You are on page 1of 11

Molecular Phylogenetics and Evolution 155 (2021) 106982

Contents lists available at ScienceDirect

Molecular Phylogenetics and Evolution


journal homepage: www.elsevier.com/locate/ympev

Repeated hybridization increased diversity in the door snail complex


Charpentieria itala in the Southern Alps
Jie Xu, Bernhard Hausdorf *
Center of Natural History, Zoological Museum, University of Hamburg, Martin-Luther-King-Platz 3, 20146 Hamburg, Germany

A R T I C L E I N F O A B S T R A C T

Keywords: The door snail species complex Charpentieria itala is widely distributed in the Southern Alps and subdivided into
ddRAD several morphologically differentiated subspecies. Thus, it can be used as a model group for understanding
Charpentieria migration and differentiation processes in the Southern Alps. We generated genome-wide double digest Re­
Southern Alps
striction Site Associated DNA (ddRAD) sequencing data for 166 specimens from 36 populations of the door snail
Phylogeography
Hybridization
Charpentieria itala and for 8 specimens of the other three Charpentieria species to reconstruct their evolutionary
Land snails history and phylogeography. Phylogenetic and STRUCTURE analyses based on the ddRAD data indicated that the
repeated separation of the populations in western and eastern groups by the Garda glacier during the glacials was
the process that most strongly shaped the population structure of C. itala. This process may also explain a similar
phylogeographic boundary in many other southern Alpine animal and plant species. Our study revealed that
some populations that resemble Charpentieria stenzii morphologically and ecologically, the ‘stenzioid’ subspecies,
originated by a hybridization event with Charpentieria stenzii. A further hybridization event between stenzioid
populations that survived the glacials in mountain refuges and non-stenzioid populations that probably came into
contact with stenzioid populations as a result of climate warming during an interglacial resulted in the origin of a
hybrid subspecies that is adapted to intermediate altitudes. Our study demonstrated that the origin of new
differentiated taxa by hybridization, is more frequent than previously assumed.

1. Introduction and Linares, 2008; Nieto Feliner et al., 2017; Schumer et al., 2014, 2018)
or the extinction of one or both parental species (Gómez et al., 2015;
Climate changes often cause shifts or expansions of distribution areas Muhlfeld et al., 2014; Rhymer and Simberloff, 1996; Seehausen et al.,
of many species that may result in contact between closely related 2008). The frequency of the different outcomes (Buerkle et al., 2003) has
species that were previously geographically isolated (Hewitt, 2000; still to be assessed across different taxa, different regions and different
Schmitt, 2007; Taberlet et al., 1998). This is true for the repeated cycles environmental histories to be able to better predict the probable con­
of climate changes during the Pleistocene, but also for the current sequences of the current human-induced climate changes.
human-induced climate changes. The encounter of closely related spe­ The few markers that could be screened in non-model organisms
cies may result in hybridization (Chunco, 2014; Hewitt, 2000; Schmitt, until recently did hardly allow the reconstruction of complex hybridi­
2007; Taberlet et al., 1998). In mountain areas geographical range zation processes. The accessibility of genomic data for non-model or­
changes may further be complicated by altitudinal range shifts. During ganisms by next generation sequencing (Andrews et al., 2016; Franchini
the Pleistocene climactic cycles, population groups became frequently et al., 2017; Peterson et al., 2012) and a plethora of new bioinformatics
isolated, begun to diverge and came into contact again in a later climatic tools for analysing admixture (Patterson et al., 2012; Pickrell and
cycle. Hybridization of incompletely isolated groups can result in com­ Pritchard, 2012) have opened new possibilities to detect and reconstruct
plex population structures and morphological patterns. The evolu­ hybridization and introgression.
tionary outcomes of hybridization range from the introgression of genes The door snail genus Charpentieria includes four species which are
for adaptive traits (Abbott et al., 2013; Arnold and Kunte, 2017; Becker mainly distributed in the Southern Alps and adjacent mountain regions
et al., 2013; Hedrick, 2013; Minder and Widmer, 2008) to hybrid (Nordsieck, 1963; Welter-Schultes, 2012) and can be used as a model
speciation (Abbott et al., 2013; Koch et al., 2016; Mallet, 2007; Mavárez group for understanding migration and differentiation processes in the

* Corresponding author.
E-mail address: hausdorf@zoologie.uni-hamburg.de (B. Hausdorf).

https://doi.org/10.1016/j.ympev.2020.106982
Received 18 June 2020; Received in revised form 6 August 2020; Accepted 2 October 2020
Available online 13 October 2020
1055-7903/© 2020 Elsevier Inc. All rights reserved.
J. Xu and B. Hausdorf Molecular Phylogenetics and Evolution 155 (2021) 106982

Southern Alps. Charpentieria itala is the most widespread species that 2012).
extends from the Lago Maggiore to the Dolomites (Supplementary In the current study, we used double-digest restriction site associated
Fig. S1) and which is represented by one subspecies in the Alpes- DNA (ddRAD) sequencing (Franchini et al., 2017; Peterson et al., 2012)
Maritimes, the Ligurian Alps and the Northern and Central Apennines. to generate a genomic dataset for all Charpentieria species and pop­
Charpentieria dyodon is restricted to the western Southern Alps from ulations from the complete range of the C. itala complex in the Southern
Hautes-Alpes in France to Ticino. Charpentieria stenzii is spread from the Alps to clarify the evolutionary history and phylogeography of the
Dolomites to the Karawanks and Julian Alps in Slovenia. Charpentieria C. itala complex in the Southern Alps and its relations to other Char­
ornata occurs from the Southeastern Alps in Austria southwards to pentieria species. We used new approaches to detect introgression events
Croatia and in an isolate in the Sudetes. to better understand the roles of parallel adaptation versus adaptive
Beside the four undisputed Charpentieria species there is a complex of introgression in the evolution of this species complex.
taxa in the Bergamasque, Brescia and Garda Prealps that combine fea­
tures of C. itala and of C. stenzii. Whereas typical C. itala live on humid 2. Material and methods
rocks covered with vegetation, usually in forests or shaded by bushes,
and can also be found directly on trees, the stenzii-like or “stenzioid” 2.1. Sampling
subspecies, clavata, variscoi, balsamoi, triumplinae, trepida, tiesenhauseni,
and lorinae (Supplementary Fig. S1), usually dwell on vertical surfaces of Charpentieria populations in the Southern Alps between Lake Lugano
more exposed calcareous rocks as C. stenzii does (Nordsieck, 1963). and the Dolomites were sampled during expeditions in 2009, 2011 and
Charpentieria stenzii and the stenzioid subspecies are characterized by a 2017. All samples were stored in 100% isopropanol at − 20 ◦ C. The
more or less protruded aperture with a continuous peristome. This is populations were classified according to Nordsieck (1963, 2007) and
presumably an adaptation to life on vertical surfaces which enables Nardi (2011). For a comparison of the classifications of Nordsieck
these snails to attach the aperture more tightly to the rock surface while (2007) and the classification proposed in this study see Supplementary
resting or to carry the shell more easily in a vertical position (Scheel and Table S1. We chose 163 specimens from 35 populations of the C. itala
Hausdorf, 2012). Because of the similarities with C. stenzii, Käufel complex from the Southern Alps (Fig. 1). Additionally, three specimens
(1928) classified the stenzioid taxa as subspecies of C. stenzii. However, from an introduced population in Weinheim in Germany, the type lo­
Nordsieck (1963) found shared characteristics in the genitalia between cality of C. itala braunii, were investigated. As outgroups, specimens of
the stenzioid subspecies and C. itala and intermediate populations be­ the three other Charpentieria species, C. ornata, C. stenzii and C. dyodon,
tween them. Thus, he suggested that the stenzioid taxa are not related to and of representatives of three further genera of the Delimini, Delima
C. stenzii, but classified them as subspecies of C. itala. Furthermore, laevissima, Papillifera papillaris and Siciliaria grohmanniana, were
Nordsieck (1963) supposed that the stenzioid taxa originated poly­ included. The morphological classification and collection data of the
phyletically from typical C. itala ancestors in adaptation to the open specimens used in this study are compiled in Supplementary Table S1.
limestone rock faces.
The stenzioid subspecies, and the non-stenzioid C. itala latestriata and
C. itala albopustulata are distributed in a mosaic pattern in the Berga­ 2.2. DNA extraction, ddRAD library preparation and sequencing
masque, Brescia and Garda Prealps (Nordsieck, 1963; Supplementary
Fig. S1). The stenzioid subspecies occur at higher altitudes almost Total genomic DNA was extracted from tissue samples of the foot
exclusively in mountain ranges that were not glaciated during the gla­ following the protocol proposed by Sokolov (2000) with slight modifi­
cials (Scheel and Hausdorf, 2012). In the western Bergamasque Prealps cations as detailed in Scheel and Hausdorf (2012). DNA concentrations
they are replaced by the non-stenzioid subspecies C. itala latestriata at were measured using the Qubit v2.0 fluorometer (Life Technologies,
intermediate altitudes and C. itala albopustulata at the lowest altitudinal Carlsbad, CA, USA) with the dsDNA HS assay kit (Thermo Fisher Sci­
levels. Populations attributed to the latter subspecies also populate the entific, Waltham, MA, USA).
valleys of the Brescia and Garda Prealps, where they are connected with For library preparation, we used the modified double-digest restric­
stenzioid subspecies by transitional populations. Nevertheless, Nord­ tion-site-associated DNA (quaddRAD) sequencing protocol proposed by
sieck (2002, 2007) separated the stenzioid subspecies from C. itala as a Franchini et al. (2017) with slight modifications. In brief, 200 ng
distinct species, Charpentieria clavata, because of sympatric occurrences genomic DNA were digested with the restriction enzymes FastDigest PstI
of stenzioid and non-stenzioid populations without intermediates in the and MspI (Thermo Fisher Scientific) and ligated to adaptors with inner
western Bergamasque Prealps (Nordsieck, 1963). barcodes and 4 random bases for the recognition of PCR duplicates in the
The distribution and the differences in morphology and ecology same reaction. The products were purified and double size-selected with
between stenzioid and non-stenzioid subspecies of Charpentieria itala can 0.45x and 0.35x AMPure XP (Beckman Coulter, Brea, USA) beads. Then
be explained by different hypotheses. The stenzioid subspecies may have equal amounts by weight of twelve samples each with different inner
a common origin and may represent relicts of an early colonization wave barcode combinations were pooled. In the following step, each pool of
that survived the ice ages in isolated mountain refuges within the Alps, the digested/ligated DNA is amplified by PCR with Phusion High-
whereas the non-stenzioid subspecies may have been displaced from the Fidelity Polymerase (Thermo Fisher Scientific, Waltham, MA, USA)
Alps during the glacials and re-colonized the south-alpine valleys only and primers with Illumina indices. After a purification step with 0.8x
postglacially (Käufel, 1928). In contrast, Nordsieck (1963) suggested AMPure XP, the DNA concentration and fragment length distribution of
that the geographically isolated stenzioid subspecies evolved through each pool was determined on a TapeStation (Agilent Technologies,
parallel adaptation of C. itala populations to life on vertical surfaces of Santa Clara, USA) and the amount of DNA of each pool in the range
exposed calcareous rocks. Later, Nordsieck (1984) himself rejected the 400–500 bp was calculated. Aliquots of 16 pools with equimolar
hypothesis that the stenzioid subspecies originated by convergent evo­ amounts of fragments in this size range were pooled into one library.
lution and suggested that their similarity originated from an introgres­ Finally, fragments in the range 400–500 bp were selected using a Blue
sion with an extinct species. Pippin electrophoresis system (Sage Science, Beverly, USA). Paired-end
A previous study of the relationships of southern Alpine C. itala next-generation sequencing (2 × 150 bp) was performed at Macrogen
populations based on mitochondrial DNA sequences and nuclear AFLP Inc. (Seoul, Korea) in one Illumina HiSeq X lane. Raw RAD-seq data of
markers confirmed the classification of the stenzioid subspecies in the this study have been deposited in the European Nucleotide Archive
C. itala complex and revealed monophyletic groups of western and (ENA) at EMBL-EBI under accession number PRJEB41328, using the
eastern stenzioid subspecies, but could not resolve their relationships to data brokerage service of the German Federation for Biological Data
each other and to other C. itala population groups (Scheel and Hausdorf, (GFBio, Diepenbroek et al., 2014).

2
J. Xu and B. Hausdorf Molecular Phylogenetics and Evolution 155 (2021) 106982

Fig. 1. Sampled Charpentieria itala populations in the Southern Alps and number of private SNPs (restricted to one sampled population)/ mean number of individuals
genotyped per locus (pSNPs/mN). Red ellipse: approximate ancestral area based on populations with high numbers of private SNPs. Encircled numbers refer to
localities listed in Supplementary Table S1, where also the classification of the populations is specified. Background map (Geologische Bundesanstalt, 2013) shows
maximum extension of the glaciers at the last glacial maximum. pSNPs/mN was not calculated for population 18, because only one specimen was sampled. (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

2.3. quaddRAD locus assembly applied to call genotypes with the gstacks program. Loci were selected
that were present in more than 80% of all individuals using the pop­
The raw fastq files obtained from the Illumina run were demulti­ ulations module (option -R). SNPs with a minimum minor allele fre­
plexed by the sequencing provider using the outer dual Illumina indices. quency of 0.05 (option –min-maf) and a maximum observed
We further processed the data using the STACKS software pipeline, version heterozygosity of 0.5 (option –max-obs-het) were called. For the popu­
2.4 (Catchen et al., 2013). First, we identified and removed PCR dupli­ lation genetic analyses and the Neighbor-net, we used only one random
cates based on the random four-base stretch incorporated into the SNP per locus (option –write-random-snp) to minimize linkage. This
adaptors using the clone_filter module (option –inline_inline). If these SNP data set can be found in Supplementary Table S2. Summary genetic
four bases at the beginning of each read and the rest of the sequence (six- diversity statistics for populations of C. itala were calculated with STACKS
base inner barcode and DNA template) of these reads were identical, a based on a dataset including one random biallelic SNP per locus with a
single copy, stripped off the four random bases at the 5′ end of each minimum minor allele frequency of 0.01 to record also private alleles, i.
paired reads, was retained. The retained reads were then demultiplexed e. alleles restricted to a population.
by the dual inner barcodes using the process_radtags module (option
–inline_inline). Two mismatches were allowed in the barcodes (option
–barcode_dist_1 –barcode_dist_2) and mutated barcodes were rescued 2.4. Population genetic structure
(option -r). Low-quality reads (options -c -q), reads shorter than 120 bp
(options -len_limit) and reads with an average quality score below 15 Individual-based clustering and admixture analyses of the SNP data
(option -s) within a sliding window were discarded. were performed with STRUCTURE 2.3.4 (Falush et al., 2007; Pritchard et al.,
The demultiplexed reads were assembled using the de novo assembly 2000). To minimize the effects of linkage, we only used one random SNP
pipeline of STACKS. We followed the recommendations of Paris et al. per locus. The input files for the STRUCTURE analyses were generated by
(2017) to determine the most suitable assembly parameters for our STACKS populations module (option –structure). A total of 20 runs with

dataset. As recommended, we required that there were at least 3 iden­ 180,000 iterations after a burn-in period of 20,000 iterations for K = 1 to
tical reads in an individual sample to create a stack (option -m in K = 10 were carried out with the admixture model allowing for corre­
ustacks). Following the r80 rule of Paris et al. (2017), we varied the lated allele frequencies between populations. We used the mean prob­
number of mismatches allowed between stacks within individuals (op­ abilities L(K) of the data for each K and the ad hoc quantity ΔK proposed
tion -M in ustacks) and selected the value that maximized the number of by Evanno et al. (2005) computed with STRUCTURE HARVESTER (Earl and
polymorphic loci found in 80% of the samples. As further recommended vonHoldt, 2012) to estimate the number of clusters. In addition, we
by Paris et al. (2017), we varied the number of mismatches allowed carried out a longer run for the optimal value of K with 800,000 gen­
between stacks between individuals (option -n in cstacks) as n = M ± 1. erations after a burn-in period of 200,000 generations. To assess the
Again, we selected the value that maximized the number of polymorphic substructure within the primary groups, subsequent analyses were
loci found in 80% of the samples. The default Marukilow model was conducted within each of the primary clusters, using the same param­
eters as described above. The STRUCTURE results were visualized using

3
J. Xu and B. Hausdorf Molecular Phylogenetics and Evolution 155 (2021) 106982

CLUMPAK (Kopelman et al., 2015). by the number of samples present at that locus < 10). In addition, one
C. ornata individual was removed because of a high number of missing
2.5. Phylogenetic and network analyses loci (twice as much as in the other two C. ornata individuals).
We determined the most suitable parameters for the de novo assem­
For RAXML, BEAST and TREEMIX analyses, we first generated all sequences bly of loci with STACKS for two datasets. For phylogenetic analyses with
for each sample by the populations module of STACKS (option –fasta- RAXML, ASTRAL-III and BEAST, we generated a dataset including 3 individuals
samples). Then, we computed consensus sequences of the two alleles of of Delimini other than Charpentieria, 3 individuals of C. stenzii, 2 of
each individual of each locus using a custom script and concatenated the C. ornata and 3 of C. dyodon as well as 14 individuals of the C. itala
alignments of the consensus sequences with the program FASCONCAT (Kück complex representing different subspecies. We optimized the STACKS pa­
and Meusemann, 2010). The alignment files and phylogenetic trees are rameters for this dataset (m = 3, M = 19, n = 19; see Supplementary
available from TreeBASE (study accession number 26911). Fig. S2) and compiled 843 loci (in total 214,730 bp) present in 80% of
Maximum likelihood (ML) analyses based on the concatenated se­ the individuals. For all other analyses except TREEMIX, we optimized the
quences of the ddRAD loci were performed with RAXML version 8.2.12 STACKS parameters with a dataset including all C. itala samples (m = 3, M
(Stamatakis, 2014) using the GTR-GAMMA model. Branch support = 7, n = 8; see Supplementary Fig. S2). With these parameters and minor
values were computed by bootstrapping with a stopping criterion that allele frequency of 0.05 3379 loci (in total 884,090 bp) present in 80% of
automatically determines if enough bootstrap replicate searches were the individuals were found. An overview about the amount of sequence
conducted for obtaining stable support values (Pattengale et al., 2010). data for the different datasets can be found in Supplementary
We used ASTRAL-III, version 5.6.3 (Zhang et al., 2018) to estimate the Table S3–4. For the calculation of summary genetic diversity statistics
species tree from unrooted gene trees under the multi-species coalescent for the populations of the C. itala complex, we set the minimum minor
model. Gene trees were generated by RAXML with the GTR-GAMMA allele frequency to 0.01 to record also private SNPs (Fig. 1, Supple­
model. mentary Table S5). The resulting dataset included one random biallelic
We dated the divergences of the major lineages within Charpentieria SNP each from 3457 loci. For the TREEMIX analyses, a dataset including
using the Bayesian algorithm implemented in BEAST version 2.4.1 one random biallelic SNP each from 2457 loci present in 80% of the
(Bouckaert et al., 2014) with the GTR + G model. A linked uncorrelated individuals of all Charpentieria individuals was generated with the same
relaxed lognormal molecular clock was used for the BEAST analysis (a STACKS parameters.
strict clock was rejected at a 1% significance level using the test
implemented in MEGA X; Kumar et al., 2018). As tree prior the birth-death 3.2. Population genetic structure
model was chosen and the analysis was run for 10,000,000 generations
with a sampling frequency set to 1000. Tracer v1.6 was used to assess The number of SNPs restricted to a population divided by the number
convergence of runs and 10% of generations were discarded as burn-in. of scored individuals is highest in some populations from the Brescia,
A maximum clade credibility tree with median node heights was Garda and western Venetian Prealps (Fig. 1; Supplementary Table S5). It
calculated with TreeAnnotator 2.1.3. According to the dated phylogeny decreases westwards in the Bergamasque Prealps and eastwards in the
of the Clausiliidae shown by Uit de Weerd and Gittenberger (2013), eastern Venetian Prealps. With a few exceptions, e.g. in the Berga­
Charpentieria separated from the other Delimini (in our dataset repre­ masque Prealps and south of Meran, it is low in areas that were glaciated
sented by Delima laevissima, Papillifera papillaris and Siciliaria groh­ during the last glacial maximum (Fig. 1).
manniana) approximately 15 Ma ago. Thus, we calibrated the tree using The mean probability L(K) of the data increased from K = 1 to K = 6
a lognormal-distributed prior assuming an age of 15 Ma for the sepa­ (Supplementary Fig. S3A) and the ad hoc quantity ΔK based on the
ration of Charpentieria from the other Delimini to get a rough time frame output of the STRUCTURE analyses of the SNP data had its highest value for
for the evolution of Charpentieria. K = 2 (Supplementary Fig. S3B). A longer run with 800,000 generations
A Neighbor-net (Bryant and Moulton, 2004) was constructed based after a burn-in period of 200,000 generations with K = 2 essentially
on shared allele distances (Bowcock et al., 1994) between Charpentieria agreed with the shorter runs and indicated a division of the C. itala
itala individuals using SPLITSTREE4 version 4.14.3 (Huson and Bryant, complex into western and eastern populations with a border at Lake
2006). We calculated the shared allele distances based on the SNP data Garda (Fig. 2A). Some populations close to Lake Garda showed admix­
using the function ’alleledist’ of PRABCLUS (Hennig and Hausdorf, 2020), ture up to 50%.
an add-on package for the statistical software R (R Core Team, 2020). To assess the substructure within the two primary groups, subse­
We used TREEMIX version 1.1 (Pickrell and Pritchard, 2012) to infer quent analyses were conducted within both primary clusters. The mean
the history of population splits and gene flow within Charpentieria, probabilities L(K) of the data increased from K = 1 to K = 4 (Supple­
allowing up to five migration events. The input file was generated by mentary Fig. S3C, E) and the ΔK had its highest value for K = 2 for both
STACKS populations program (–treemix). This method first constructs a subgroups (Supplementary Fig. S3D, F). The western populations were
bifurcating maximum-likelihood tree of populations. It then identifies subdivided into stenzioid and non-stenzioid populations (Fig. 2B). The
populations that are poor fits to the tree model, and models migration populations of the western stenzioid subspecies (C. clavata clavata,
events involving these populations. C. clavata variscoi and C. clavata balsamoi) showed no admixture with
non-stenzioid populations, whereas the populations of eastern stenzioid
3. Results subspecies (C. itala triumplinae, C. itala trepida, C. itala tiessenhauseni, and
C. itala lorinae) are composed of genomic shares of stenzioid as well as
3.1. Sequencing and SNP calling non-stenzioid origin. Some stenzioid shares were also found in the
C. itala latestriata populations from the Bergamasque Prealps and in the
We obtained a total of 886 million paired-end sequences with read C. itala allatollae population from the Garda Prealps. The eastern pop­
length of 151 bp. The raw data consisted of 16 pools (including 12 in­ ulation group was subdivided into the easternmost populations classi­
dividuals each) with unique barcode combination. 5.4–10.2% of the fied as C. itala serravalensis versus the other eastern populations with
reads of each pool were identified as PCR duplicates and removed. The some admixture (Fig. 2B).
output number of paired-end reads for each individual varied from
792,660 to 11,829,670. Samples with adapter barcode i501 had fewer 3.3. Evolutionary history
output reads then others indicating a quality problem with that adaptor.
After a test run with the de novo assembly pipeline of STACKS, we removed The maximum-likelihood tree calculated with RAXML based on the
8 samples due to low read coverage per locus (mean coverage, weighted concatenated alignments of 843 loci including all outgroup samples and

4
J. Xu and B. Hausdorf Molecular Phylogenetics and Evolution 155 (2021) 106982

Austria

22 Switzerland
Weinheim
Germany

24 25 28 30 36

1 4 6 9 10 13 14 16 17 19 23 26 31 34 35

2 3 5 7 8 11 12 15
18 20 21 27 29 32 33

A 50 km

Austria

22 Switzerland
Weinheim
Germany

24 25 28 30 36

1 4l 6s 9 10 13s 14 16s 17 19 23 26 31 34 35

2 3s 5s 7l 8s 11 12s 15s
18 20 21 27 29 32 33

B 50 km

Fig. 2. A. STRUCTURE plots based on 3379 random SNPs (one per locus) of 166 individuals from 36 populations of the Charpentieria itala complex showing their
estimated ancestries in two assumed genetic clusters. B. STRUCTURE plot showing the estimated ancestries in two subclusters within the western group and two
subclusters within the eastern group. s = stenzioid subspecies; l = Charpentieria itala latestriata. Numbers refer to localities listed in Supplementary Table S1, where
also the classification of the populations is specified.

a subset of the C. itala individuals showed that the C. itala complex and clavata, C. clavata variscoi and C. clavata balsamoi form a clade. Char-
C. dyodon as well as C. stenzii and C. ornata are sister species (Fig. 3). pentieria c. clavata is sister to C. clavata variscoi plus C. clavata balsamoi.
Within the C. itala complex, the deepest split separated the populations The population of C. clavata variscoi from Val Taléggio (population 5; see
from west of Lake Garda from those from east of Lake Garda. The Figs. 1 and 2) are more closely related to C. clavata balsamoi from Ambria
stenzioid subspecies form a clade that is nested in populations classified (population 8) than to C. clavata variscoi from Valtorta (population 6). In
as C. itala albopustulata from west of Lake Garda. Also the populations the eastern clade, C. itala baldensis from higher altitudes of Mount Baldo
classified as C. itala latestriata form a clade that is nested in this group. (population 20) is sister to all other populations. Charpentieria i. itala,
Among the stenzioid subspecies, the western subspecies, C. clavata C. itala braunii and C. itala rubiginea are not separated in the tree. The

5
J. Xu and B. Hausdorf Molecular Phylogenetics and Evolution 155 (2021) 106982

Fig. 3. Trees based on sequences of 843 loci of 14 specimens of the Charpentieria itala complex and C. ornata, C. stenzii, C. dyodon, Delima laevissima, Papillifera
papillaris and Siciliaria grohmanniana. Left side: maximum likelihood tree based on concatenated sequences calculated with RAXML with bootstrap support values. Right
side: species tree based on gene trees calculated with ASTRAL-III with posterior probabilities. Encircled numbers refer to populations and colours refer to groups within
the Charpentieria itala complex (see Fig. 1 and Supplementary Table S1).

easternmost populations classified as C. itala serravalensis (populations of the Late Miocene (C. itala – C. dyodon 10.4 Ma ago, 95% HPD:
31–36) form a clade that is nested in the other eastern populations. The 7.8–13.4 Ma; C. stenzii – C. ornata 10.3 Ma ago, 95% HPD: 7.6–13.1 Ma).
relationships within the C. itala complex were largely confirmed in a The western and the eastern lineage of the C. itala complex diverged
second maximum-likelihood tree based on the concatenated alignments probably at the beginning of the Pliocene (5.3 Ma ago; 95% HPD:
of 3379 loci of all individuals of the C. itala complex, rooted according to 3.8–6.9 Ma). The origin of the stenzioid subspecies (c. 3.9 Ma ago; 95%
the maximum-likelihood tree with the outgroups (Supplementary HPD: 2.8–5.0 Ma) also predated the Pleistocene.
Fig. S4). Also the population tree calculated with TREEMIX (Fig. 5) based on one
The coalescent-based species tree estimated with ASTRAL-III based on random biallelic SNP each from 2457 loci with C. stenzii and C. ornata as
842 gene trees including all outgroup samples and a subset of the C. itala outgroups confirmed the relationships of the major groups within the
individuals (Fig. 3) agreed largely with maximum likelihood tree. C. itala complex inferred with the other approaches. Charpentieria itala
However, C. itala latestriata was sister to the stenzioid subspecies in the latestriata is nested in the non-stenzioid C. itala albopustulata pop­
maximum likelihood tree based on the concatenated sequences, whereas ulations. However, the TREEMIX analysis showed in addition gene flow
it was sister to non-stenzioid C. itala albopustulata populations in the between C. stenzii and the clade including the stenzioid subspecies. It
coalescent-based species tree. indicated also several mixture events within the C. itala complex. The
The Neighbor-net based on shared allele distances calculated with migration edge with the highest weight goes from the stenzioid
3379 SNPs of the individuals of the C. itala complex confirmed the C. clavata variscoi plus C. clavata balsamoi clade to the C. itala late-striata
subdivision of the complex into a western and an eastern group of group indicating that C. itala latestriata originated by hybridization be­
populations separated by Lake Garda and the distinctness of the clusters tween this stenzioid lineage and C. itala albopustulata. Furthermore, the
including the stenzioid populations and the C. itala latestriata pop­ Lugano population (population 1) received admixture from an older
ulations, respectively, within the western group (Supplementary lineage, and there was gene flow from a population from the northern
Fig. S5). Val di Non (population 24) to a population in the Eisack valley (popu­
The BEAST analysis based on the concatenated alignment (Fig. 4) lation 30) and from an older lineage to a C. itala serravalensis population
indicated that the divergence between the C. itala complex and in the Valsugana (population 33).
C. dyodon on the one hand and C. stenzii and C. ornata on the other
started at the end of the Middle Miocene (11.3 Ma before present, 95%
highest posterior density interval (HPD): 8.4–14.2 Ma) and that the
divergence of each species pair occurred only little later at the beginning

6
J. Xu and B. Hausdorf Molecular Phylogenetics and Evolution 155 (2021) 106982

Fig. 4. Dated maximum clade credibility tree of Charpentieria based on concatenated sequences of 843 ddRAD loci. Values at the nodes represent median ages of
nodes in million years (Ma). The arrow indicates the node used for calibration. Stars at the branches indicate posterior probabilities * > 0.95 and ** >0.99. Numbers
refer to individuals, encircled numbers refer to populations and colours refer to groups within the Charpentieria itala complex (see Fig. 1 and Supplemen­
tary Table S1).

4. Discussion that extended furthest southwards in the central part of the Alps during
the glacials (Fig. 1; Geologische Bundesanstalt, 2013; Penck and
4.1. Origin and differentiation of Charpentieria itala along the Garda line Brückner, 1909; Seguinot et al., 2018; Voges, 1995), was essential for
their continued divergence. During the long glacials western and eastern
A high number of private SNPs per individual, an indicator for the population groups could differentiate by genetic drift and gene flow
longstanding of populations, in many of the populations from the between adjacent populations resulted in a homogenization within the
Brescia, Garda and western Venetian Prealps (Fig. 1, red ellipse; Sup­ population groups separated by the Garda glacier. Only in the shorter
plementary Table S5) indicated that these regions were part of the interglacial periods, the retreat of the Garda glacier enabled gene flow
ancestral area of the C. itala complex. This is also corroborated by the between western and eastern metapopulations.
position of populations from these regions close to the root within Similar phylogeographic breaks into western and eastern pop­
C. itala in all phylogenetic analyses (Figs. 3–5; Supplementary Fig. S4). ulations in the Southern Alps along the Garda-Etsch-Brenner line have
All phylogenetic analyses (Figs. 3–5; Supplementary Fig. S4) as well been found in several other animals and plants (Cornetti et al., 2015;
as the STRUCTURE analyses (Fig. 2A) indicated a primary division of the Haubrich and Schmitt, 2007; Marchesini et al., 2017; Schönswetter
Alpine populations of the C. itala complex into a western and an eastern et al., 2002; Thiel-Egenter et al., 2009, 2011; Vernesi et al., 2016). This
group at Lake Garda. According to the BEAST analysis (Fig. 4), the line has already been recognized as a border between west and east
divergence between these groups predated the Pleistocene (the highest Alpine plant species by Kerner (1871) and has sometimes been called the
posterior density interval of the estimated divergence time does not “Brenner zone” or “Brenner line” (Cornetti et al., 2015; Marchesini et al.,
overlap with the beginning of the Pleistocene). However, the admixture 2017; Thiel-Egenter et al., 2011; Vernesi et al., 2016). However, because
between the eastern and the western populations (Fig. 2A) and the of the essential role of the Garda glacier for separating western and
intermingling of eastern and western lineages in the mitochondrial gene eastern populations, it is more appropriate to call the boundary the
tree (Scheel and Hausdorf, 2012: Fig. 3) demonstrated that gene flow “Garda line” rather than the “Brenner line”. Some of the species that
across the contact zone occurred. Thus, the repeated geographical sep­ show the pattern, like C. itala, do not occur up to the Brenner pass, which
aration of these populations by the Garda glacier, which was the glacier itself is not a barrier.

7
J. Xu and B. Hausdorf Molecular Phylogenetics and Evolution 155 (2021) 106982

stenzioid subspecies were first explained by convergent adaptation to


the life on exposed rocks by Nordsieck (1963). However, later Nordsieck
(1984) supposed that the character discordances were caused by hy­
bridization with an extinct species. We could resolve the cause of the
observed discordances with our genomic data and new tools for the
detection of introgression like TREEMIX (Pickrell and Pritchard, 2012).
The maximum-likelihood tree as well as the coalescent-based species
tree (Fig. 3) showed that C. dyodon is the sister species of the C. itala
complex as has been supposed by Nordsieck (2002) based on the simi­
larity of their genitalia, whereas C. stenzii is sister to C. ornata.
The stenzioid subspecies formed a monophyletic group within the
C. itala complex in all phylogenetic analyses (Figs. 3–5; Supplementary
Fig. S4). Thus, we can reject the hypothesis that the stenzioid subspecies
evolved through parallel adaptation of C. itala populations to life on
vertical surfaces of exposed calcareous rocks as supposed by Nordsieck
(1963). Instead, the TREEMIX analysis of the ddRAD data indicated
introgression from C. stenzii to the stenzioid subspecies (Fig. 5). This
introgression of C. stenzii genes is most likely the cause for the
morphological and ecological similarities between the stenzioid sub­
species and C. stenzii. Currently, there are no populations of C. stenzii in
the region occupied by the stenzioid populations. The geographically
next population of C. stenzii can be found approximately 50 km towards
the northeast of the easternmost stenzioid populations. The origin of the
stenzioid subspecies probably predated the Pleistocene (Fig. 4). We
suggest that propagules of C. stenzii that spread further west before the
Pleistocene were lost by the hybridization with C. itala populations that
resulted in the origin of the stenzioid populations. Nordsieck (1984) was
the first to suggest that introgression was involved in the origin of the
stenzioid subspecies. However, our TREEMIX analysis (Fig. 5) showed that
his assumption that an unknown, extinct species was involved is not
necessary but that introgression from C. stenzii can explain the similar­
ities of the stenzioid populations with C. stenzii best.
Fig. 5. Tree of 35 Charpentieria populations with five mixture events inferred The ability of the stenzioid subspecies to live at exposed rocks with a
with TREEMIX. Migration arrows are coloured according to their weight and sparse cover of lichens, similar as C. stenzii, enabled the stenzioid sub­
indicate the gene flow direction. Horizontal branch lengths are proportional to
species to colonize higher altitudes than the non-stenzioid populations
the amount of genetic drift that has occurred on the branch. The scale bar shows
and to survive the glacials in unglaciated mountain areas, “massifs de
ten times the average standard error of the entries in the sample covariance
matrix. Encircled numbers refer to populations and colours refer to groups refuge” (Holdhaus, 1954; Stehlik, 2000; Tribsch, 2004; Tribsch and
within the Charpentieria itala complex (see Fig. 1 and Supplementary Table S1). Schönswetter, 2003). They are still geographically restricted to regions
that remained unglaciated during the glacials (Scheel and Hausdorf,
2012). Their occurrences indicate the locations of mountain refuges. In
Thiel-Egenter et al. (2011) suggested that this break zone repre­
contrast, most non-stenzioid populations probably survived the glacials
sented a dispersal barrier for eastern and western genetic lineages of
at lower altitudes at the margin of the Alps as suggested by Käufel
alpine plants during post-glacial recolonization. They assumed that the
(1928).
low and open valleys and high elevations above the permanent snowline
The hybridization between C. itala and C. stenzii that resulted in the
in this break zone can be considered as barriers because these areas lack
origin of the stenzioid subspecies was not the only hybridization event
or only marginally harbour suitable habitats for alpine plants. For
that increased the genetic and morphological diversity within the C. itala
C. itala, however, the valleys do not represent barriers. Charpentieria
complex. The stenzioid subspecies had a common origin, but their
itala can be found on the valley bottom as well as on adjacent mountain
evolutionary trajectories were different. The STRUCTURE analyses for the
slopes. Actually, the species spread through the Etsch and Eisack valley
western group of C. itala populations (Fig. 2B) showed that the western
upwards to Franzensfeste (Nordsieck, 1963) after the retreat of the
subspecies with stenzioid morphology, C. clavata clavata, C. clavata
glaciers in the Holocene (and probably also during the interglacials). We
variscoi and C. clavata balsamoi, are genetically homogeneous, whereas
suggest that also in most other animal and plant groups the repeated
the eastern subspecies with stenzioid morphology have a high propor­
geographical separation of western and eastern populations during the
tion of ancestry in non-stenzioid populations. This is in accordance with
longer glacials by the Garda glacier was more important for their dif­
the fact that there are few morphologically intermediate specimens in
ferentiation than the lack of suitable habitats in the contact zone during
the west, whereas the eastern stenzioid subspecies are connected with
the shorter interglacials.
non-stenzioid populations by a range of morphologically transitional
populations (e.g. the populations named C. itala allatollae; see Nord­
4.2. Increase of diversity by repeated hybridizations sieck, 1963). In the west, stenzioid and non-stenzioid populations even
co-occur at a few sites on the same rocks without mixing (Nordsieck,
Although hybridization between species was known for a long time, 1963). One hypothesis to explain this is that the western stenzioid
it has often been supposed that it usually does not result in introgression subspecies had more time to differentiate from C. itala than the eastern
(Mayr, 1942, 1963, 1992). However, there is increasing evidence for ones. The eastern stenzioid subspecies are distributed in the west of the
introgression and its evolutionary importance (Abbott et al., 2013; approximate ancestral area of C. itala in the Brescia, Garda and western
Arnold and Kunte, 2017; Hedrick, 2013; Mallet, 2007; Mavárez and Venetian Prealps (see above; Fig. 1). Thus, they probably came earlier
Linares, 2008). Also in Charpentieria the discordant distribution of and more often in contact with non-stenzioid populations. These con­
characters of the genitalia and the shell in C. itala, C. stenzii and the tacts resulted in admixture and counteracted the build-up of isolation

8
J. Xu and B. Hausdorf Molecular Phylogenetics and Evolution 155 (2021) 106982

mechanisms. In contrast, the western stenzioid subspecies are far from classification of the stenzioid taxa as subspecies of C. itala. None of the
the ancestral area of C. itala and could differentiate for a longer time two classifications reflects both, the sympatric co-occurrences of sten­
without contact to non-stenzioid populations. The continuous admixture zioid and non-stenzioid populations without fusing in the Bergamasque
between neighbouring populations of stenzioid and non-stenzioid pop­ Prealps and the continuous intergradation of stenzioid and non-
ulations in the east was not shown in the TREEMIX analysis (Fig. 5) as an stenzioid populations in the Brescia and Garda Prealps.
admixture event, but may be the reason for the placement of the eastern Our study demonstrated that the stenzioid taxa had a common origin
stenzioid populations at the base of the stenzioid clade, close to the (Figs. 3–5; Supplementary Fig. S4), but the STRUCTURE analyses (Fig. 2B)
neighbouring non-stenzioid populations. also showed that the trajectories of the stenzioid populations in the
TREEMIX revealed an additional hybridization event between the Bergamasque Prealps and the Brescia and Garda Prealps were different.
stenzioid C. clavata variscoi/balsamoi lineage and neighbouring non- The stenzioid taxa from the Bergamasque Prealps themselves show no
stenzioid populations (Fig. 5). Probably during an early interglacial admixture with neighbouring non-stenzioid populations (Fig. 2B),
the C. clavata variscoi/balsamoi populations came into contact with non- although they locally co-occur with them. Accordingly, we suggest
stenzioid populations that colonized the valleys after the retreat of the classifying them as a separate species C. clavata including the subspecies
glaciers and hybridized with them. This was the origin of C. itala lates­ C. c. clavata, C. clavata variscoi and C. clavata balsamoi (for a comparison
triata that was classified as a non-stenzioid subspecies so far. The of the classification proposed by Nordsieck (2007) and the classification
contribution of the stenzioid populations to the genome of C. itala late- proposed here see Supplementary Table S1). C. clavata variscoi proved to
striata is also evident from the STRUCTURE analyses (Fig. 2B). C. itala be paraphyletic with regard to C. clavata balsamoi. Nevertheless, the
latestriata occurs at intermediate altitudes. It became probably isolated separation of these subspecies may be maintained because they are
from the stenzioid subspecies by the advancing glaciers in the next morphologically well differentiated and are distributed in disjunct
glacial. How it was isolated from other non-stenzioid populations is less mountain regions separated by the Val Brembana so that they will
clear. One possibility is that no other non-stenzioid populations probably continue to diverge.
remained in this region and that C. itala albopustulata, which is found at In accordance with the presence of morphologically intermediate
lower altitudes in that region now, re-colonized this region only later. populations between stenzioid and non-stenzioid populations in the
Our analyses revealed two instances in the C. itala complex in which Brescia and Garda Prealps, the STRUCTURE analyses (Fig. 2B) have shown
hybridization and introgression contributed to the evolution of new that the eastern stenzioid populations share a large part of the genepool
differentiated taxa. The finding of an introgression from C. stenzii into with non-stenzioid C. itala. In contrast, there is no evidence for recent
C. itala and the change of morphological and ecological characteristics gene flow between the stenzioid populations in the Bergamasque Prealps
of exactly the clade that received the introgression towards a phenotype and in the Brescia and Garda Prealps in accordance with the large
similar to C. stenzii, makes it very likely that these changes are attrib­ geographical gap between their ranges. Consequently, we suggest clas­
utable to introgressive alleles of C. stenzii. Hybridization does not only sifying the eastern stenzioid populations as a subspecies of C. itala (see
act as an additional source of adaptive genetic variation within pop­ Supplementary Table S1). The stenzioid populations from the Brescia
ulations that may contribute more to adaptation than mutation (Abbott Prealps are genetically not separated from those of the Garda Prealps,
et al., 2013; Arnold and Kunte, 2017; Grant and Grant, 1994; Hedrick, although they were separated by the Chiese glacier during the glacials.
2013) but the increase in variation contributes essentially to differen­ This is in agreement with the fact that the populations from the Brescia
tiation of population groups and the evolution of new taxa. However, the Prealps do not share morphological characteristics that distinguish them
hybridization events themselves probably did not led to the emergence from those of the Garda Prealps (Nordsieck, 1963). On both sides of the
of reproductive isolation, a criterion required in some definitions of Chiese valley ribbed populations (C. i. tiesenhauseni in the Garda Prealps
hybrid speciation (Schumer et al., 2014, 2018; but see Nieto Feliner and C. i. trepida in the Brescia Prealps) as well as populations with more
et al., 2017). The above described contrast in the isolation of western or less obsolete ribbing (C. i. lorinae in the Garda Prealps and C. i. tri­
and eastern stenzioid populations from adjacent non-stenzioid pop­ umplinae in the Brescia Prealps) occur. Thus, we follow Nordsieck (1963)
ulations either indicates that reproductive isolation between stenzioid in including all stenzioid populations from the Brescia and Garda Pre­
and non-stenzioid populations evolved only in the western stenzioid alps in C. i. lorinae. Later, Nordsieck (2007) and Nardi (2011) recognized
populations after the hybridization or that it was (partly) associated the mentioned named local forms and even some populations that are
with the hybridization, but broke down in the eastern stenzioid pop­ intermediate between stenzioid and non-stenzioid populations (C. i.
ulations. The extent of gene flow between the hybrid subspecies C. itala allatollae) as distinct subspecies. This splitting obscures the close re­
latestriata and the neighbouring C. itala albopustulata is unclear. Despite lationships and minor genetic differentiation of the stenzioid pop­
there is no evidence that the hybridization directly led to the emergence ulations from the Brescia and Garda Prealps.
of reproductive isolation, hybridization contributed to an increase of The morphologically distinct non-stenzioid populations from the
biodiversity in both cases. Depending on the perspective, this may be Bergamasque Prealps that originated by a hybridization event with
considered more relevant for an evaluation of the importance of hy­ C. clavata (see above) were classified as C. itala latestriata by Nordsieck
bridization than the question whether a hybridization event directly (1963, 2007). Given the unique origin and morphologically distinctness
resulted in reproductive isolation, i.e. hybrid speciation in the strict of these populations, classifying them as a distinct subspecies might be
sense. acceptable. However, their relations to neighbouring non-stenzioid
populations should be investigated in more detail.
4.3. Delimitation of species and subspecies The other non-stenzioid populations belonging to the western group
of C. itala largely corresponds to C. itala albopustulata as delimited by
Not surprisingly, the delimitation of species and subspecies in a Nordsieck (1963). However, Nordsieck (1963) included also some
group that has undergone repeated cycles of isolation, secondary contact populations from east of Lake Garda as well as populations considered
and hybridization poses problems. Nordsieck (1963) suggested classi­ transitional forms between C. itala albopustulata and C. itala rubiginea by
fying the stenzioid populations as subspecies despite he was aware that him, but later classified as separate subspecies C. itala baldensis and
some of them co-occur locally with non-stenzioid populations without C. itala braunii by Nordsieck (2007), in C. itala albopustulata. These
mixing. Later, Nordsieck (2002, 2007) changed his mind and separated populations were placed in the eastern group in the phylogenetic and
the stenzioid subspecies from C. itala as a distinct species, C. clavata. admixture analyses (Figs. 2–5, Supplementary Fig. S4). Later, Nordsieck
However, this necessitates an arbitrary subdivision of the continuous (2007) classified these taxa as separate subspecies. However, the ranges
series of intermediate populations in the Brescia and Garda Prealps. of C. itala braunii and C. itala rubiginea as given by Nordsieck (1963)
Thus, Scheel and Hausdorf (2012) preferred to maintain the overlap (Supplementary Fig. S1) and the alleged characteristics like

9
J. Xu and B. Hausdorf Molecular Phylogenetics and Evolution 155 (2021) 106982

shell size and shape, ribbing of the upper whorls and continuous peri­ References
stome vary continuously and show no clear geographical pattern. The
molecular phylogenies also showed that the populations that were Abbott, R., Albach, D., Ansell, S., Arntzen, J.W., Baird, S.J.E., Bierne, N., et al., 2013.
Hybridization and speciation. J. Evol. Biol. 26, 229–246.
classified as C. itala braunii and C. itala rubiginea do not form coherent Andrews, K.R., Good, J.M., Miller, M.R., Luikart, G., Hohenlohe, P.A., 2016. Harnessing
units (Figs. 3–5; Supplementary Fig. S4). Rather, these populations form the power of RADseq for ecological and evolutionary genomics. Nat. Rev. Genet. 17,
a morphologically diverse paraphyletic group from which the popula­ 81–92.
Arnold, M.L., Kunte, K., 2017. Adaptive genetic exchange: a tangled history of admixture
tion groups classified as C. itala itala and C. itala serravalensis by Nord­ and evolutionary innovation. Trends Ecol. Evol. 32 (8), 601–611.
sieck (1963) are derived. Monte Baldo was isolated by the Garda und the Becker, M., Gruenheit, N., Steel, M., Voelckel, C., Deusch, O., Heenan, P.B., Lockhart, P.
Etsch glacier during the glacials, but remained free of ice. The high J., 2013. Hybridization may facilitate in situ survival of endemic species through
periods of climate change. Nat. Clim. Change 3, 1039–1043.
number of private alleles found in the population on Monte Baldo Bouckaert, R., Heled, J., Kühnert, D., Vaughan, T., Wu, C.-H., Xie, D., et al., 2014. BEAST
(population no. 20; Fig. 1) indicated that the C. itala population survived 2: a software platform for Bayesian evolutionary analysis. PLoS Comput. Biol. 10,
the glacials in this mountain refuge. The Monte Baldo population rep­ e1003537.
Bowcock, A.M., Ruiz-Linares, A., Tomfohrde, J., Minch, E., Kidd, J.R., Cavalli-Sforza, L.
resents the sister group to the remaining populations of the eastern
L., 1994. High resolution of human evolutionary trees with polymorphic
group of C. itala. Nordsieck (1963) considered it as a transitional form microsatellites. Nature 368, 455–457.
between C. itala albopustulata and C. itala rubiginea. Actually, it shows Bryant, D., Moulton, V., 2004. Neighbor-net: an agglomerative method for the
admixture with C. itala albopustulata (Fig. 2A). Although similar small construction of phylogenetic networks. Mol. Biol. Evol. 21, 255–265.
Buerkle, C.A., Wolf, D.E., Rieseberg, L.H., 2003. The origin and extinction of species
morphotypes can also be found in other regions, Nordsieck (2007) through hybridization. In: Brigham, C.A., Schwartz, M.W. (Eds.), Population
classified C. itala baldensis as a subspecies. The populations from the Viability in Plants. Springer, Berlin, Heidelberg, Germany, pp. 117–141.
foothills of the Alps (Monte Berici and Colli Euganei) represent the Catchen, J., Hohenlohe, P.A., Bassham, S., Amores, A., Cresko, W.A., 2013. Stacks: an
analysis tool set for population genomics. Mol. Ecol. 22, 3124–3140.
nominotypical C. itala itala. They are somewhat variable. Generally, Chunco, A.J., 2014. Hybridization in a warmer world. Ecol. Evol. 4, 2019–2031.
they do not differ from C. itala albopustulata except in the on average Cornetti, L., Ficetola, G.F., Hoban, S., Vernesi, C., 2015. Genetic and ecological data
larger shell, but some of the populations resemble C. itala serravalensis. reveal species boundaries between viviparous and oviparous lizard lineages.
Heredity 115, 517–526.
The populations from the Valsugana eastwards were classified as Earl, D.A., vonHoldt, B.M., 2012. Structure Harvester: a website and program for
C. itala serravalensis by Nordsieck (1963), but differ morphologically visualizing Structure output and implementing the Evanno method. Conserv. Genetic
hardly from some of the neighbouring populations classified as C. itala Resour. 4, 359–361.
Evanno, G., Regnaut, S., Goudet, J., 2005. Detecting the number of clusters of individuals
itala and C. itala rubiginea. These populations form a monophyletic group using the software Structure: a simulation study. Mol. Ecol. 14, 2611–2620.
in the phylogenetic trees (Fig. 5; Supplementary Fig. S4) and are Falush, D., Stephens, M., Pritchard, J.K., 2007. Inference of population structure using
recognized as a distinct cluster in the STRUCTURE analyses (Fig. 2B) of the multilocus genotype data: dominant markers and null alleles. Mol. Ecol. Notes 7,
574–578.
eastern C. itala subgroup with some admixture with the neighbouring
Franchini, P., Monné Parera, D., Kautt, A.F., Meyer, A., 2017. quaddRAD: a new high-
cluster in the lower Valsugana and elsewhere. multiplexing and PCR duplicate removal ddRAD protocol produces novel
We propose to subsume all C. itala populations east of the Garda zone evolutionary insights in a nonradiating cichlid lineage. Mol. Ecol. 26, 2783–2795.
into C. itala itala (see Supplementary Table S1). The alternative would be Geologische Bundesanstalt (Ed.), 2013. Der Alpenraum zum Höhepunkt der letzten
Eiszeit – Posterkarte. Wien, Austria: Geologische Bundesanstalt. Retrieved from https
to divide these populations into a morphologically and genetically ://opac.geologie.ac.at/wwwopacx/wwwopac.ashx?command=getcontent&server=i
highly variable western group (C. itala itala) and a morphologically and mages&value=Poster_Alpenraum%20Eiszeit_opt.pdf.
genetically uniform eastern group (C. itala serravalensis). The low Gómez, J.M., González-Megías, A., Lorite, J., Abdelaziz, M., Perfectti, F., 2015. The silent
extinction: climate change and the potential hybridization-mediated extinction of
numbers of private alleles in the eastern group (populations no. 31–36; endemic high-mountain plants. Biodivers. Conserv. 24, 1843–1857.
Fig. 1; Supplementary Table S5) showed that these populations do not Grant, P.R., Grant, B.R., 1994. Phenotypic and genetic effects of hybridization in
represent a group that evolved separately for a long time, but that their Darwin’s finches. Evolution 48, 297–316.
Haubrich, K., Schmitt, T., 2007. Cryptic differentiation in alpine-endemic, high-altitude
homogeneity is the result of a recent colonization of the eastern region. butterflies reveals down-slope glacial refugia. Mol. Ecol. 16, 3643–3658.
Moreover, they are connected with the western group by morphologi­ Hedrick, P.W., 2013. Adaptive introgression in animals: examples and comparison to
cally intermediate populations. Thus, we prefer not to separate these new mutation and standing variation as sources of adaptive variation. Mol. Ecol. 22,
4606–4618.
populations as a separate subspecies from the other C. i. itala Hennig, C., Hausdorf, B. 2020. Package ‘prabclus’: functions for clustering of presence-
populations. absence, abundance and multilocus genetic data. R package version 2.3-2. Available
from: http://CRAN.R-project.org/package=prabclus.
Hewitt, G.M., 2000. The genetic legacy of the Quaternary ice ages. Nature 405, 907–913.
Author contribution
Holdhaus, K., 1954. Die Spuren der Eiszeit in der Tierwelt Europas. Abhandlungen der
Zoologisch-botanischen Gesellschaft in Wien 18 (1–493), 1–52.
B.H. conceptualized the study; B.H. and J.X. collected the samples; J. Huson, D.H., Bryant, D., 2006. Application of phylogenetic networks in evolutionary
X. performed the laboratory analyses; J.X. and B.H. analysed the data; B. studies. Mol. Biol. Evol. 23, 254–267.
Käufel, F., 1928. Beitrag zur Kenntnis der Verbreitung und Formenbildung der Clausilien
H. and J.X. wrote the manuscript. in den Südalpen. Archiv für Molluskenkunde 92, 69–107, pl. 2.
Kerner, A., 1871. Die natürlichen Floren im Gelände der deutschen Alpen. In: Schaubach,
Acknowledgements A. (ed.), Die deutschen Alpen für Einheimische und Fremde geschildert. 1. Theil.
Allgemeine Schilderung, second ed. Frommann, Jena, Germany, pp. 126–189.
Koch, E.L., Neiber, M.T., Walther, F., Hausdorf, B., 2016. Presumable incipient hybrid
We are grateful to K.-L. Nägele, H. Nordsieck and F. Walther for speciation of door snails in previously glaciated altitudes in the Caucasus. Mol.
samples, P. Franchini for information about the quaddRAD protocol, D. Phylogenet. Evol. 97, 120–128.
Kopelman, N.M., Mayzel, J., Jakobsson, M., Rosenberg, N.A., Mayrose, I., 2015. CLUMPAK:
Granse for help with flow cytometry and S. Bamberger for comments on a program for identifying clustering modes and packaging population structure
the manuscript. This work was funded by the Land­ inferences across K. Mol. Ecol. Resour. 15 (5), 1179–1191.
esforschungsförderung Hamburg and benefitted from the sharing of Kück, P., Meusemann, K., 2010. FASconCAT: convenient handling of data matrices. Mol.
Phylogenet. Evol. 56, 1115–1118.
expertise within the DFG priority program SPP 1991 Taxon-Omics and Kumar, S., Stecher, G., Li, M., Knyaz, C., Tamura, K., 2018. MEGA X: Molecular
support from DFG HA 2763/6-1. Evolutionary Genetics Analysis across computing platforms. Mol. Biol. Evol. 35,
1547–1549.
Mallet, J., 2007. Hybrid speciation. Nature 446, 279–283.
Appendix A. Supplementary data
Marchesini, A., Ficetola, G.F., Cornetti, L., Battisti, A., Vernesi, C., 2017. Fine-scale
phylogeography of Rana temporaria (Anura: Ranidae) in a putative secondary contact
Supplementary data to this article can be found online at https://doi. zone in the southern Alps. Biol. J. Linn. Soc. 122 (4), 824–837.
org/10.1016/j.ympev.2020.106982. Mavárez, J., Linares, M., 2008. Homoploid hybrid speciation in animals. Mol. Ecol. 17,
4181–4185.
Mayr, E., 1942. Systematics and the Origin of Species. Columbia Univ. Press, New York.
Mayr, E., 1963. Animal Species and Evolution. Belknap Press, Cambridge, MA.

10
J. Xu and B. Hausdorf Molecular Phylogenetics and Evolution 155 (2021) 106982

Mayr, E., 1992. A local flora and the biological species concept. Am. J. Bot. 79, 222–238. Schmitt, T., 2007. Molecular biogeography of Europe: Pleistocene cycles and postglacial
Minder, A.M., Widmer, A., 2008. A population genomic analysis of species boundaries: trends. Front. Zool. 4, 11.
neutral processes, adaptive divergence and introgression between two hybridizing Schönswetter, P., Tribsch, A., Barfuss, M., Niklfeld, H., 2002. Several Pleistocene refugia
plant species. Mol. Ecol. 17 (6), 1552–1563. detected in the high alpine plant Phyteuma globulariifolium Sternb. & Hoppe
Muhlfeld, C.C., Kovach, R.P., Jones, L.A., Al-Chokhachy, R., Boyer, M.C., Leary, R.F., (Campanulaceae) in the European Alps. Mol. Ecol. 11, 2637–2647.
et al., 2014. Invasive hybridization in a threatened species is accelerated by climate Schumer, M., Rosenthal, G.G., Andolfatto, P., 2014. How common is homoploid hybrid
change. Nat. Clim. Change 4, 620–624. speciation? Evolution 68, 1553–1560.
Nardi, G., 2011. Clausiliidae (Gastropoda, Pulmonata) from Lombardy (northern Italy), Schumer, M., Rosenthal, G.G., Andolfatto, P., 2018. What do we mean when we talk
with the description of a new subspecies. Basteria 75, 95–103. about hybrid speciation? Heredity 120, 379–382.
Nieto Feliner, G., Álvarez, I., Fuertes-Aguilar, J., Heuertz, M., Marques, I., Moharrek, F., Seehausen, O., Takimoto, G., Roy, D., Jokela, J., 2008. Speciation reversal and
Villa-Machío, I., 2017. Is homoploid hybrid speciation that rare? An empiricist’s biodiversity dynamics with hybridization in changing environments. Mol. Ecol. 17
view. Heredity 118, 513–516. (1), 30–44.
Nordsieck, H., 1963. Zur Anatomie und Systematik der Clausilien, II. Die Formenbildung Seguinot, J., Ivy-Ochs, S., Jouvet, G., Huss, M., Funk, M., Preusser, F., 2018. Modelling
des Genus Delima in den Südalpen. Archiv für Molluskenkunde 92, 169–203, pl. last glacial cycle ice dynamics in the Alps. Cryosphere 12 (10), 3265–3285.
3–3a. Sokolov, E.P., 2000. An improved method for DNA isolation from mucopolysaccharide-
Nordsieck, H., 1984. Neue Taxa rezenter europäischer Clausilien, mit Bemerkungen zur rich molluscan tissues. J. Molluscan Stud. 66, 573–575.
Bastardierung bei Clausilien (Gastropoda: Clausiliidae). Archiv für Molluskenkunde Stamatakis, A., 2014. RAxML Version 8: A tool for phylogenetic analysis and post-
114, 189–211. analysis of large phylogenies. Bioinformatics 30, 1312–1313.
Nordsieck, H., 2002. Contributions to the knowledge of the Delimini (Gastropoda: Stehlik, I., 2000. Nunataks and peripheral refugia for alpine plants during quaternary
Stylommatophora: Clausiliidae). Mitteilungen der deutschen malakozoologischen glaciation in the middle parts of the Alps. Bot. Helv. 110, 25–30.
Gesellschaft 67, 27–39. Taberlet, P., Fumagalli, L., Wust-Saucy, A.-G., Cosson, J.-F., 1998. Comparative
Nordsieck, H., 2007. Worldwide Door Snails (Clausiliidae), Recent and Fossil. phylogeography and postglacial colonization routes in Europe. Mol. Ecol. 7,
ConchBooks, Hackenheim, Germany. 453–464.
Paris, J.R., Stevens, J.R., Catchen, J.M., 2017. Lost in parameter space: a road map for Thiel-Egenter, C., Alvarez, N., Holderegger, R., Tribsch, A., Englisch, T., Wohlgemuth, T.,
STACKS. Methods Ecol. Evol. 8 (10), 1360–1373. Gugerli, F., 2011. Break zones in the distributions of alleles and species in alpine
Pattengale, N.D., Alipour, M., Bininda-Emonds, O.R.P., Moret, B.M.E., Stamatakis, A., plants. J. Biogeogr. 38, 772–782.
2010. How many bootstrap replicates are necessary? J. Comput. Biol. 17, 337–354. Thiel-Egenter, C., Holderegger, R., Brodbeck, S., Intrabiodiv Consortium, Gugerli, F.,
Patterson, N., Moorjani, P., Luo, Y., Mallick, S., Rohland, N., Zhan, Y., Reich, D., 2012. 2009. Concordant genetic breaks, identified by combining clustering and tessellation
Ancient admixture in human history. Genetics 192, 1065–1093. methods, in two co-distributed alpine plant species. Mol. Ecol. 18 (21), 4495–4507.
Penck, A., Brückner, E., 1909. Die Alpen im Eiszeitalter, vol. 3. Tauchnitz, Leipzig, Tribsch, A., 2004. Areas of endemism of vascular plants in the Eastern Alps in relation to
Germany. Pleistocene glaciation. J. Biogeogr. 31, 747–760.
Peterson, B.K., Weber, J.N., Kay, E.H., Fisher, H.S., Hoekstra, H.E., 2012. Double digest Tribsch, A., Schönswetter, P., 2003. Patterns of endemism and comparative
RADseq: an inexpensive method for de novo SNP discovery and genotyping in model phylogeography confirm palaeoenvironmental evidence for Pleistocene refugia in
and non-model species. PLoS ONE 7, e37135. the Eastern Alps. Taxon 52, 477–497.
Pickrell, J.K., Pritchard, J.K., 2012. Inference of population splits and mixtures from Uit de Weerd, D.R., Gittenberger, E., 2013. Phylogeny of the land snail family
genome-wide allele frequency data. PLoS Genet. 8, e1002967. Clausiliidae (Gastropoda: Pulmonata). Mol. Phylogenet. Evol. 67, 201–216.
Pritchard, J.K., Stephens, M., Donnelly, P., 2000. Inference of population structure using Vernesi, C., Hoban, S.M., Pecchioli, E., Crestanello, B., Bertorelle, G., Rosà, R., Hauffe, H.
multilocus genotype data. Genetics 155, 945–959. C., 2016. Ecology, environment and evolutionary history influence genetic structure
R Core Team. 2020. R: A language and environment for statistical computing, version in five mammal species from the Italian Alps. Biol. J. Linn. Soc. 117 (3), 428–446.
4.0.2. R Foundation for Statistical Computing, Vienna, Austria. Available from: Voges, A., 1995. International Quaternary Map of Europe. Sheet 10 Bern. Hannover,
http://CRAN.R-project.org/. Germany: Bundesanstalt für Geowissenschaften und Rohstoffe/UNESCO.
Rhymer, J.M., Simberloff, D., 1996. Extinction by hybridization and introgression. Annu. Welter-Schultes, F., 2012. European non-marine molluscs, a guide for species
Rev. Ecol. Syst. 27, 83–109. identification. Planet Poster Editions, Göttingen, Germany.
Scheel, B., Hausdorf, B., 2012. Survival and differentiation of subspecies of the land snail Zhang, C., Rabiee, M., Sayyari, E., Mirarab, S., 2018. ASTRAL-III: polynomial time
Charpentieria itala in mountain refuges in the Southern Alps. Mol. Ecol. 21, species tree reconstruction from partially resolved gene trees. BMC Bioinf. 19, 153.
3794–3808.

11

You might also like