Research Article
Print
Research Article
On the verge of extinction – revision of a highly endangered Swiss alpine snail with description of a new genus, Raeticella gen. nov. (Gastropoda, Eupulmonata, Hygromiidae)
expand article infoJeannette Kneubühler§, Markus Baggenstos|, Eike Neubert§
‡ Natural History Museum Bern, Bern, Switzerland
§ University of Bern, Bern, Switzerland
| Oekologische Beratung Markus Baggenstos, Stans, Switzerland
Open Access

Abstract

The phylogenetic status of the alpine land snail Fruticicola biconica has remained questionable since it was described by Eder in 1917. Considered a microendemic species from mountain tops in Central Switzerland, the shell is specially adapted for life under stones. Herein, we show via molecular and anatomical investigations that F. biconica neither belongs to the land snail genus Trochulus, nor to any other genus within Trochulini, but rather warrants placement within the newly established genus Raeticella Kneubühler, Baggenstos & Neubert, 2022. Phylogenetic analyses reveal that R. biconica is clearly separated from Trochulus. These findings are supported by morphological investigations of the shell and genitalia.

Keywords

Endemism, integrative taxonomy, LGM, mountains, nunataks, phylogeny, Switzerland, Trochulus

Introduction

Discovered in the Bannalp, Nidwalden and known from only a few localities in the Central Swiss Alps, Fruticicola biconica was described by the Swiss zoologist Leo Eder in 1917.

Later, F. biconica, known as the Nidwaldner hairy snail, was moved to the widely used genus Trichia W. Hartmann, 1840 and circulated throughout the European literature under this designation (e.g., Kerney et al. 1983). The generic name, Trichia, was subsequently replaced by Trochulus Chemnitz, 1786 due to homonymy with Trichia De Haan, 1839 (Crustacea, Xanthidae).

Previous studies (Pfenninger et al. 2005; Dépraz et al. 2009; Duda et al. 2014; Kruckenhauser et al. 2014; Proćków et al. 2021) included T. biconicus individuals in their genetic analyses of Trochulus species. Pfenninger et al. (2005) and Dépraz et al. (2009) used the same sequence of T. biconicus collected at the type locality at Bannalp. This sequence clustered within the so far known Trochulus species and some newly identified lineages, which were not further described (fig. 2 in Pfenninger et al. 2005; fig. 1 in Dépraz et al. 2009). Most likely, Pfenninger et al. (2005) and Dépraz et al. (2009) used misidentified specimens in their phylogenetic studies, or some samples were mixed. Since these authors did not publish images of the investigated specimens, an unequivocal identification is not possible. Duda et al. (2014) and Kruckenhauser et al. (2014) found that T. biconicus, “T. oreinos oreinos” (A.J. Wagner, 1915), and “T. oreinos scheerpeltzi” (Mikula, 1957) form basal lineages in comparison to specimens of Trochulus s. str. The latter two taxa were elevated from subspecies to species level (Bamberger et al. 2020) and are today known to belong to the newly described genus Noricella (Neiber et al. 2017). Proćków et al. (2021) found the same phylogenetic pattern as Duda et al. (2014) and Kruckenhauser et al. (2014) and questioned the affiliation of biconicus to Trochulus. Already Turner et al. (1998) had disputed the phylogenetic position of T. biconicus. Until today, the phylogenetic position of T. biconicus within the Trochulini remained unclear. Hence, an integrative taxonomic approach is applied in this study to investigate the phylogenetic affiliation of T. biconicus.

Materials and methods

Specimens investigated

Living individuals of T. biconicus were collected in September 2020 at 11 sites of the known distribution area in Central Switzerland (see Fig. 1 for detailed sampling localities). Trochulus biconicus is classified as Vulnerable by Swiss law (Federal Office of Environment) and is protected. It is also considered Endangered by the IUCN (https://www.iucnredlist.org/species/22107/9360310). Collecting permits were obtained from the cantonal administrations of Nidwalden, Obwalden, and Uri. At each site, 3–5 snails were collected from large populations (>20 individuals) from under rocks on stony outcrops. The individual snails were preserved in 80% ethanol to keep the body tissue soft for proper anatomical investigations and DNA extraction. In Table 1, sampling localities and GenBank accession numbers are listed for all sequenced specimens of T. biconicus, Trochulus spp., and Edentiella edentula. Usually, two specimens of T. biconicus per population were sequenced. Those not destroyed in the extraction process are deposited at the NMBE as voucher material. The map was produced with QGIS (2016, v. 2.18.13) using the Natural Earth data set.

Table 1.

Sequenced T. biconicus specimens from Central Switzerland. Asterisk (*) marks the type localities of the species studied. Additionally, Edentiella edentula (Draparnaud, 1805) and some species of Trochulus were sequenced and included for phylogenetic analyses.

Voucher-No. Species Locality Coordinates Altitude [m] GenBank accession number COI GenBank accession number 16S GenBank accession number ITS2
NMBE 567164 T. biconicus Bannalp Schonegg* 46.87°N, 8.46°E 2232 MW435154 MW433778 MW433799
NMBE 567165 T. biconicus Bannalp Schonegg* 46.87°N, 8.46°E 2232 MW435155 MW433779 MW433800
NMBE 567167 T. biconicus Chaiserstuel 46.87°N, 8.46°E 2263 MW435156 MW433780 MW433801
NMBE 567168 T. biconicus Chaiserstuel 46.87°N, 8.46°E 2263 MW435157 MW433781 MW433802
NMBE 567149 T. biconicus Wissberg I 46.81°N, 8.47°E 2335 MW435158 MW433782 MW433803
NMBE 567150 T. biconicus Wissberg I 46.81°N, 8.47°E 2335 MW435159 MW433783 MW433804
NMBE 567152 T. biconicus Wissberg II 46.81°N, 8.47°E 2355 MW435160 MW433784 MW433805
NMBE 567153 T. biconicus Wissberg II 46.81°N, 8.47°E 2355 MW435161 MW433785 MW433806
NMBE 567155 T. biconicus Widderfeld I 46.83°N, 8.33°E 2120 MW435162 MW433786 MW433807
NMBE 567156 T. biconicus Widderfeld I 46.83°N, 8.33°E 2120 MW435163 MW433787 MW433808
NMBE 567159 T. biconicus Widderfeld II 46.83°N, 8.33°E 2290 MW435164 MW433788 MW433809
NMBE 567161 T. biconicus Brisen I 46.90°N, 8.45°E 2045 MW435165 MW433789 MW433810
NMBE 567137 T. biconicus Brisen I 46.90°N, 8.45°E 2045 MW435166 MW433790 MW433811
NMBE 567139 T. biconicus Brisen II 46.90°N, 8.46°E 2130 MW435167 MW433791 MW433812
NMBE 567140 T. biconicus Brisen II 46.90°N, 8.46°E 2130 MW435168 MW433792 MW433813
NMBE 567142 T. biconicus Brisen III 46.90°N, 8.46°E 2090 MW435169 MW433793 MW433814
NMBE 567143 T. biconicus Brisen III 46.90°N, 8.46°E 2090 MW435170 MW433794 MW433815
NMBE 567145 T. biconicus Gitschen I 46.88°N, 8.57°E 1890 MW435171 MW433795 MW433816
NMBE 567146 T. biconicus Gitschen I 46.88°N, 8.57°E 1890 MW435172 MW433796 MW433817
NMBE 567148 T. biconicus Gitschen II 46.88°N, 8.57°E 1970 MW435173 MW433797 MW433818
NMBE 567162 T. biconicus Gitschen II 46.88°N, 8.57°E 1970 MW435174 MW433798 MW433819
NMBE 568100 T. hispidus Sweden, prov. Uppland, Uppsala, Linnaeus Garden* 59.8619°N, 17.6342°E ON477947
NMBE 568103 T. hispidus Sweden, Östergötland, Vist 58.3294°N, 15.729°E 70 ON477948 ON479908 ON479901
NMBE 564609 Trochulus sp. Bullet, Le Chasseron 46.8517°N, 6.5377°E 1606 ON477944 ON479905 ON479898
NMBE 564607 Trochulus sp. Mervelier, Scheltental 47.336°N, 7.5153°E 615 ON477943 ON479904 ON479897
NMBE 543063 Trochulus sp. St-Cergue, Route de Cuvaloup 46.4487°N, 6.123°E 1208 ON477941 ON479902 ON479895
NMBE 564601 Trochulus sp. Zernez 46.6998°N, 10.0943°E 1473 ON477942 ON479903 ON479896
NMBE 568094 Trochulus sp. Lac du Mont d’Orge 46.2321°N, 7.333°E 624 ON477946 ON479907 ON479900
NMBE 565821 T. alpicola Bannalp Schonegg* 46.8706°N, 8.2491°E 2234 ON477945 ON479906 ON479899
MNHW_S_15_29_101 T. villosus Montagne De Cernier 47.0763°N, 6.8888°E 1385 MW440985 MW447773 MW440678
MNHW_S_15_29_02 T. clandestinus Montagne De Cernier 47.0763°N, 6.8888°E 1385 MW440983 MW447772 MW440676
MNHW_S_15_27_12 Trochulus sp. Gorges de Court 47.2553°N, 7.3439°E 608 MW440984 MW621002 MW440677
MNHW_S_15_21_02 T. caelatus Gorges de Moutier* 47.2856°N, 7.3819°E 477 MW440982 MW621001 MW440675
MNHW_S_Er_50 E. edentula Erschwil 47.3673°N, 7.555°E 459 MW440986 MW621003 MW440679
MNHW_S_Er_51 E. edentula Erschwil 47.3673°N, 7.555°E 459 MW440987 MW621004 MW440680
Figure 1. 

Sampling locations of the investigated individuals of T. biconicus.

Acronyms of collections

NMBE Natural History Museum Bern, Switzerland;

MNHW Museum of Natural History Wrocław, University of Wrocław, Poland.

Shell morphology and anatomical study of the genitalia

One animal was selected from each population for investigations of the shell morphology and the genital organs. The dissection of the genitalia was performed under a Leica MZ12 stereomicroscope using thin tweezers. The genital organs were removed from the body, spread on a wax-lined bowl and properly pinned with small needles. The total length of the situs was measured using Mitutoyo callipers. Proportions between different parts of the genitalia were estimated using the total situs length as a reference. Additionally, the inner structures of the penis and the penial papilla were investigated. Pictures of the situs and the shells were taken with a Leica M205 microscope camera using an image-processing program (Leica LAS X v. 3.6.0.20104, Switzerland). The shells were imaged in frontal, lateral, apical, and ventral position. Shell height and shell width were measured using the callipers to assess perpendicularity with the shell axis.

Abbreviations used in the anatomical descriptions and figures

AG albumen gland;

BC bursa copulatrix;

DS dart sacs;

Ep epiphallus;

Flflagellum;

HD hermaphroditic duct;

MG mucous glands;

Pe penis;

PP penial papilla;

sh shell height;

sw shell width;

Va vagina;

VD vas deferens.

DNA extraction, PCR amplification and sequence determination

For total DNA extraction of the specimens, the Qiagen Blood and Tissue Kit (Qiagen; Hilden, Germany) was used in combination with a QIAcube extraction robot. Circa 0.5 cm3 of tissue was cut and placed in a mixture of 180 µl ATL buffer and 20 µl Proteinase K. It was then incubated for ca. 4 hours at 56 °C in a heater (Labnet, Vortemp 56, witec AG, Littau, Switzerland). For subsequent DNA extraction, the QIAcube extraction robot was used with the Protocol 430 (DNeasy Blood Tissue and Rodent tails Standard). In this study, two mitochondrial markers (COI and 16S) and one nuclear marker (5.8S rRNA+ITS2) were investigated. PCR mixtures consisted of 12.5 µl GoTaq G2 HotStart Green Master Mix (Promega M7423), 4.5 µl ddH2O, 2 µl forward and reverse primer each, and 4 µl DNA template. The primer pairs implemented for the PCR are listed in Table 2. The following PCR cycles were used: for COI, 2 min at 94 °C, followed by 40 cycles of 1 min at 95 °C, 1 min at 47 °C and 1 min at 72 °C and finally, 5 min at 72 °C; for 16S, 3 min at 96 °C, followed by 40 cycles of 30 s at 94 °C, 30 s at 50 °C and 30 s at 72 °C, and finally, 1 min at 72 °C; and for 5.8S rRNA+ITS2, 3 min at 94 °C, followed by 40 cycles of 30 s at 94 °C, 30 s at 50 °C and 30 s at 72 °C, and finally, 5 min at 72 °C (SensoQuest Tabcyclet and Techne TC-512, witec AG, Littau, Switzerland). The purification and sequencing of the PCR product was performed by LGC (LGC Genomics Berlin, Germany).

Table 2.

Primer pairs used for PCR.

Gene Primer Sequence Sequence length (bp) Reference
COI LCO1490 5′-GGTCAACAAATCATAAAGATATTGG-3′ 655 Folmer et al. 1994
HCO2198 5′-TAAACTTCAGGGTGACCAAAAAATCA-3′
16S 16S cs1 5′-AAACATACCTTTTGCATAATGG-3′ 440 Chiba 1999
16S cs2 5′-AGAAACTGACCTGGCTTACG-3′
ITS2 ITS2 LSU1 5′-GCTTGCGGAGAATTAATGTGAA-3′ 900 Wade and Mordan 2000
ITS2 LSU3 5′-GGTACCTTGTTCGCTATCGGA-3′

Phylogenetic analyses

The phylogenetic analyses were conducted using sequences obtained from GenBank and from this study, which were included as outgroup: Ichnusotricha berninii Giusti & Manganelli, 1987, Plicuteria lubomirskii (Ślósarski, 1881), Petasina unidentata (Draparnaud, 1805), Noricella oreinos (A.J. Wagner, 1915), Noricella scheerpeltzi (Mikula, 1957) (GenBank numbers and sampling localities published by Neiber et al. 2017), Edentiella edentula (Draparnaud, 1805), and several ingroup specimens of Trochulus (Table 1). These species were selected to identify the phylogenetic position of T. biconicus.

For sequence processing and editing, the software package Geneious v. 9.1.8 (Biomatters Ltd) was used. Topologies were estimated using two different phylogenetic methods: Bayesian Inference (BI) and Maximum Likelihood (ML). Bayesian Inference was performed using Mr. Bayes v. 3.2.6 x64 (Huelsenbeck and Ronquist 2001; Ronquist and Huelsenbeck 2003; Altekar et al. 2004) via the HPC cluster from the University of Bern (http://www.id.unibe.ch/hpc). Evolutionary models for each subset were set to mixed models. The Monte Carlo Markov Chain (MCMC) parameter was set as follows: starting with four chains and four separate runs for 20 million generations with a tree sampling frequency of 1000 and a burn in of 25%. RAxML plug-in for Geneious (Stamatakis 2014) was implemented for computing ML inference, using Geneious’ plug-in with rapid bootstrapping setting, the search for the best scoring ML tree and 1000 bootstrapping replicates. The model, GTR CAT I was implemented.

Results

Phylogenetic analyses

The BI analysis of the concatenated data set (Fig. 2) shows two major clades within the tribe Trochulini. These two clades are separated with full support. One clade contains representative specimens of Edentiella and Noricella which form a polytomy. The second major clade within Trochulini contains representatives of Petasina, Trochulus, and the investigated T. biconicus specimens. Trochulus biconicus is the sister lineage to the selected Trochulus specimens. This node has full posterior probability support. Trochulus hispidus from the type locality in Sweden clusters together with a second specimen from Sweden and forms the sister group to two Swiss Trochulus specimens from Zernez and Lac du Mont d’Orge. The resolution within the T. biconicus clade is moderate because the investigated individuals differ in only few nucleotides in all three investigated markers.

Figure 2. 

Bayesian Inference (BI) tree based on the concatenated data set of COI, 16S, and 5.8S rRNA+ITS2. Numbers represent Bayesian posterior probabilities.

The ML analysis of the concatenated data set (Fig. 3) shows a similar topology as that of the BI analysis. The difference in the ML and the BI tree is the relationship of Edentiella and Noricella. In the ML tree, E. edentula clusters together with N. scheerpeltzi. This node has low support value (bootstrap support of 51 in Fig. 3), whereas in the BI analysis (Fig. 2), Edentiella and Noricella show a polytomy. In both analyses, T. biconicus forms the sister lineage to the selected Trochulus species. This node has full ML support. The support values within the Trochulus clade are moderate to high.

Figure 3. 

Maximum Likelihood (RAxML) tree based on the concatenated data set of COI, 16S, and 5.8S rRNA+ITS2. Numbers represent bootstrap support values from the ML analysis.

The p-distance, which shows the number of base differences per site from between sequences (Kumar et al. 2018) for the COI was calculated using MEGA v. 10.1.8 (https://www.megasoftware.net/). The p-distance for T. biconicus and the remaining investigated Trochulus species ranges from 0.153–0.189, for T. biconicus and E. edentula from 0.183–0.189, for T. biconicus and Noricella species from 0.128–0.166, for T. biconicus and P. unidentata from 0.171–0.176, for T. biconicus and I. berninii from 0.142–0.147 and for T. biconicus and P. lubomirksi from 0.177–0.188 (see Suppl. material 1). The genetic investigations in this study clearly show that T. biconicus is neither a member of the Trochulus clade nor does it belong to another known genus in the Trochulini. It thus, warrants designation in a separate new genus.

Figure 4. 

Trochulus biconicus (NMBE 567151) collected from Wissberg I A shell, sw = 5.56 mm, sh = 2.55 mm B situs C penis (Pe) with penial papilla (PP) D cross section of the penial papilla. Shell × 5.

Shell morphology

The shell of T. biconicus is flattened, tightly coiled, and beige to brownish. The mean shell width of the investigated individuals (N = 13) is 5.63 mm (range: 5.3–6.1 mm; SD = 0.23 mm) with mean shell height reaching 2.67 mm (range: 2.34–2.9 mm; SD = 0.17 mm) (Table 3). The shell bears 5.5–6 whorls which increase only slightly in width towards the perimeter. The umbilicus is entirely open and wide. The crescent-shaped aperture contains a white, poorly developed lip. Neither juveniles nor adults show hairs on the shell (Figs 410).

Table 3.

Morphological analysis: measurements of the shell and genital organs of T. biconicus and T. clandestinus. Additionally, some collected dry shells from Bannalp Schonegg (NMBE 567170) and Chaiserstuel (NMBE 567171) were included in the analysis. Asterisk (*) marks the type locality of T. biconicus. Umbilicus minor diameter is measured according to Proćków (2009). All measurements are in mm.

Voucher No. Species Locality Coordinates Altitude [m] shell width shell height umbilicus minor diameter penis length epiphallus length flagellum length Figure number
NMBE 567151 T. biconicus Wissberg I 46.81°N, 8.47°E 2335 5.56 2.55 0.8 1.81 2.01 5.98 Fig. 4
NMBE 567160 T. biconicus Widderfeld II 46.83°N, 8.33°E 2290 5.73 2.59 0.88 2.79 3.42 8.13 Fig. 5
NMBE 567138 T. biconicus Brisen I 46.90°N, 8.45°E 2045 5.61 2.34 0.73 2.86 3.06 7.25 Fig. 6
NMBE 567163 T. biconicus Gitschen II 46.88°N, 8.57°E 1970 5.67 2.87 0.84 2.23 3.67 6.38 Fig. 7
NMBE 567166 T. biconicus Bannalp Schonegg* 46.87°N, 8.46°E 2232 5.75 2.76 0.96 1.84 1.98 4.26 Fig. 8
NMBE 567169 T. biconicus Chaiserstuel 46.87°N, 8.46°E 2263 5.3 2.46 0.99 2.66 3.21 4.1 Fig. 9
NMBE 567170_1 T. biconicus Bannalp Schonegg* 46.87°N, 8.46°E 2232 5.7 2.82 1.19 Fig. 10A
NMBE 567170_2 T. biconicus Bannalp Schonegg* 46.87°N, 8.46°E 2232 6.03 2.73 1.08 Fig. 10B
NMBE 567170_3 T. biconicus Bannalp Schonegg* 46.87°N, 8.46°E 2232 5.39 2.54 0.99 Fig. 10C
NMBE 567170_4 T. biconicus Bannalp Schonegg* 46.87°N, 8.46°E 2232 6.1 2.9 1.07 Fig. 10D
NMBE 567171_1 T. biconicus Chaiserstuel 46.87°N, 8.46°E 2263 5.46 2.76 1.08 Fig. 10E
NMBE 567171_2 T. biconicus Chaiserstuel 46.87°N, 8.46°E 2263 5.41 2.61 0.79 Fig. 10F
NMBE 567171_3 T. biconicus Chaiserstuel 46.87°N, 8.46°E 2263 5.46 2.83 0.89 Fig. 10G
NMBE 571318 T. clandestinus Bern, Bümpliz 46.9435°N, 7.3922°E 540 9.64 5.57 1.29 4.24 5.79 4.69 Fig. 11
Figure 5. 

Trochulus biconicus (NMBE 567160) collected from Widderfeld II A shell, sw = 5.73 mm, sh = 2.59 mm B situs C penis (Pe) with penial papilla (PP) D cross section of the penial papilla. Shell × 5.

Figure 6. 

Trochulus biconicus (NMBE 567138) collected from Brisen I A shell, sw = 5.61 mm, sh = 2.34 mm B situs C penis (Pe) with penial papilla (PP). Shell × 5.

Figure 7. 

Trochulus biconicus (NMBE 567163) collected from Gitschen II A shell, sw = 5.67 mm, sh = 2.87 mm B situs C penis (Pe) with penial papilla (PP). Shell × 5.

Morphology of the genitalia

The genitalia are characterised by four stylophores, symmetrically placed in two pairs on both sides of the vagina (see fig. 11 in Proćków 2009). The inner dart sacs are somewhat longer and slenderer than the outer sacs. The outer stylophores contain the love darts (see also Proćków 2009). The mucous glands consist of four simple and thin tubes branching off the free oviduct directly above the dart sacs. The vagina is a rather long tube, which is almost smooth inside or shows some faint elongate tissue folds that connect to the atrium (not shown in the figures). The bursa copulatrix branches off from the free oviduct above the dart sacs and the mucous glands and is terminated by an elongated vesicle.

Figure 8. 

Trochulus biconicus (NMBE 567166) collected from Bannalp Schonegg A shell, sw = 5.75 mm, sh = 2.76 mm B situs C penis (Pe) with penial papilla (PP). Shell × 5.

Figure 9. 

Trochulus biconicus (NMBE 567169) collected from Chaiserstuel A shell, sw = 5.30 mm, sh = 2.46 mm B situs C penis (Pe) with penial papilla (PP). Shell × 5.

Figure 10. 

Shells of Trochulus biconicus from Bannalp Schonegg (A–D) and from Chaiserstuel (E–G).

The penis is fusiform and contains a club-shaped penial papilla which points into the lumen of the penial chamber. The epiphallus is as long as the penis; the penis retractor muscle inserts at the transition zone between epiphallus and penis. The flagellum is about 1.5× the length of the penis and epiphallus each. The epiphallial lumen contains longitudinal tissue ridges (e.g., Fig. 4C). The penial chamber is characterised by smooth walls. The penial papilla contains a lateral subapical pore. The cross section of the penial papilla (Figs 4D, 5D) reveals a central duct surrounded by small folds.

The anatomy of the genitalia of T. clandestinus differs from T. biconicus by having eight long, thin mucous glands (Fig. 11). The inner dart sacs of the investigated T. clandestinus are slightly longer in length than the outer dart sacs. The flagellum has about the same length as the bulbous penis, and the epiphallus is slightly longer than the penis. The cross section of the penial papilla differs in T. clandestinus by having several tissue layers around the main tube of the penial papilla (Fig. 11D).

Figure 11. 

Trochulus clandestinus (NMBE 571318) collected from Bümpliz, Bern, Switzerland A shell, sw = 9.64 mm, sh = 5.57 mm B situs C penis (Pe) with penial papilla (PP) D cross section of the penial papilla. Shell × 3.

Taxonomic and systematic implications

The fully supported split between T. biconicus and currently known Trochulus species (Figs 2, 3) warrants description of a new genus, Raeticella gen. nov., based on Fruticicola biconica.

Raeticella gen. nov.

Type species

Fruticicola biconica Eder, 1917.

Trochulus Chemnitz, 1786

Trochulus biconicus (Eder, 1917)

Diagnosis

Shell flattened and thin-walled, translucent, compressed in the direction of the axis; no trichome formation; whorls 5.5–6, gradually increasing so that the body whorl is only about twice as wide as the first whorl; the aperture is oblique, narrow, crescent-shaped; lip sharp, whitish and slightly reflexed; the four mucous glands are long, thick and pointed; penis and epiphallus are about the same length; the flagellum is barely separated from the epiphallus.

Differential diagnosis

Raeticella gen. nov. differs from Trochulus by having a flat, biconical shell, devoid of any periostracal hairs, even in juveniles, and in having only four instead of occasionally six or eight (see Duda et al. 2014) mucous glands. It differs from Noricella by lacking a basal tooth, being devoid of any periostracal hairs, the absence of coarse ripples and the absence of an additional fold and bulge in the penial papilla, which occurs in N. oreinos (Duda et al. 2014).

Etymology

The name is derived from the Roman province of Raetia, which comprised within its larger expansion, the area of what is now known as eastern and central Switzerland. It also refers to the generic name, Noricella, which is another recently detected spin-off from Trochulus and whose name derives in part from the eastern border province of Raetia (Noricum – now Austria and Slovenia).

Discussion

Neiber et al. (2017) clarified the phylogenetic positions of some species within the Trochulini by establishing the new genus Noricella Neiber, Razkin & Hausdorf, 2017. In their study it was proven that N. oreinos and N. scheerpeltzi differed from the closest related genus Edentiella Poliński, 1929 in some apomorphic nucleotide substitutions and by morphological characters. Edentiella contains at least one longitudinal septum separating an additional lacuna in the penial papilla which is lacking in N. oreinos, in most Trochulus species, and in Petasina (Neiber et al. 2017). These authors also included some representatives of Trochulus but did not have specimens of R. biconica available. Turner et al. (1998) had already considered R. biconica to be only distantly related to Trochulus s. str. because of 1) the lack of periostracal hair even in juveniles, 2) a very long flagellum, and 3) only four instead of six or eight mucous glands. Hence, Turner (1991) suggested to move R. biconica into a subgenus of Trochulus. The questionable position of biconicus in Trochulus was recently re-addressed by Proćków et al. (2021). In our analysis, the calculated p-distance of R. biconica and the investigated Trochulus specimens comprises the highest values. The p-distance of R. biconica and Noricella species is lower than for Trochulus, which means that Raeticella is genetically closer, based on COI, to Noricella than to Trochulus. Even Ichnusotricha, which belongs to the tribe of Ganulini is genetically more similar to Raeticella than Trochulus is to Raeticella.

The shell morphology of R. biconica differs from all known Trochulus species by having a flat shell with a low spire. The last whorl is bluntly keeled. Adults are always hairless (Proćków 2009). In this regard, it is most like the shells of the two Noricella species (Duda et al. 2011, 2014), but the anatomy of the genital organs of these species is different. Both Noricella species have four pairs of mucous glands, compared to two pairs in R. biconica. Noricella oreinos possesses an additional fold and bulge in the penial papilla, which seems to be unique to this species (Duda et al. 2014). The section of the penial papilla in R. biconica shows similar internal features as in T. caelatus (Proćków 2009), T. striolatus (Proćków 2009; Duda et al. 2014; Proćków et al. 2021), and T. suberectus (Proćków 2009). Raeticella biconica does not possess periostracal hairs, neither as a juvenile nor as an adult. This, however, is considered a typical feature for Trochulus species (Proćków 2009).

Hewitt (2004) observed that many taxa in temperate refugial regions in Europe and North America show relatively deep DNA divergence, indicating their presence over several ice ages and suggesting a mode of speciation by repeated allopatry. On the one hand, this possibly explains the deep split between Raeticella and Trochulus and shifts the splitting event of these groups to the Pliocene. On the other hand, we observed a low genetic diversity within our analysed populations. So, this species probably underwent a bottleneck event during the Pleistocene and the Last Glacial Maximum (LGM). Some isolated populations obviously survived this icy period. The LGM lasted about 30–19 ka in the Alps. During that period, this area was covered by massive ice sheets, and the glaciers reached out to the forelands of both, the northern and southern side of the main alpine chains. However, mountain tops above more than 2000 m were not covered by ice during the LGM. The recession of the glaciers from their maximum extent started around 24 ka (see Ivy-Ochs 2015). We hypothesize that the original distribution area of R. biconica was much larger, but only a few individuals survived on neighbouring nunataks (glacial islands) during the LGM. A similar scenario is assumed for the evolution of the two Noricella species (Duda et al. 2011, 2014; Kruckenhauser et al. 2014). Gittenberger et al. (2004) also hypothesized the survival of Arianta arbustorum alpicola (A. Férussac, 1821) on nunataks. A similarly fragmented distribution pattern can be observed in the eastern alpine mollusc species Cylindrus obtusus (Draparnaud, 1805) (Schileyko 2012: 95, fig. 2). Schileyko argued that the missing fossil record for this species proves that it was formed at the end of the Würm glaciation approximately 10–12 ka ago. As a species adapted to cold environmental conditions, this species was then assumed to be forced to follow the retreating snow and ice fields, which subsequently lead to habitat fragmentation. This assumption requires an ancestor from interglacials (which is also not found in the fossil record), and has to explain the rapid transformation of an Ariantine species from a globular or even depressed shell to a turriform shell. This is most unlikely. Based on COI sequences, Cadahía et al. (2014) estimated 1.5–12 mya for the split between Arianta and Cylindrus. So, we assume that Raeticella gen. nov., like the monotypic genus Cylindrus, evolved much earlier and survived the Pleistocene by chance on nunatak mountain tops.

The current distribution pattern does not necessarily and strictly reflect the “survivor” populations. ARNAL (2018) found a limited gene flow between the “isolated” populations of R. biconica. This shows that dispersal is not completely impossible, but, due to the high-altitude adaptation of the species, it is rather limited to other, hitherto unpopulated high alpine areas. Possible vectors may be large pasturing animals like sheep and goats, but also ibex, chamois, or birds.

In alpine environments, microendemic species with a relict distribution pattern may occur, which were much more widespread in earlier times. They are now restricted to a very small area due to changes in environmental condition (Turner 1991; Cook 2008; Veron et al. 2019). The distribution area of R. biconica is currently known to encompass 150 isolated sites on both sides of the Engelberger valley, all situated between 1890 and 2575 m of altitude (Baggenstos 2010).

The habitat of R. biconica is very special, and only few other snail species are known to survive in this harsh environment (Eder 1917; Baggenstos 2010). Apart from the occurrence of limestone scree, the snails very much depend on small-scale relief. Slope edges or hilltops, ridges and summits as well as rocky heads and rocky steps are more likely to be colonised by the snail than slope hollows and slope foothills. The highest density of R. biconica is reached in areas with more than 50% of rocky scree (Baggenstos 2010). All these sites are covered with snow for a relatively short time in winter. With its flat shell, R. biconica is perfectly adapted to live under or between stones (Figs 12, 13). Flatness was interpreted as an adaptation to the cold climate at the top of the mountains and may protect the animals from predators (Baur 1987). When it gets too hot, the snails retreat into the ground. The individuals are mainly active during night (Baggenstos 2010). Almost all known R. biconica habitats are blue grass meadows. These are alpine grasslands rich in flowers with a great diversity and a remarkably high proportion of Leguminosae. The prominent structural elements are Sesleria caerulea and Carex sempervirens. The soil cover is relatively thin, interspersed with gravel and stones and dries out quickly (Delarze et al. 2008). Wigger (2007) observed that R. biconica mainly feeds on decaying leaves of blue grass (Sesleria caerulea). The landscape of these meadows is strongly influenced by extensive pasturing and hiking tourism. Pasture animals like sheep, goats, and cows can modify the position of large stones and thus create new micro habitats for the snails. However, stronger interventions, such as the removal of stones or a climate-related transfer of the rubble-rich sites into closed meadows or woodland formations would cause the snail to disappear (Turner 1991).

Figure 12. 

Typical habitat of Raeticella biconica. This photograph was taken on 25.05.2009 on Chaiserstuel (46.8762°N, 8.4671°E, 2263 m) by Markus Baggenstos.

This stenoecious species is prone to extinction because of climate change. Over the last 100 years temperatures have increased by about 0.12–0.20 °C per decade in the Swiss Alps and the snow seasons have shortened (Kohler et al. 2014). Raeticella biconica already reached the summits of the mountains in their vicinity, and there is no more alternative for avoiding unsuitable climate conditions. Considering that global warming is ongoing, R. biconica may well become extinct in just a few years.

Figure 13. 

Close-up of Raeticella biconica crawling on the underside of a stone. This photograph was taken on 09.09.2009 on Chaiserstuel (46.8762°N, 8.4671°E, 2263 m) by Markus Baggenstos.

Conclusion

Long known morphological characteristics in conjunction with our genetic analyses show that R. biconica should be assigned to a new genus. Morphologically, the investigated individuals of R. biconica strongly resemble N. oreinos (Duda et al. 2011). But the genetic analyses of several different species from all genera within Trochulini reveal that R. biconica does not belong to any currently known genus. Therefore, a new monotypic genus within Trochulini is introduced.

Acknowledgements

We are indebted to Małgorzata Proćków for providing tissue samples of E. edentula and some Trochulus specimens, to Ted von Proschwitz for providing T. hispidus samples from the Swedish type locality, to Adrienne Jochum for the linguistic revision, to Tom Burri for paleoecological insights, to the Swiss Federal Office of Environment (FOEN) for financial support (contract no. 110010344 / 8T30/00.5147.PZ/0006), and to the cantonal authorities of Nidwalden (Felix Omlin), Obwalden (Andreas Bacher), and Uri (Georges Eich) for providing sampling permits.

References

  • Altekar G, Dwarkadas S, Huelsenbeck JP, Ronquist F (2004) Parallel Metropolis-coupled Markov chain Monte Carlo for Bayesian phylogenetic inference. Bioinformatics (Oxford, England) 20(3): 407–415. https://doi.org/10.1093/bioinformatics/btg427
  • ARNAL Büro für Natur und Landschaft AG (2018) Projekt Naturschutzgenetik – Schlussbericht: Methodik und Resultate (Insbeso. «Verbund»). Report, Herisau, 45 pp.
  • Baggenstos M (2010) Verbreitung und Biologie der Nidwaldner Haarschnecke (Trochulus biconicus). Report, Stans, 45 pp.
  • Bamberger S, Duda M, Tribsch A, Haring E, Sattmann H, Macek O, Affenzeller M, Kruckenhauser L (2020) Genome-wide nuclear data confirm two species in the Alpine endemic land snail Noricella oreinos s.l. (Gastropoda, Hygromiidae). Journal of Zoological Systematics and Evolutionary Research 58(4): 982–1004. https://doi.org/10.1111/jzs.12362
  • Baur B (1987) Richness of land snail species under isolated stones in a karst area on Öland, Sweden. Basteria 51: 129–133.
  • Cadahía L, Harl J, Duda M, Sattmann H, Kruckenhauser L, Fehér Z, Zopp L, Haring E (2014) New data on the phylogeny of Ariantinae (Pulmonata, Helicidae) and the systematic position of Cylindrus obtusus based on nuclear and mitochondrial DNA marker sequences. Journal of Zoological Systematics and Evolutionary Research 52(2): 163–169. https://doi.org/10.1111/jzs.12044
  • Chemnitz JH (1786) Neues systematisches Conchylien-Cabinet. Neunten Bandes zwote Abtheilung, enthaltend die ausführliche Beschreibung von den Land- und Flußschnecken, oder von solchen Conchylien welche nicht im Meer, sondern auf der Erde und in süssen Wassern zu leben pflegen. Mit zwanzig nach der Natur gemalten und durch lebendige Farben erleuchteten. Raspe, Nürnberg, [xxvi +] 194 pp. [pls 117–136] https://doi.org/10.5962/bhl.title.125416
  • Chiba S (1999) Accelerated evolution of land snails Mandarina in the oceanic Bonin Islands: Evidence from mitochondrial DNA sequences. Evolution; International Journal of Organic Evolution 53(2): 460–471. https://doi.org/10.1111/j.1558-5646.1999.tb03781.x
  • Delarze R, Gonseth Y, Galland P (2008) Lebensräume der Schweiz: Ökologie, Gefährdung, Kennarten. Ott Verlag, Thun, 424 pp.
  • Dépraz A, Hausser J, Pfenninger M (2009) A species delimitation approach in the Trochulus sericeus/hispidus complex reveals two cryptic species within a sharp contact zone. BMC Evolutionary Biology 9(1): e171. https://doi.org/10.1186/1471-2148-9-171
  • Duda M, Sattmann H, Haring E, Bartel D, Winkler H, Harl J, Kruckenhauser L (2011) Genetic differentiation and shell morphology of Trochulus oreinos (Wagner, 1915) and T. hispidus (Linnaeus, 1758) (Pulmonata: Hygromiidae) in the northeastern Alps. Journal of Molluscan Studies 77(1): 30–40. https://doi.org/10.1093/mollus/eyq037
  • Duda M, Kruckenhauser L, Sattmann H, Harl J, Jaksch K, Haring E (2014) Differentiation in the Trochulus hispidus complex and related taxa (Pulmonata: Hygromiidae): morphology, ecology and their relation to phylogeography. Journal of Molluscan Studies 80(4): 371–387. https://doi.org/10.1093/mollus/eyu023
  • Eder L (1917) Eine neue Schweizer Helicide. Revue Suisse de Zoologie 25(15): 442–452.
  • Folmer O, Black M, Hoe W, Lutz R, Vrijenhoek R (1994) DNA primers for amplification of mitochondrial cytochrome c oxidase subunit I from diverse metazoan invertebrates. Molecular Marine Biology and Biotechnology 3(5): 294–299.
  • Gittenberger E, Piel WH, Groenenberg DSJ (2004) The Pleistocene glaciations and the evolutionary history of the polytypic snail species Arianta arbustorum (Gastropoda, Pulmonata, Helicidae). Molecular Phylogenetics and Evolution 30(1): 64–73. https://doi.org/10.1016/S1055-7903(03)00182-9
  • Hartmann JDW (1840–1844) Erd- und Süsswasser-Gasteropoden der Schweiz. Mit Zugabe einiger merkwürdigen exotischen Arten. Scheitlin und Zollikofer, St. Gallen, [i–xx,] 36 pp. [pls 1, 2 [30-06-1840]; 37–116, pls 13–36 [1841]; 117–156, pls 37–60 [1842]; 157–204, pls 61–72 [1843]; 205–227, pls 73–84 [1844]]
  • Hewitt GM (2004) Genetic consequences of climatic oscillations in the Quaternary. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences 359(1442): 183–195. https://doi.org/10.1098/rstb.2003.1388
  • Ivy-Ochs S (2015) Glacier variations in the European Alps at the end of the last glaciation. Cuadernos de Investigación Geográfica 42(2): 295–315. https://doi.org/10.18172/cig.2750
  • Kerney MP, Cameron RAD, Jungbluth JH (1983) Die Landschnecken Nord- und Mitteleuropas. Parey-Verlag, Hamburg/Berlin, 384 pp.
  • Kohler T, Wehrli A, Jurek M (2014) Mountains and climate change: a global concern. Sustainable Mountain Development Series. Centre for Development and Environment (CDE), Swiss Agency for Development and Cooperation (SDC) and Geographica Bernensia, Bern, 136 pp. https://doi.org/10.1659/mrd-journal-d-09-00086.1
  • Kruckenhauser L, Duda M, Bartel D, Sattmann H, Harl J, Kirchner S, Haring E (2014) Paraphyly and budding speciation in the hairy snail (Pulmonata, Hygromiidae). Zoologica Scripta 43(3): 273–288. https://doi.org/10.1111/zsc.12046
  • Kumar S, Steche G, Li M, Knyaz C, Tamura K (2018) MEGA X: Molecular Evolutionary Genetics Analysis across computing platforms. Molecular Biology and Evolution 35(6): 1547–1549. https://doi.org/10.1093/molbev/msy096
  • Neiber MT, Razkin O, Hausdorf B (2017) Molecular phylogeny and biogeography of the land snail family Hygromiidae (Gastropoda: Helicoidea). Molecular Phylogenetics and Evolution 111: 169–184. https://doi.org/10.1016/j.ympev.2017.04.002
  • Pfenninger M, Hrabáková M, Steinke D, Dépraz A (2005) Why do snails have hairs? A Bayesian inference of character evolution. BMC Evolutionary Biology 5(1): e59. https://doi.org/10.1186/1471-2148-5-59
  • Proćków M, Kuznik-Kowalska E, Pienkowska JR, Zeromska A, Mackiewicz P (2021) Speciation in sympatric species of land snails from the genus Trochulus (Gastropoda, Hygromiidae). Zoologica Scripta 50(1): 16–42. https://doi.org/10.1111/zsc.12458
  • Schileyko AA (2012) On the origin of Cochlopupa (= Cylindrus auct.) obtusa (Gastropoda, Pulmonata, Helicidae). Ruthenica: Rossiiskii Malakologicheskii Zhurnal = Russian Malacological Journal 22: 93–100.
  • Turner H (1991) Die Nidwaldner Haarschnecke gibt es sonst nirgends auf der Welt: ein kleines zoologisches Geheimnis lebt auf der Bannalp. Nidwaldner Volksblatt 240(17 Oktober): 13.
  • Turner H, Kuiper JGJ, Thew N, Bernasconi R, Rüetschi J, Wüthrich M, Gosteli M (1998) Atlas der Mollusken der Schweiz und Lichtensteins. Fauna Helvetica 2. Center Suisse de cartographie de la faune, Schweizerische Entomologische Gesellschaft, Eidg. Forschungsanstalt für Wald, Schnee und Landschaft, Neuchâtel, 527 pp.
  • Wade CM, Mordan PB (2000) Evolution within the gastropod molluscs; using the ribosomal RNA gene-cluster as an indicator of phylogenetic relationships. The Journal of Molluscan Studies 66(4): 565–570. https://doi.org/10.1093/mollus/66.4.565
  • Wigger F (2007) Der mikroklimatische und zeitabhängige Aktivitätsrhythmus von Trochulus biconicus. Eine Feldstudie im Rahmen der Projektarbeit in Biogeographie der Universität Basel in Zusammenarbeit mit der Oekologischen Beratung Markus Baggenstos. Basel, unpublished report Basel, 12 pp.

Supplementary material

Supplementary material 1 

Calculated p-distances of the COI of the investigated specimens.

Jeannette Kneubühler, Markus Baggenstos, Eike Neubert

Data type: excel file

Explanation note: Calculated p-distances of the COI of the investigated specimens.

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0/). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this Dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (57.50 kb)
login to comment