reported from several European countries [82,86,87], most likely introduced as unintended imports together with koi carps (Cyprinus rubrofuscus) [86]. As this species has hitherto not been reported for Austria, our finding of C. langsdorfii in the Schwarzaubach in Styria is the first evidence for its occurrence in Austria. Frequent hybridization among Carassius species, and between Carassius and other cyprinid species, as well as the presence of both sexually reproducing and gynogenetic populations complicate species identification in this genus. In fact, the only species to be reliably identified based on morphology is C. carassius, whereas genetic data are indispensable for identifying the other species in the genus (e.g. [82]). Indeed, knowledge about the present distribution of C. langsdorfii in Europe is almost exclusively based on mtDNA data [87]. However, a caveat of this strategy is that Carasissus species have a high propensity to hybridize, and thus hybridization and introgression might lead to erroneous species identifications when based on mtDNA alone. Nonetheless, the discovery of a C. langsdorfii haplotype at least confirms the presence of C. langsdorfii mtDNA in Austria. Whether our specimen is indeed C. langsdorfii or a hybrid will have to be confirmed by additional, ideally nuclear genetic/genomic data. Phenotypically, this individual has a lower body (with fewer scale rows) than C. gibelio sensu stricto caught at the same site (see S1 Fig). The specimen also differed from C. gibelio sensu stricto specimens by its lighter ventral and darker dorsal side (compare with [86]), suggesting it might indeed be C. langsdorfii.

Nomenclatural issues

Uncertainties in nomenclature such as in the above-mentioned example of C. langsdorfii, but also taxonomic revisions or even ‘under-studied’ groups constitute an un-negligible issue for online repositories such as BOLD but also museum collections. This becomes apparent when, e.g., looking at gudgeons. Both, [42] and [43] listed Gobio kesslerii as present in Austria, whereas [44] already used Romanogobio kesslerii. According to [45], however, the correct species name should be Romanogobio carpathorossicus, and here we follow this suggestion but note that R. carpathorossicus is listed as a synonym of R. kessleri in Eschmeyer’s catalogue of fishes [71]. A similar situation is found in gudgeons of the genus Gobio, where [42,43] only list G. gobio, whereas [44] report G. gobio and G. obtusirostris from the Danube system with the potential existence of a hybrid zone. The most recent work by [47] however, found three distinct lineages (likely corresponding to G. gobio, G. obtusirostris and a third, Balkans-derived lineage) to which we also adhere in this study and which was confirmed by [80]. The distribution of these lineages throughout Austria (and adjacent countries) is currently unresolved, and further complicated by high morphological variability and hybridization [47].

Systematics and taxonomy change over time simply due to the accumulation of new or more comprehensive data [45,8891]. Therefore, museum collections as well as digital (sequence) repositories need to be periodically updated to reflect currently accepted nomenclature. In museum collections, this translates to an iterative additive labelling of physical objects (the verbatim labels are never changed) as well as an immaculate concurrent (digital) documentation [92]. Regarding BOLD, skilled personal observing and incorporating current changes and novelties in the taxonomic backbone are crucial to uphold user confidence and integrity with regards to content. Despite the undisputable requirement of additional effort and resources, this accuracy and timeliness will ensure maximum reliability and use of reference barcode data (in the sense of voucher-related DNA sequences) as well as museum collections for future applications.

This barcode-based inventory of the Austrian fish fauna has brought some new additions [45,47,48,93] and while some of these novelties are shared with adjacent countries [e.g. [10,33], others are original to Austria [45] underscoring the need to update a national Red List. We argue that national red lists should increasingly be augmented by genetic data [10,9496], which allows for non-invasive monitoring [54] and might illuminate the need for further detailed ecological or systematic study for problematic or ambiguous taxa [31,32]. Here, we provide the first comprehensive DNA barcode reference set for Austrian fishes, which may serve as a basis for a regularly updated Austrian Red List of fish species, aid in sample/specimen identification for both basic and applied monitoring, provide the basis for sound fisheries management and conservation of native fish populations and facilitate read determination in eDNA or meta-barcoding studies. Furthermore, our data update helps to increase the coverage of barcoding data at the European scale and thus will likely be useful in a wider biogeographic context.

Supporting information

S1 Fig. Pictures of Carassius samples from Schwarzaubach, Styria.

A) C. langsdorfii, B-D) C. gibelio.

https://doi.org/10.1371/journal.pone.0268694.s001

(TIF)

S1 Table. Summary of species delimitation analyses results.

https://doi.org/10.1371/journal.pone.0268694.s002

(DOCX)

Acknowledgments

We are grateful to Günter Parthl, Gerhard Woschitz, Albert Rechberger, Josef Melcher, Peter Mehlmauer, Harald Ellinger, Michael Jung, Clemens Ratschan, Klemens Koblmüller, Mario Poglitsch and the Fischereiverein Leibnitz for their help acquiring various samples during electric fishing surveys. We also thank Taraneh E. H. Westergerling, Raphael Schmid and Elisabeth Glatzhofer for their help with the laboratory work. We also kindly acknowledge Sandra Kirchner and Oliver Macek from the Natural History Museum in Vienna for managing the reference specimens and providing Museum IDs and for help with database work, respectively. We would also like to thank Anja Palandačić for her comments and suggestions on the topic of minnows in Austria.

References

  1. 1. Hebert PDN, Cywinska A, Ball SL, DeWaard JR. Biological identifications through DNA barcodes. Proc R Soc B Biol Sci. 2003 Feb 7;270(1512):313–21. pmid:12614582
  2. 2. Will KW, Rubinoff D. Myth of the molecule: DNA barcodes for species cannot replace morphology for identification and classification. Cladistics. 2004 Feb;20(1):47–55. pmid:34892971
  3. 3. Ebach MC, Holdrege C. DNA barcoding is no substitute for taxonomy [2]. Vol. 434, Nature. Nature Publishing Group; 2005. p. 697.
  4. 4. DeSalle R, Goldstein P. Review and Interpretation of Trends in DNA Barcoding. Front Ecol Evol. 2019 Sep 10;7:302.
  5. 5. Lahaye R, Van Der Bank M, Bogarin D, Warner J, Pupulin F, Gigot G, et al. DNA barcoding the floras of biodiversity hotspots. Vol. 105, PNAS. 2008. pmid:18258745
  6. 6. Hebert PDN, Ratnasingham S, Zakharov E V., Telfer AC, Levesque-Beaudin V, Milton MA, et al. Counting animal species with DNA barcodes: Canadian insects. Philos Trans R Soc B Biol Sci. 2016 Sep 5;371(1702). pmid:27481785
  7. 7. Hawlitschek O, Morinière J, Dunz A, Franzen M, Rödder D, Glaw F, et al. Comprehensive DNA barcoding of the herpetofauna of Germany. Mol Ecol Resour. 2016;16(1):242–53. pmid:25892157
  8. 8. Zangl L, Daill D, Schweiger S, Gassner G, Koblmüller S. A reference DNA barcode library for Austrian amphibians and reptiles. PLoS One. 2020;15(3). pmid:32163447
  9. 9. Truong C, Mujic AB, Healy R, Kuhar F, Furci G, Torres D, et al. How to know the fungi: combining field inventories and DNA-barcoding to document fungal diversity. Vol. 214, New Phytologist. Blackwell Publishing Ltd; 2017. p. 913–9.
  10. 10. Knebelsberger T, Dunz AR, Neumann D, Geiger MF. Molecular diversity of Germany’s freshwater fishes and lampreys assessed by DNA barcoding. Mol Ecol Resour. 2015 May 1;15(3):562–72. pmid:25186809
  11. 11. Zemlak TS, Ward RD, Connell AD, Holmes BH, Hebert PDN. DNA barcoding reveals overlooked marine fishes. Mol Ecol Resour. 2009 May;9(SUPPL. 1):237–42. pmid:21564983
  12. 12. Weigt LA, Baldwin CC, Driskell A, Smith DG, Ormos A, Reyier EA. Using DNA barcoding to assess Caribbean reef fish biodiversity: Expanding taxonomic and geographic coverage. PLoS One. 2012 Jul 17;7(7). pmid:22815912
  13. 13. Landi M, Dimech M, Arculeo M, Biondo G, Martins R, Carneiro M, et al. DNA barcoding for species assignment: The case of Mediterranean marine fishes. PLoS One. 2014 Sep 15;9(9). pmid:25222272
  14. 14. Victor BC, Hanner R, Shivji M, Hyde J, Caldow C. Identification of the larval and juvenile stages of the Cubera Snapper, Lutjanus cyanoptems, using DNA barcoding. Zootaxa. 2009;(2215):24–36.
  15. 15. Hubert N, Delrieu-Trottin E, Irisson JO, Meyer C, Planes S. Identifying coral reef fish larvae through DNA barcoding: A test case with the families Acanthuridae and Holocentridae. Mol Phylogenet Evol. 2010 Jun 1;55(3):1195–203. pmid:20188843
  16. 16. Ko HL, Wang YT, Chiu TS, Lee MA, Leu MY, Chang KZ, et al. Evaluating the Accuracy of Morphological Identification of Larval Fishes by Applying DNA Barcoding. PLoS One. 2013 Jan 31;8(1). pmid:23382845
  17. 17. Wibowo A, Wahlberg N, Vasemägi A. DNA barcoding of fish larvae reveals uncharacterised biodiversity in tropical peat swamps of New Guinea, Indonesia. Mar Freshw Res. 2017;68(6):1079–87.
  18. 18. Holmes BH, Steinke D, Ward RD. Identification of shark and ray fins using DNA barcoding. Fish Res. 2009 Jan 14;95(2–3):280–8.
  19. 19. Cline E. Marketplace substitution of Atlantic salmon for Pacific salmon in Washington State detected by DNA barcoding. Food Res Int. 2012 Jan 1;45(1):388–93.
  20. 20. Galal-Khallaf A, Ardura A, Mohammed-Geba K, Borrell YJ, Garcia-Vazquez E. DNA barcoding reveals a high level of mislabeling in Egyptian fish fillets. Food Control. 2014 Dec 1;46:441–5.
  21. 21. Wong EHK, Hanner RH. DNA barcoding detects market substitution in North American seafood. Food Res Int. 2008 Oct 1;41(8):828–37.
  22. 22. Cawthorn DM, Steinman HA, Witthuhn RC. DNA barcoding reveals a high incidence of fish species misrepresentation and substitution on the South African market. Food Res Int. 2012 Apr 1;46(1):30–40.
  23. 23. Kristensen P, Whalley C, Zal FNN, Christiansen T. European waters assessment of status and pressures 2018. EEA Rep. 2018;(No.7/2018).
  24. 24. Albert JS, Destouni G, Duke-Sylvester SM, Magurran AE, Oberdorff T, Reis RE, et al. Scientists’ warning to humanity on the freshwater biodiversity crisis. Ambio. 2021;50(1):85–94. pmid:32040746
  25. 25. Vörösmarty CJ, McIntyre PB, Gessner MO, Dudgeon D, Prusevich A, Green P, et al. Global threats to human water security and river biodiversity. Nat 2010 4677315. 2010 Sep 29;467(7315):555–61. pmid:20882010
  26. 26. Dudgeon D. Multiple threats imperil freshwater biodiversity in the Anthropocene. Curr Biol. 2019 Oct 7;29(19):R960–7. pmid:31593677
  27. 27. Fraser DJ, Bernatchez L. Adaptive evolutionary conservation: towards a unified concept for defining conservation units. Mol Ecol. 2001 Dec;10(12):2741–52. pmid:11903888
  28. 28. Crandall KA, Bininda-Emonds ORR, Mace GM, Wayne RK. Considering evolutionary processes in conservation biology. Trends Ecol Evol. 2000 Jul 1;15(7):290–5. pmid:10856956
  29. 29. Sheth BP, Thaker VS. DNA barcoding and traditional taxonomy: An integrated approach for biodiversity conservation. Genome. 2017;60(7):618–28. pmid:28431212
  30. 30. Trivedi S, Ansari AA, Ghosh SK, Rehman H. DNA barcoding in marine perspectives: Assessment and conservation of biodiversity. DNA Barcoding in Marine Perspectives: Assessment and Conservation of Biodiversity. 2016. 1–330 p.
  31. 31. Thalinger B, Wolf E, Traugott M, Wanzenböck J. Monitoring spawning migrations of potamodromous fish species via eDNA. Sci Rep. 2019 Dec 1;9(1):1–11.
  32. 32. Thalinger B, Oehm J, Mayr H, Obwexer A, Zeisler C, Traugott M. Molecular prey identification in Central European piscivores. Mol Ecol Resour. 2016 Jan 1;16(1):123–37. pmid:26053612
  33. 33. Belle CC, Stoeckle BC, Cerwenka AF, Kuehn R, Mueller M, Pander J, et al. Genetic species identification in weatherfish and first molecular confirmation of Oriental Weatherfish Misgurnus anguillicaudatus (Cantor, 1842) in Central Europe. Knowl Manag Aquat Ecosyst. 2017 Aug;(418):31.
  34. 34. Ramler D, Palandačić A, Delmastro GB, Wanzenböck J, Ahnelt H. Morphological divergence of lake and stream Phoxinus of Northern Italy and the Danube basin based on geometric morphometric analysis. Ecol Evol. 2017 Jan 1;7(2):572–84. pmid:28116054
  35. 35. Palandačić A, Naseka A, Ramler D, Ahnelt H. Contrasting morphology with molecular data: An approach to revision of species complexes based on the example of European Phoxinus (Cyprinidae). BMC Evol Biol. 2017 Aug 9;17(1).
  36. 36. Wanzenböck J, Hopfinger M, Wanzenböck S, Fuxjäger L, Rund H, Lamatsch DK. (2021). First successful hybridization experiment between native European weatherfish (Misgurnus fossilis) and non-native Oriental weatherfish (M. anguillicaudatus) reveals no evidence for postzygotic barriers. NeoBiota. 2021 (69):29–50.
  37. 37. Schindel DE, Miller SE. DNA barcoding a useful tool for taxonomists. Nature. 2005;435(7038):17–17. pmid:15874991
  38. 38. Chan A, Chiang LP, Hapuarachchi HC, Tan CH, Pang SC, Lee R, et al. DNA barcoding: Complementing morphological identification of mosquito species in Singapore. Parasites and Vectors. 2014;7(1).
  39. 39. Ratnasingham S, Hebert PDN. BOLD: The Barcode of Life Data System: Barcoding. Mol Ecol Notes. 2007 May;7(3):355–64. pmid:18784790
  40. 40. Carr CM, Hardy SM, Brown TM, Macdonald TA, Hebert PDN. A tri-oceanic perspective: DNA barcoding reveals geographic structure and cryptic diversity in Canadian polychaetes. PLoS One. 2011;6(7). pmid:21829451
  41. 41. Čandek K, Kuntner M. DNA barcoding gap: reliable species identification over morphological and geographical scales. Mol Ecol Resour. 2015 Mar 1;15(2):268–77. pmid:25042335
  42. 42. Wolfram G, Mikschi E. Rote Liste der Fische (Pisces) Österreichs. Rote List gefährdeter Tiere Österreichs Checklisten, Gefährdungsanalysen, Handlungsbedarf Tl 2 Grüne R des Leb Band 14/2. 2007;61–198.
  43. 43. Fische Hauer W., Krebse & Muscheln in heimischen Seen und Flüssen. Leopold Stocker Verlag; 2007. 231 p.
  44. 44. Kottelat M, Freyhof J. Handbook of European freshwater fishes. Publications Kottelat; 2007.
  45. 45. Friedrich T, Wiesner C, Zangl L, Daill D, Freyhof J, Koblmüller S. Romanogobio skywalkeri, a new gudgeon (Teleostei: Gobionidae) from the upper Mur River, Austria. Zootaxa. 2018;4403(2):336–50. pmid:29690236
  46. 46. Palandačić A, Kruckenhauser L, Ahnelt H, Mikschi E. European minnows through time: museum collections aid genetic assessment of species introductions in freshwater fishes (Cyprinidae: Phoxinus species complex). Heredity (Edinb). 2020;124(3):410–22. pmid:31896822
  47. 47. Zangl L, Daill D, Gessl W, Friedrich T, Koblmüller S. Austrian gudgeons of the genus Gobio (Teleostei: Gobionidae): A mixture of divergent lineages. J Zool Syst Evol Res. 2020 Feb 25;58(1):327–40.
  48. 48. Zangl L, Jung M, Gessl W, Koblmüller S, Ratschan C. Oriental or not: First record of an Alien Weatherfish (misgurnus) species in Austria verified by molecular data. BioInvasions Rec. 2020;9(2):375–83.
  49. 49. Clusa L, García-Vázquez E. A simple, rapid method for detecting seven common invasive fish species in Europe from environmental DNA. Aquat Conserv Mar Freshw Ecosyst. 2018;28(3):619–29.
  50. 50. Imoto JM, Saitoh K, Sasaki T, Yonezawa T, Adachi J, Kartavtsev YP, et al. Phylogeny and biogeography of highly diverged freshwater fish species (Leuciscinae, Cyprinidae, Teleostei) inferred from mitochondrial genome analysis. Gene. 2013;514(2):112–24. pmid:23174367
  51. 51. Schönhuth S, Vukić J, Šanda R, Yang L, Mayden RL. Phylogenetic relationships and classification of the Holarctic family Leuciscidae (Cypriniformes: Cyprinoidei). Mol Phylogenet Evol. 2018;127:781–99. pmid:29913311
  52. 52. Perea S, Böhme M, Zupančič P, Freyhof J, Šanda R, Özulu M, et al. Phylogenetic relationships and biogeographical patterns in Circum-Mediterranean subfamily Leuciscinae (Teleostei, Cyprinidae) inferred from both mitochondrial and nuclear data. BMC Evol Biol. 2010;10(1). pmid:20807419
  53. 53. Behrens-Chapuis S, Herder F, Esmaeili HR, Freyhof J, Hamidan NA, Özuluğ M, et al. Adding nuclear rhodopsin data where mitochondrial COI indicates discrepancies–can this marker help to explain conflicts in cyprinids? DNA Barcodes. 2016;3(1):187–99.
  54. 54. Behrens-Chapuis S, Herder F, Geiger MF. Adding DNA barcoding to stream monitoring protocols–What’s the additional value and congruence between morphological and molecular identification approaches? PLoS One. 2021 Jan 1;16(1 January). pmid:33395693
  55. 55. Corse E, Meglécz E, Archambaud G, Ardisson M, Martin JF, Tougard C, et al. A from-benchtop-to-desktop workflow for validating HTS data and for taxonomic identification in diet metabarcoding studies. Mol Ecol Resour. 2017 Nov 1;17(6):e146–59. pmid:28776936
  56. 56. Geiger MF, Herder F, Monaghan MT, Almada V, Barbieri R, Bariche M, et al. Spatial heterogeneity in the Mediterranean Biodiversity Hotspot affects barcoding accuracy of its freshwater fishes. Mol Ecol Resour. 2014 Nov 1;14(6):1210–21. pmid:24690331
  57. 57. De Santis V, Delmastro GB, Vanetti I, Britton JR, Zaccara S. Species composition of introduced and natural minnow populations of the Phoxinus cryptic complex in the westernmost part of the Po River Basin (north Italy). Biol Invasions. 2021 Mar 1;23(3):657–68.
  58. 58. Denys GPJ, Dettai A, Persat H, Daszkiewicz P, Hautecœur M, Keith P. Revision of Phoxinus in France with the description of two new species (Teleostei, Leuciscidae). Cybium. 2020;44(3):205–37.
  59. 59. Norén M, Kullander S, Nydén T, Johansson P. Multiple origins of stone loach, Barbatula barbatula (Teleostei: Nemacheilidae), in Sweden based on mitochondrial DNA. J Appl Ichthyol. 2018 Feb 1;34(1):58–65.
  60. 60. Gauliard C, Dettaï A, Persat H, Keith P, Denys GP, Denys Barbatula leoparda Actinopterygii GP, et al. Barbatula leoparda (Actinopterygii, Nemacheilidae), a new endemic species of stone loach of French Catalonia. Cybium. 2019;43(2):169–77.
  61. 61. Richlen M, Notes PB-ME, 2005 undefined. A technique for the rapid extraction of microalgal DNA from single live and preserved cells. Mol Ecol Notes. 2005 Sep;5(3):688–91.
  62. 62. Ivanova N V., Zemlak TS, Hanner RH, Hebert PDN. Universal primer cocktails for fish DNA barcoding. Mol Ecol Notes. 2007 Jul 1;7(4):544–8.
  63. 63. Ward RD, Zemlak TS, Innes BH, Last PR, Hebert PDN. DNA barcoding Australia’s fish species. Philos Trans R Soc B Biol Sci. 2005 Oct 29;360(1462):1847–57. pmid:16214743
  64. 64. Koblmãller S, Salzburger W, Obermãller B, Eigner E, Sturmbauer C, Sefc KM. Separated by sand, fused by dropping water: Habitat barriers and fluctuating water levels steer the evolution of rock-dwelling cichlid populations in Lake Tanganyika. Mol Ecol. 2011 Jun 1;20(11):2272–90. pmid:21518059
  65. 65. Duftner N, Koblmüller S, Sturmbauer C. Evolutionary relationships of the Limnochromini, a tribe of benthic deepwater cichlid fish endemic to Lake Tanganyika, East Africa. J Mol Evol. 2005;60(3):277–89. pmid:15871039
  66. 66. Tamura K, Stecher G, Peterson D, Filipski A, Kumar S. MEGA6: Molecular Evolutionary Genetics Analysis Version 6.0. Mol Biol Evol. 2013 Dec 1;30(12):2725–9. pmid:24132122
  67. 67. Puillandre N, Lambert A, Brouillet S, Achaz G. ABGD, Automatic Barcode Gap Discovery for primary species delimitation. Mol Ecol. 2012 Apr;21(8):1864–77. pmid:21883587
  68. 68. Puillandre N, Brouillet S, Achaz G. ASAP: assemble species by automatic partitioning. Mol Ecol Resour. 2021 Feb 1;21(2):609–20. pmid:33058550
  69. 69. Zhang J, Kapli P, Pavlidis P, Stamatakis A. A general species delimitation method with applications to phylogenetic placements. Bioinformatics. 2013 Nov 15;29(22):2869–76. pmid:23990417
  70. 70. Trifinopoulos J, Nguyen L-T, von Haeseler A, Minh BQ. W-IQ-TREE: a fast online phylogenetic tool for maximum likelihood analysis. Nucleic Acids Res. 2016 Jul 8;44(W1):W232–5. pmid:27084950
  71. 71. Fricke, R., Eschmeyer WN& R van der L. Eschmeyer’s Catalog of Fishes: Genera, Species, References.(http://researcharchive.calacademy.org/research/ichthyology/catalog/fishcatmain.asp). Electronic version accessed 21.12.2021 [Internet]. Fricke, R., Eschmeyer, W. N. & R. van der Laan. 2021 [cited 2021 Dec 21]. Available from: https://www.calacademy.org/scientists/projects/catalog-of-fishes.
  72. 72. Lerceteau-Köhler E, Schliewen U, Kopun T, Weiss S. Genetic variation in brown trout Salmo trutta across the Danube, Rhine, and Elbe headwaters: a failure of the phylogeographic paradigm?. BMC Evol Biol. 2013;13(1):1–18.
  73. 73. Pamminger-Lahnsteiner B, Weiss S, Winkler KA, Wanzenböck J. Composition of native and introduced mtDNA lineages in Coregonus sp. in two Austrian lakes: Evidence for spatio-temporal segregation of larvae? Hydrobiologia. 2009;632(1):167–75.
  74. 74. Winkler KA, Pamminger-Lahnsteiner B, Wanzenböck J, Weiss S. Hybridization and restricted gene flow between native and introduced stocks of Alpine whitefish (Coregonus sp.) across multiple environments. Mol Ecol. 2011 Feb 1;20(3):456–72. pmid:21199024
  75. 75. Hänfling B, Brandl R. Phylogenetics of European cyprinids: insights from allozymes. J Fish Biol. 2000 Aug 1;57(2):265–76.
  76. 76. Nowak M, Koščo J, Popek W, Epler P. First record of the black bullhead Ameiurus melas (Teleostei: Ictaluridae) in Poland. J Fish Biol. 2010 Apr 1;76(6):1529–32. pmid:20537032
  77. 77. Walter RP, Gnyra ES, Söderberg LI, Heath DD. Rapid genetic identification of brown bullhead (Ameiurus nebulosus), black bullhead (Ameiurus melas) and their hybrids. Conserv Genet Resour. 2014 Feb 26;6(3):507–9.
  78. 78. Cheng L, Chang YM, Lu CY, Cao DC, Sun XW. [DNA barcoding and species and subspecies classification within genus Carassius]. Kunming Inst Zool Chinese Acad Sci. 2012 Nov 16;33(5):463–72. pmid:23019027
  79. 79. Podlesnykh A V., Apalikova O V., Brykov VA. Phylogenetic relationships of silver crucian carp in Carassius auratus complex based on mtDNA analysis. Russ J Genet. 2012 Dec;48(12):1207–17. pmid:23516900
  80. 80. Takács P, Ferincz Á, Imecs I, Kovács B, Nagy AA, Ihász K, et al. Increased spatial resolution of sampling in the Carpathian basin helps to understand the phylogeny of central European stream-dwelling gudgeons. BMC Zool. 2021 Dec 1;6(1).
  81. 81. Šedivá A, Janko K, ŠlechtovÁ V, Kotlík P, Simonović P, Delic A, et al. Around or across the Carpathians: Colonization model of the Danube basin inferred from genetic diversification of stone loach (Barbatula barbatula) populations. Mol Ecol. 2008 Mar 1;17(5):1277–92. pmid:18302688
  82. 82. Rylková K, Kalous L, Bohlen J, Lamatsch DK, Petrtýl M. Phylogeny and biogeographic history of the cyprinid fish genus Carassius (Teleostei: Cyprinidae) with focus on natural and anthropogenic arrivals in Europe. Aquaculture. 2013 Mar 4;380–383:13–20.
  83. 83. Ribeiro F, Collares-Pereira MJ, Moyle PB. Non-native fish in the fresh waters of Portugal, Azores and Madeira Islands: A growing threat to aquatic biodiversity. Fish Manag Ecol. 2009 Aug 1;16(4):255–64.
  84. 84. Gorgan LD, Ciorpac M. Identification of Carassius gibelio migration patterns in Europe. AACL Bioflux. 2013;6(2):92–6.
  85. 85. Ağdamar S, Tarkan AS. High genetic diversity in an invasive freshwater fish species, Carassius gibelio, suggests establishment success at the frontier between native and invasive ranges. Zool Anz. 2019 Nov 1;283:192–200.
  86. 86. Kalous L, ŠLechtová V, Bohlen J, Petrtýl M, ŠVátora M. First European record of Carassius langsdorfii from the Elbe basin. J Fish Biol. 2007 Mar;70(SUPPL. A):132–8.
  87. 87. Kalous L, Rylková K, Bohlen J, Šanda R, Petrtýl M. New mtDNA data reveal a wide distribution of the Japanese ginbuna Carassius langsdorfii in Europe. J Fish Biol. 2013 Feb 1;82(2):703–7. pmid:23398078
  88. 88. Cai T, Cibois A, Alström P, Moyle RG, Kennedy JD, Shao S, et al. Near-complete phylogeny and taxonomic revision of the world’s babblers (Aves: Passeriformes). Mol Phylogenet Evol. 2019 Jan 1;130:346–56. pmid:30321696
  89. 89. Tortoli E, Meehan CJ, Grottola A, Fregni Serpini G, Fabio A, Trovato A, et al. Genome-based taxonomic revision detects a number of synonymous taxa in the genus Mycobacterium. Infect Genet Evol. 2019 Nov 1;75:103983. pmid:31352146
  90. 90. Thomson SA, Pyle RL, Ahyong ST, Alonso-Zarazaga M, Ammirati J, Araya JF, et al. Taxonomy based on science is necessary for global conservation. PLOS Biol. 2018 Mar 14;16(3):e2005075. pmid:29538381
  91. 91. Thiele KR, Conix S, Pyle RL, Barik SK, Christidis L, Costello MJ, et al. Towards a global list of accepted species I. Why taxonomists sometimes disagree, and why this matters. Org Divers Evol. 2021;
  92. 92. Mulcahy DG, Ibáñez R, Jaramillo CA, Crawford AJ, Ray JM, Gotte SW, et al. DNA barcoding of the National Museum of Natural History reptile tissue holdings raises concerns about the use of natural history collections and the responsibilities of scientists in the molecular age. Plos one. 2022; 17(3), e0264930. pmid:35245325
  93. 93. Rabitsch W, Milasowszky N, Nehring S, Wiesner C, Wolter C, Essl F. The times are changing: temporal shifts in patterns of fish invasions in central European fresh waters. J Fish Biol. 2013 Jan;82(1):17–33. pmid:23331135
  94. 94. Ward RD, Holmes BH, White WT, Last PR. DNA barcoding Australasian chondrichthyans: Results and potential uses in conservation. Mar Freshw Res. 2008;59(1):57–71.
  95. 95. Swartz ER, Mwale M, Hanner R. A role for barcoding in the study of African fish diversity and conservation. Vol. 104, South African Journal of Science. 2008. p. 293–8.
  96. 96. Lakra WS, Verma MS, Goswami M, Lal KK, Mohindra V, Punia P, et al. DNA barcoding Indian marine fishes. Mol Ecol Resour. 2011 Jan;11(1):60–71. pmid:21429101