Open Access
How to translate text using browser tools
2 February 2023 A Revision of the Rhinolophus hipposideros group (Chiroptera: Rhinolophidae) with Definition of an Additional Species from the Middle East
Petr Benda, Marek Uvizl, Peter Vallo, Antonín Reiter, Marcel Uhrin
Author Affiliations +
Abstract

Initially, the Rhinolophus hipposideros group was defined by two morphological traits, the structure of the nose-leaf and the shape of basioccipital bone of the skull. Originally, it consisted of two species, R. hipposideros and R. midas, whereas currently it is considered to contain a single species, R. hipposideros, under whose rank both original species have been joined. The interpretation of geographic variability within the group has traditionally been based on variation in body and skull size, nose-leaf shape, and several selected skull and tooth characters. This approach resulted in delimitations of up to seven subspecies, mostly in the Mediterranean area, a conception introduced more than a hundred years ago and accepted by many authors till today. We investigated the phylogenetic relationships among populations of R. hipposideros with the help of molecular genetic, morphological, and acoustic examinations. Our analysis uncovered the existence of an unexpected diversity within the R. hipposideros group, challenging its current phylogenetic and taxonomic arrangements. The molecular genetic analysis of almost 100 samples and morphological examinations of about 300 specimens showed two main, geographically exclusive, phylogenetic lineages within the group, well delimited by molecular characteristics and possessing two distinct morphotypes and two distinct echotypes. These two lineages are isolated deep enough to be considered separate species. One of them, R. hipposideros s.str., is widespread over the south-western Eurasia and north-western and north-eastern Africa, and the other, R. midas, is distributed in a small range around the Strait of Hormuz and Gulf of Oman. The extensive range of R. hipposideros s.str. is inhabited at least by two subspecies, separated mainly by the genetic characters, whereas the morphological and echolocation traits do not distinguish the populations sufficiently. The western R. h. hipposideros occurs in the Maghreb and Europe west of the Dnieper River, Bosporus, and the Strait of Karpathos, and the eastern R. h. minimus lives east of this boundary, including the populations of Crimea, Caucasus, the Middle East, and north-eastern Africa (Sudan to Djibouti). The two subspecies also differ in karyotype, with 2n = 58 in R. h. minimus and 2n = 54–56 in R. h. hipposideros. The taxonomic position of the easternmost populations of R. hipposideros s.str. (West Turkestan, Afghanistan, Kashmir) remains unresolved and has to be investigated more elaborately and using a more extensive sample set.

Introduction

The Rhinolophus hipposideros group is one of the numerous groups that divide the genus Rhinolophus Lacépède, 1799, the only genus of the chiropteran family Rhinolophidae. The group currently contains a single species, the lesser horseshoe bat, Rhinolophus hipposideros (André, 1797). Originally, it was defined by Andersen (1905) as the Rhinolophus midas group, comprising two species, R. hipposideros and R. midas Andersen, 1905. This definition of the group was based on a typical structure of the sella of the nose-leaf, bearing a very low and rounded off posterior connecting process, and an extremely narrow basioccipital bone of the skull, reported to be distinct in both characters from other groups of the genus Rhinolophus. Since Andersen (1918) joined the two species into one under the prior name R. hipposideros, this name was also transferred to the group name. The R. hipposideros group was then reported as a separate and monotypic unit within the genus by numerous followers, despite the variable numbers and contents of other groups considered (Allen, 1939; Ellerman and Morrison-Scott, 1951; Koopman, 1994; Horáček et al., 2000; Csorba et al., 2003; Simmons, 2005; Burgin, 2019; etc.).

Besides Andersen's (1905) original definition made on the simple comparison of a few morphological characters, justification of the determination of the R. hipposideros group within the genus Rhinolophus was supported by the results of additional analyses of morphometric data by Bogdanowicz (1992) and genetic data by Guillén Servent et al. (2003), Stoffberg et al. (2010), Foley et al. (2015), and Dool et al. (2016). The basal and very separate position of this group within the genus Rhinolophus was stressed by Guillén Servent et al. (2003), who suggested delimiting it into the subgenus Phyllorhina Leach, 1816.

Because the group currently consists of a single species, R. hipposideros, its intraspecific variation also represents the only variation detectable in the group. This bat is a typical faunal element of the western Palaearctic (Fig. 1), where it occurs in a broad belt of the Mediterranean and temperate zones of Europe, North Africa, and western Asia (Csorba et al., 2003; Gaisler, 2013; Burgin, 2019; Bendjeddou et al., 2022); its distribution range comprises the Mediterranean Maghreb (Morocco to Tripolitania), southern, western and central Europe (from Portugal, Ireland, and Germany to western and southern Ukraine, as well as the Balkans), numerous Mediterranean islands; the Levant, including Sinai; Anatolia; Crimea; the Caucasus region; Iran; Afghanistan; Kashmir; and West Turkestan. Moreover, R. hipposideros also marginally extends to the Afrotropics; it occurs in south-western Arabia, Eritrea, Djibouti, Ethiopia, and Sudan (Fig. 1). Within this broad range, the bat is considered a polytypic species; up to seven subspecies have been defined and recognised (Andersen, 1918; Ellerman and Morrison-Scott, 1951; Koopman, 1994). Although several attempts to analyse the intraspecific structure of R. hipposideros have been made, this issue is still considered unresolved (see Burgin, 2019).

Fig. 1.

Map of the distribution range of the Rhinolophus hipposideros group (pale grey; after numerous sources) and the localities of origin and grouping of the examined samples (some symbols can denote more sites); full squares indicate the samples examined in both molecular genetic and morphological comparisons, open squares indicate the samples examined in the morphological comparison only, and open circles the samples used in molecular genetic analysis only. Grey dots denote isolated records out of the known distribution range

img-z2-6_269.jpg

Based on the body size, structure of the infraorbital region of the skull, and the presence and position of the small lower premolars, Andersen (1905, 1907, 1918) defined six subspecies in R. hipposideros, and these taxa have been listed as tentatively valid by various authors until present (see Ellerman and Morrison-Scott, 1951; Koopman, 1994; Horáček et al., 2000; Roer and Schober, 2001; Csorba et al., 2003; Simmons, 2005; Burgin, 2019). Csorba et al. (2003) and Burgin (2019) defined the distribution ranges of these subspecies as follows: R. h. hipposideros (André, 1797) [type locality (t.l.) Germany] in continental Europe north of the Alps, from the Netherlands to southern Ukraine; R. h. minutus (Montagu, 1808) [t.l. Wiltshire, England] in western Ireland and south-western Great Britain; R. h. minimus von Heuglin, 1861 [t.l. Kérén in den Bogosländern (= western Eritrea)] in Mediterranean Europe from Portugal to the Levant, including Sinai and the Mediterranean islands, and in western Arabia, southern Sudan, Eritrea, Djibouti, and central Ethiopia; R. h. midas Andersen, 1905 [t.l. Jask, Persian Gulf (southern Iran)] in western Asia, from the Caucasus, Transcaucasia, and northern Iraq, to southern Kazakhstan, western Kirghizstan, and Kashmir; R. h. majori Andersen, 1918 [t.l. Patrimonia, northern Corsica] in Corsica; and R. h. escalerae Andersen, 1918 [t.l. Mogador (= Essaouira), Ha-ha, Morocco] in the Mediterranean zone of north-western Africa.

With the exception of the latter two names, all of the above-mentioned forms were originally described as separate species that were, however, soon synonymised with R. hipposideros (Blasius, 1857; Peters, 1871; Dobson, 1876; Trouessart, 1879; Andersen 1904, 1918). Moreover, several authors have demonstrated the morphological inadequacies of these numerous subspecies and showed them to be difficult to identify, because the particular characters exhibit a variable occurrence in particular populations of the species (Miller, 1912; Grulich, 1949; Panouse, 1951; Saint Girons and Caubère, 1966; Felten et al., 1977; Palmeirim, 1990). Although such variability led to the description of additional taxa, both at the species and subspecies levels (currently invalid), see e.g., R. h. alpinus Koch, 1865 [t.l. the Alps], R. phasma Cabrera, 1904 [t.l. Madrid, Spain], R. h. vespa Laurent, 1937 [t.l. Korifla, Morocco], or R. moravicus Kostroň, 1943 [t.l. Ponikev and Kadeřín, Moravia (= Czech Republic)], the mosaic-like occurrence of traditional identification characters resulted in the abandonment of taxonomic division at small geographic scales (Corbet, 1978). The variability in various morphological characters was thus interpreted as an individual variation influenced by local environmental conditions rather than a result of phylogenetic separation (Saint Girons and Caubère, 1966; Palmeirim, 1990; Salinas-Ramos et al., 2021).

Felten et al. (1977) proposed the only revision of intraspecific taxonomy in R. hipposideros. Using an evaluation of the characters suggested by Andersen (1905, 1918) — body size, shape of rostrum, and size and position of premolars — Felten et al. (1977) delimited four population groups in the species and tentatively identified with the subspecies: R. h. hipposideros in Europe (including Corsica) and the Levant, R. h. minimus in northeastern Africa and Crete, R. h. midas in the Middle East from north-eastern Turkey to Afghanistan, and an unnamed form in the islands of the central Mediterranean (Sicily, Pantelleria) and in western Turkey [Felten et al. (1977) did not evaluate some populations, e.g., those of North Africa, British Isles, or West Turkestan]. This geographic division of morphotypes in R. hipposideros was revised only to a small extent and only to certain populations; the examined specimens of the Middle Eastern populations were found to fit the morphotypes defined by Felten et al. (1977) — see Benda et al. (2006) and Benda and Gaisler (2015). However, bats from Crete did not fit the morphological criteria that Felten et al. (1977) gave for R. h. minimus when a large set of samples was examined (Benda et al., 2009). In general, subsequent authors did not follow the conclusions that Felten et al. (1977) suggested regarding intraspecific relationships in R. hipposideros and their taxonomic arrangement.

Another type of evidence of the geographic variability in R. hipposideros was found and widely documented in karyotype (Zima et al., 1992; Zima, 2004; Volleth et al., 2013; Arslan and Zima, 2014; Kacprzyk et al., 2016); three chromosome races were described in R. hipposideros, (1) the populations of 2n = 54 from Ireland, Spain, Germany, and Switzerland; (2) 2n = 56 from Italy, Greece, Bulgaria, Czech Republic, and Slovakia; and (3) 2n = 58 from Jordan, Syria, Turkey, and Iran. These chromosome races thus seem to be geographically well-defined forms, with one living in the western part of Europe, another in the eastern part of Europe, and a third in the Middle East.

Molecular genetic analyses focused on intraspecific variation in R. hipposideros (Kůs, 2008; Dool et al., 2013; Shahabi et al., 2019) have shown — in both mitochondrial and nuclear markers — a split of the species into two main lineages, the western one comprising the Maghrebian and European populations (Maghreb, British Isles, central and southern Europe, Sardinia, Malta, and Crete) and the eastern one covering the Asian populations (Turkey, Cyprus, Levant, Iran, Tajikistan).

Burgin (2019) recently summarised the main message of the review presented above, although he did not propose a taxonomic synthesis revising the old intraspecific arrangement of R. hipposideros (suggested already by Andersen, 1905). However, the available data suggest this arrangement is untenable and the bat's intraspecific relationships need a profound revision. Thus, to identify the phylogenetic pattern in R. hipposideros, we carried out a morphological examination of a set of more than 270 museum specimens with the aim of defining the positions of particular populations from its whole distribution range. Simultaneously, we subjected a geographically representative subset of these specimens to a molecular genetic comparison. In addition, we compared the echolocation data from various parts of the species range. Results of these approaches are presented here, and we propose a revised view of the systematic relationships within the R. hipposideros group, including its taxonomic interpretation.

Nomeclatural Note

Although R. hipposideros ranks among the most common and most frequently mentioned bats of Europe and the western Palaearctic as well, the author and year of description of this species was confused for a long time. For almost 150 years, the creation of this name was attributed to J. M. Bechstein; initially to Bechstein (1801) (see e.g., Blasius, 1857; Kolenati, 1860; Koch, 1865; Peters, 1871; Dobson, 1876; Trouessart, 1879; Méhely, 1900; Cabrera, 1904), later to Bechstein (1800 [= 1799; see Benda and Mlíkovský, 2022]) (see e.g., Andersen, 1905; Miller, 1912; Ellerman and Morrison-Scott, 1951; Lay, 1967; Corbet, 1978; DeBlase, 1980; Qumsiyeh, 1985; Harrison and Bates, 1991; Horáček et al., 2000; Simmons, 2005; etc.). It was only recently that Tupinier (2001) and Kožurina (2006) pointed out that an older mention of this bat name was published by Borkhausen (1797), and the nomenclatural authority of this author over R. hipposideros has been nowadays accepted by numerous authors (see e.g. Benda et al., 2008, 2009, 2010, 2012, 2016; Kruskop, 2012; Lino et al., 2014; Benda and Gaisler, 2015; Downs et al., 2016; Burgin, 2019; Bendjeddou et al., 2022). However, Benda and Mlíkovský (2022) demonstrated that Borkhausen (1797) was not the oldest publication of the name hipposideros, while the available evidence shows that André (1797) published it earlier than Borkhausen (1797). Because the official publication dates for the purposes of zoological nomenclature are 19 April 1797 for André (1797), and 30 September 1797 for Borkhausen (1797), the former work takes priority over the latter and the author of the name hipposideros is André (1797).

Materials and Methods

Molecular Genetic Analysis

Sampling, amplification, and sequencing

In the molecular genetic analysis, we used muscle tissue samples of 92 specimens of R. hipposideros from the collection of the National Museum, Prague, Czech Republic (NMP) to extract DNA (Fig. 1 and  Supplementary Table S1A (01-AC-24-2-p-269-298-Supplement.pdf)). The genomic DNA was extracted from the alcohol-preserved tissue samples using Geneaid Genomic DNA Mini Kit. We targeted one mitochondrial marker (mtDNA), including 1,128 bp of the cytochrome b gene (Cyt-b) and five nuclear markers (nDNA), consisting of 536 bp of acyl-coenzyme A oxidase 2 intron (ACOX), 616 bp of biglycan intron (BGN), 741 bp of COP9 signalsom subunit 7A intron (COPS), 480 bp of the rogdi atypical leucine zipper (ROGDI), and 521 bp of the signal transducer and activator of transcription 5A intron (STAT). We sequenced both strands for all sequences. We used primers that have been specifically designed for the order Chiroptera and provided good amplification in previous studies (see, e.g., Puechmaille et al., 2011; Salicini et al., 2011; Thong et al., 2012; Dool et al., 2016).

We supplemented this dataset with 155 Cyt-b sequences from previous studies (Ibáñez et al., 2006; Li et al., 2006; García-Mudarra et al., 2009; Çoraman et al., 2013; Dool et al., 2013, 2016). As a multiple outgroup, we added 38 GenBank sequences of 28 other Rhinolophus species (Dool et al., 2016; Taylor et al., 2018) and sequences of three Hipposideros species from the sister family Hipposideridae (for details see  Supplementary Table S1A (01-AC-24-2-p-269-298-Supplement.pdf)). The largest possible set of shorter sequences of the Cyt-b gene of R. hipposiderosSupplementary Table S1B (01-AC-24-2-p-269-298-Supplement.pdf)) from GenBank was used for the test of geographic grouping of particular mtDNA haplotypes. For the primer names, their sequences, and annealing temperatures, see  Supplementary Table S2 (01-AC-24-2-p-269-298-Supplement.pdf). The PCR products were Sanger-sequenced from both sides using the PCR primers by Macrogen, Inc. (Amsterdam, the Netherlands).

Phylogenetic reconstruction

Sequences were edited and aligned using the MAFFT plug-in (Katoh and Standley, 2013) in Geneious 11.0.5 ( https://www.geneious.com), subsequently manually edited and trimmed using Gblocks (Castresana, 2000). Heterozygous positions in the nDNA markers were coded with IUPAC codes and ambiguous positions or missing data were coded with ‘N’. Indels were treated as gaps. Sequences of protein-coding markers were translated to amino acids to check for the presence of stop codons, which would indicate that pseudogenes have been amplified. Alleles of nuclear markers were estimated using PHASE (Flot, 2010) with the probability threshold set to 0.7. The two final multilocus datasets were made according to the mode of inheritance of the markers, mitochondrial and nuclear datasets. The mitochondrial dataset contained Cyt-b sequences of a total length of 1,128 bp. The nuclear dataset contained ACOX, BGN, COPS, ROGDI, and STAT sequences of a total length of 2,894 bp. The latter dataset was partitioned by gene.

Phylogenetic analyses of both datasets were run using Bayesian inference (BI) and maximum likelihood (ML). The appropriate nucleotide substitution model for each partition was selected based on the Bayesian information criterion (BIC) ModelFinder ( Supplementary Table S3 (01-AC-24-2-p-269-298-Supplement.pdf)Kalyaanamoorthy et al., 2017). We used MrBayes v3.2.6 (Ronquist and Huelsenbeck, 2003) to run the BI analysis. Appropriate substitution models were specified for each partition and all parameters were unlinked across partitions. We ran two independent runs for 20 million generations with trees sampled every 1,000 generations. All other parameters were set to default. Stationarity and convergence of the runs were inspected in Tracer v1.6 (Rambaut et al., 2014) and the values of the average standard deviations of the split frequencies were lower than 0.01. The burn-in fraction was left as the default at 25% of sampled trees. Thus, from the 20,000 produced trees, 5,000 were discarded. A majority-rule consensus tree was produced from the post-burnin trees with posterior probability (PP) values embedded. The BI analyses were run through CIPRES Science Gateway (Miller et al., 2010). Then, we inferred the maximum-likelihood tree using the partition model in IQ-TREE (Nguyen et al., 2015; Chernomor et al., 2016). Searching for the best-scoring ML was performed by ultrafast bootstrap (UFBoot — Hoang et al., 2018) with 1,000 bootstrap and 1,000 topology replicates. To verify robustness of the ML tree the branch supports were evaluated using SH-like approximate likelihood ratio test (SH-aLRT — Guindon et al., 2010) and a Bayesian-like transformation of aLRT (aBayes — Anisimova et al., 2011). SH-aLRT was performed with 1,000 replications. aBayes branch support was used instead Bayesian posterior probabilities because aBayes is more conservative, more robust to model violation and moreover exhibits the more confident resolution (Anisimova et al., 2011). The ML, SH-aLRT and aBayes analysis were run on IQtree web server (Trifinopoulos et al., 2016). To see whether the single nuclear markers show the same or different topology we prepared the phylogenetic trees for each nuclear marker.

Species delimitation and divergence time estimation

For the species delimitation and molecular dating analyses, we used only pruned nuclear dataset employed in phylogenetic analyses constituted from phased sequences of ACOX, BGN, COPS and STAT. For R. hipposideros, we used sequences of only two individuals, one from each diverged lineage (see below) from Cyprus and Oman. Furthermore, the data set was truncated by species represented by less than three markers, and therefore the sequences of R. landeri and R. pearsonii were omitted.

The species delimitation was conducted by Bayesian phylogenetics and phylogeography (BPP v3; Rannala and Yang, 2003; Yang and Rannala, 2010). This analysis was carried out to evaluate the phylogenetic species boundaries. The species tree topology, which was reconstructed using only nuclear loci (see above), was used as a fixed guide tree (algorithm A10 — Rannala and Yang, 2003; Yang and Rannala, 2010). We replicated twice the runs for each of four different combinations of priors on divergence depth and effective population sizes (τ and θ, respectively — see Table 1 in Demos et al., 2019), as the probability of delimitation by BPP is sensitive to these two parameters (Leaché and Fujita, 2010; Yang and Rannala, 2010). Each replicate was conducted with either the reversible-jump Markov chain Monte Carlo algorithm 0 (with parameter e = 1) or 1 (with parameters a = 2, m = 1 — Yang and Rannala, 2010). All eight BPP analyses were then run with the default settings. Lineages were considered statistically supported when the generated delimitation posterior probabilities (PP) exceeded 0.95 under all four prior combinations.

The divergence time estimation was set up in BEAUti and run in BEAST v1.8.4. We followed the settings from Dool et al. (2016) and used strict molecular clocks and Yule speciation process (Yule, 1925; Gernhard, 2008) for all genes. The substitution model was taken from phylogenetic reconstructions (see above). As a calibration point, we employed the age of the root of the family Rhinolophidae which was estimated at 37 Ma (Stoffberg et al., 2010). For an alternative divergence time reconstruction, we also used a family root age of 16.92 Ma (Foley et al., 2015). We used a lognormal prior distribution for this calibration point. BEAST was run three times for 20 million generations and parameters and trees were saved every 1,000 generation. Tracer v1.6 was used to confirm adequate mixing of the MCMC chains and acceptable effective sample sizes (ESS > 200). LogCombiner was used for burn-in (25%) and merging trees files, TreeAnnotator was used for identifying the maximum clade credibility tree. All analyses were run through CIPRES Science Gateway (Miller et al., 2010).

Uncorrected p-distances between haplotypes were calculated for the Cyt-b in MEGA11 (Tamura et al., 2021). The bootstrap was performed with 1,000 replications.

Morphometric Comparison

For the comparative morphometric analysis and for the description of morphological trends in particular populations, we used cranial and dental measurements and the forearm length (LAt) as a standardised dimension referring to the body size. The skulls and teeth were measured using mechanical and optical callipers with accuracy to 0.02 mm and 0.01 mm, respectively; horizontal dental dimensions were taken on cingulum margins of teeth. The examined museum materials are given in Appendix I (see also Fig. 1). We evaluated 18 cranial and 19 dental dimensions (i.e., plain dimensions) in each skull (see the measurements taken below); the skull and tooth shapes were described with the help of relative dimensions (indices) calculated from the plain dimensions; nine cranial and 17 dental indices were used (see  Supplementary Tables S5 and S (01-AC-24-2-p-269-298-Supplement.pdf)6). In accordance with Felten's et al. (1977) findings, sexual dimorphism was not considered in the morphometric comparisons.

For the statistical evaluation and definition of trends in morphological characters, the examined museum specimens were grouped into six sample sets, with respect to the geographic origin of the samples and to the geographic separation of lineages shown by the molecular genetic analysis that preceded the morphological comparison. The compared sample sets were defined as follows (see Tables 2 and 3): Central Europe (CEU) — 55 samples from the Czech Republic and Slovakia; West Mediterranean (WMT) — 106 samples from Morocco, Algeria, Croatia, Serbia, Albania, Kosovo, North Macedonia, Bulgaria, and Greece (including Crete); East Mediterranean (EMT) — 83 samples from Syria, Crimea (Ukraine), Rhodes (Greece), Cyprus, Lebanon, Jordan, Turkey; Central Asia (CAS) — 25 samples from Iran, Azerbaijan, Turkmenistan, Uzbekistan, Kirghizstan, Tajikistan, and Afghanistan; Oman (OMA) — four samples from north-eastern Oman; north-eastern Africa (NEA) — two samples from Ethiopia and Sudan. Two type specimens examined (R. midas Andersen, 1905 and R. h. escalerae Andersen, 1918) were evaluated separately off the sets to avoid affecting the statistical results.

Statistical analyses were performed using Statistica 6.0 software. In the cluster analysis, the unweighted pair group method with arithmetic mean was employed (UPGMA; Euclidean distances); the analysis was used to calculate differences between the mean values of morphometric traits among the particular sets of samples, and it was employed separately for 27 plain and relative dimensions of the skull and for 36 plain and relative dimensions of the teeth, respectively. Stepwise discriminant function analysis was performed as a test of importance of particular dimensions and their indices for geographic variation; statistically significant parameters most affecting morphological variation were selected and employed in a subsequent canonical analysis that was used to test grouping or separation of population sample sets of similar or different morphotypes, respectively. Statistical significance of differences in skull measurements between groups were assessed using ANOVA (one-way analysis of variance).

The following measurements were taken: (1) External dimension — LAt = forearm length; (2) Cranial dimensions — LCr = greatest length of skull incl. praemaxillae; LOc = occipitocanine length; LCc = condylocanine length; LaZ = zygomatic width; LaI = width of interorbital constriction; LaInf = rostral width between infraorbital foramens; LaNc = neurocranium width; LaM = mastoidal width of skull; ANc = neurocranium height; LBT = largest horizontal length of tympanic bulla; CC = rostral width between canines (incl.); M3M3 = rostral width between third upper molars (incl.); CM3 = length of upper tooth-row between canine and third molar (incl.); LMd = condylar length of mandible; ACo = height of coronoid process; CM3 = length of lower tooth-row between canine and third molar (incl.); (3) Dental dimensions, upper dentition — M1M3 = length of tooth-row between first and third molars (incl.); LCs = largest mesio-distal length of canine; LaCs = largest palato-labial width of canine; LP2 = largest mesio-distal length of first premolar; LaP2 = largest labio-palatal width of first premolar; LP41 = largest mesio-distal length of large premolar on the labial cingulum; LP42 = smallest mesio-distal length of large premolar taken over the talon constriction; LP43 = mesio-distal length of large premolar on palatal cingulum (largest dimension taken over the palato-mesial to palato-distal points of the talon); LaP4 = largest palato-labial width of large premolar taken over the mesio-labial and palato-distal cingulum margins; LM1 = largest mesio-distal length of first molar taken over parastyle and metastyle; LaM1 = largest palato-labial width of first molar taken over parastyle and palato-distal part of talon; LM3 = largest mesio-distal length of third molar; LaM3 = largest palato-labial width of third molar (taken over parastyle and palatal cingulum); (4) Dental dimensions, lower dentition — M1M3 = length of tooth-row between first and third molars (incl.); LCi = largest mesio-distal length of canine; LP2 = largest mesio-distal length of first premolar; LaP2 = largest labio-lingual width of first premolar; LP3 = largest mesio-distal length of second (small) premolar; LP4 = largest mesio-distal length of last premolar; LaP4 = largest labio-lingual width of last premolar; LMi = largest mesio-distal length of first molar taken over paraconid and hypoconulid. Other abbreviations included: n = number of samples; x̄ = mean; min, max = range margins; SD = standard deviation.

Echolocation Call Recordings and Analysis

In the Rhinolophus bats, the constant frequency component represents a dominant part of the echolocation call in the search phase. This characteristic has maximum energy and thus makes it acceptable to analyse calls from hand-held and flying bats, while avoiding pseudoreplication during the recording of flying bats, respectively. For the echolocation call analysis in the R. hipposideros group, we made the acoustic recordings using a portable ultrasound detector D-240x (Pettersson Elektronik AB, Uppsala, Sweden) set on time-expansion mode connected to Edirol R-09HR recorder (Roland Corp., Japan) and an ultrasound detector Batlogger M (Elekon AG, Switzerland). The analysed bat calls were recorded in free flight under natural conditions, usually near the sites where the bats were also mist-netted. Additionally, some echolocation call sequences were recorded when handling the bat in a resting position or hand-releasing the bat.

The recordings were analysed with BatExplorer 2.1.7.0 software (Elekon AG, Switzerland) to evaluate oscillograms, power spectra, and spectrograms. For each echolocation call, the following parameters were measured: pulse duration (PDUR), start frequency (SF), end frequency (EF), frequency of maximum energy (FMAXE) and inter-pulse interval (IPI, the time between two consecutive calls). In most cases, we used only high-quality recordings for analyses, in which all or most of the basic characters were measurable, and only the search phase calls were measured.

For comparison of the geographic variability, mostly published data were used (see Table 5). Original data were obtained from Slovakia, Tajikistan, Saudi Arabia, and Oman; the calls were recorded at the following sites: at the Aksamitka Cave, Slovakia (49°23′N, 20°27′E), 31 August 2015, several individuals, rec. M. Ceľuch; at a small cave near Zingrogh, Tajikistan (38°27′N, 70°49′E), 12 May 2016, one ind., rec. M. Uhrin; at the Umm Jirsan Cave, Saudi Arabia (25°35′N, 39°45′E), 26 October 2022, several inds., rec. M. Uhrin; at a water reservoir near Al Khutaymi, Oman (23°06′N, 57°33′E), 27 March 2011, one ind., rec. M. Uhrin; in Wadi Qatam, Oman (23°05′N, 57°38′E), 31 October 2019, several inds., rec. P. Benda; in a small oasis near Misfah, Oman (23°14′N, 57°08′E), 9 April 2011, one ind., rec. M. Uhrin; and at a pool near Tayma, Oman (22°31′N, 59°20′E), 3 April 2011, one ind., rec. M. Uhrin.

Results

Molecular Genetic Analysis

The resulting Cyt-b dataset comprised 81 sequences which were pruned to 54 unique haplotypes. The nuclear dataset comprised 47 ACOX, 63 BGN, 44 COPS, 11 ROGDI, and 47 STAT sequences that were pruned to 46 haplotypes. For other Rhinolophus species, we added 129 sequences from GenBank in total. The Cyt-b sequences contained 403 parsimony informative positions (35.73% of total length) and this marker showed a much larger genetic differentiation within Rhinolophus species than the nuclear markers (due to the faster mutation rate). The amount of parsimony informative positions in concatenated nuclear dataset was 386, i.e. 13.33% of its total length (for substitution models of mitochondrial and nuclear trees see  Supplementary Table S3 (01-AC-24-2-p-269-298-Supplement.pdf)).

The ML and BI tree of the nuclear dataset showed slightly different topologies, nonetheless, the different nodes had a low branch support. We showed the ML tree (Fig. 2). The genus Rhinolophus was divided into four well supported clades. Rhinolophus hipposideros formed a separate clade, however, its exact position remained unclear due to the low branch support of deep nodes. Other groups were: (1) pusillus group including the species R. shameli, R. pearsonii, and R. pusillus; (2) trifoliatus group including R. trifoliatus and R. luctus; and (3) Afro-Palaearctic clade that includes the species groups euryale, fumigatus, ferrumequinum, capensis, and landeri.

Fig. 2.

Maximum likelihood tree of reconstructed phylogenetic relationships of the Rhinolophus hipposideros group with selected species of the families Rhinolophidae and Hipposideridae based on the nuclear (left) and mitochondrial (right) datasets, respectively. Branch support values are shown by pie charts on the nodes

img-z7-5_269.jpg

The phylogenetic trees obtained by both ML and BI analyses of the Cyt-b dataset showed slightly different topologies. The ML tree was fully resolved and had a higher branch support than the BI tree, therefore we showed the ML tree (Fig. 2). The tree topology of the genus Rhinolophus was different than the topology of the nuclear tree. Within Rhinolophus, R. hipposideros formed a separate branch with an uncertain position. Another separate branch led to the trifoliatus group including only R. luctus. The rest of all Rhinolophus species formed a third branch within the genus tree supported by SH-aLRT and aBayes. This clade comprised African, European, and Asian species including the supported species groups ferrumequinum, fumigatus, maclaudi, capensis, and landeri. Nevertheless, the relationships among these groups and other ungrouped species were not satisfactorily resolved and neither were the relationships between three major clades.

Intraspecifically, R. hipposideros split into two lineages in both nuclear and mitochondrial trees. The first lineage ranged from Morocco and Ireland through Central Europe, the Balkans, Levant, and Iran to Tajikistan and Ethiopia; the second lineage comprised samples from north-eastern Oman. In the nuclear tree, the first lineage was not internally branched, and therefore the lineage was genetically uniform through almost the whole R. hipposideros range, except Oman which formed the second lineage. In the Cyt-b tree, the first lineage was formed by samples from the majority of the R. hipposideros distribution range and was further divided into four well-supported sublineages: (1) Afro-European, comprising samples from the Maghreb (Morocco), Central Europe (Slovakia), and the Balkans (Albania, Bulgaria, Greece); (2) Ponto-Levantine, with the samples from Crimea, Rhodes, and the Levant (Cyprus, Syria, Lebanon, Jordan); (3) Central Asian, with the samples from Iran and Tajikistan; and (4) Ethiopian, which includes a single sample from Ethiopia. The relationships among the sublineages were not resolved due to the low branch support. The uncorrected p-distances on Cyt-b between the two lineages were 8.93–10.75% and between the sublineages 2.30–7.02% (Table 1). The resolution of the nuclear gene trees was in accordance with the genetic variation of each nuclear marker ( Supplementary Figs. S2–S6 (01-AC-24-2-p-269-298-Supplement.pdf)). However, the basic split of R. hipposideros into two lineages was evident in all gene trees for which we obtained sequences from both lineages.

Table 1.

Percentage values of uncorrected genetic p-distances of Cyt-b among mitochondrial subgroups (lineage/sublineage) of the Rhinolophus hipposideros group (below the diagonal). The diagonal corresponds to the within-group genetic divergence estimated for Cyt-b in each subgroup

img-z8-6_269.gif

For the R. hipposideros only mitochondrial tree ( Supplementary Fig. S1 (01-AC-24-2-p-269-298-Supplement.pdf)), we added 155 Cyt-b sequences from GenBank to make a dataset of 213 sequences with the total length of 1,103 bp ( Supplementary Table S1B (01-AC-24-2-p-269-298-Supplement.pdf)). In this tree, five basic lineages in R. hipposideros were recovered. It corresponded to the topology with the above division based on the whole Cyt-b gene, and it covered almost a complete distribution range of the species; viz. (1) the Afro-European lineage, comprising sequences from the Maghreb (Morocco, Tunisia), Mediterranean Europe (Spain, Italy, France, Slovenia, Albania, Bulgaria, Greece, European Turkey), Mediterranean islands (Malta, Crete), British Isles (Ireland, Great Britain), and Central Europe (Austria, Slovakia, Romania); (2) Ponto-Levantine lineage, composed of the sequences from the Levant (Rhodes, western Anatolia, Cyprus, Syria, Lebanon, Israel, Jordan) and Crimea; (3) Eastern lineage comprising the sequences from the eastern part of the Middle East (eastern Anatolia, Iran) and West Turkestan (Tajikistan); (4) the Ethiopian lineage comprising one sequence from northern Ethiopia; and the last and most distant (5) Omani lineage from the sequences from north-eastern Oman. All five lineages had a high branch support (0.99–1.00 posterior probability [PP] and 97–100 bootstrap percentage [BP]). However, the relationships between the lineages did not always show high support, only the sister position of lineages 2 and 3 had marginal to moderate high support (0.82 PP and 93 BP), and the ML analysis supported the crown position of lineages 1–4 (98 BP). The uncorrected p-distances within lineages were 0–3.42%, between sublineages 2.30–7.02%. The Omani lineage differed from other lineages with the distances of 8.93–11.40%.

Our results of the Bayesian phylogenetics and phylogeography (BPP) analyses demonstrated the delimitation probabilities of the replicated runs being affected by the prior choice of parameters. It was especially apparent when a large effective population size was chosen in our pruned dataset ( Supplementary Table S4 (01-AC-24-2-p-269-298-Supplement.pdf)). Nevertheless, all the results for R. hipposideros and its populations had PP ≥ 0.95. It means that two clades, one from Oman and another from the rest of the distribution range, were strongly delimited within this lineage.

The topology of the calibrated tree (Fig. 3) showed the same four clades of the genus Rhinolophus as displayed by the topology of the nuclear ML/BI tree, however, their positions differed. The basal split occurred 37.8 Ma (95% highest posterior density [HPD]: 37.1–39.0 Ma) and divided the trifoliatus group from the rest of Rhinolophus species. A second split took place 32.0 Ma (95% HPD: 27.4–36.6 Ma) between the Indomalayan group and the Afro-Palaearctic group including R. hipposideros. Finally, the Afro-Palaearctic group diverged from R. hipposideros 29.4 Ma (95% HPD: 24.8–34.4 Ma). In the tree, all the nodes were statistically supported except three: between the Indomalayan group and the Afro-Palaearctic group including R. hipposideros; between the groups euryale and landeri (including only R. alcyone); and between R. fumigatus and R. hildebrandtii. In the R. hipposideros clade, two lineages used in our study split 7.1 Ma (95% HPD: 4.3–10.0 Ma). For the reconstruction based on a younger root calibration see  Supplementary Fig. S7 (01-AC-24-2-p-269-298-Supplement.pdf). The topology of both reconstructions remained identical, however, the splits of each group estimated in the alternative reconstruction occurred much later (16.7 Ma [16.1–17.4 Ma], 14.3 Ma [12.3–16.4 Ma], 13.2 Ma [11.0–15.3 Ma], and 3.2 Ma [2.0–4.9 Ma], respectively).

Fig. 3.

Chronogram of the family Rhinolophidae based on a Bayesian inference of the nuclear dataset (according to the model by Stoffberg et al., 2010). The numbers at nodes show mean divergence time estimates (Ma) and horizontal boxes 95% highest posterior density intervals of these estimates. The asterisk (*) indicates nodes with low branch support, the rest of the nodes were supported (PP ≥ 0.95)

img-z9-7_269.jpg

Morphometric Comparison

In accordance with the geographic separation of lineages in the examined mitochondrial markers (see above) and the origin of the comparative material (Appendix I), all of examined material of R. hipposideros was sorted into six sample sets (Tables 24). The comparison of morphometric characters of the population sets documented a remarkable variation in the body, skull, and tooth sizes as well as in the skull and tooth shapes. In most dimensions, both in their absolute and relative values, the dimension ranges in particular sets overlapped with or exceeded the ranges of other sets. However, metric trends in the population sample sets were easily detectable from the comparison of the basic statistical values (Tables 2 and 3).

Regarding body size, two basic groups could be delimited among the examined samples, the large bats (x̄ LAt > 39 mm) from Central Europe and Central Asia, and the small bats (x̄ LAt < 38 mm) from the Mediterranean and Oman; a single sample from north-eastern Africa is medium-sized in this respect (LAt 38.7 mm).

Table 2.

External and cranial dimensions of the examined sample sets of the Rhinolophus hipposideros group (* — after DeBlase, 1980); midas, escalerae = dimensions of the respective type specimens; for the sample set delimitations and dimension abbreviations see Materials and Methods. Mean values shown in bold

img-z10-2_269.gif

Large skull size (x̄ LCc > 14.0 mm) was found in the bats from Central Europe; a small skull (x̄ LCc < 13.5 mm) was seen in the samples from the East Mediterranean, Oman, and north-eastern Africa; and a medium-sized skull (x̄ LCc 13.5–13.8 mm) was observed in the bats from the West Mediterranean and Central Asia (Fig. 4). An absolutely and relatively wide skull (x̄ LaZ > 7.4 mm; x̄ LaZ/LCc > 0.545) was observed in the bats from Central Europe, West Mediterranean, and Central Asia, whereas an absolutely and relatively narrow skull (x̄ LaZ < 7.4 mm; x̄ LaZ/LCc < 0.545) was seen in the bats from Oman and an absolutely narrow but relatively wide skull (x̄ LaZ < 7.4 mm; x̄ LaZ/LCc > 0.545) was found in the bats from the East Mediterranean and north-eastern Africa.

An absolutely and relatively wide braincase (x̄ LaNc > 6.5 mm; x̄ LaNc/LCc > 0.475) was found in the samples from the West Mediterranean and north-eastern Africa, whereas an absolutely and relatively narrow braincase (x̄ LaNc < 6.5 mm; x̄ LaNc/LCc < 0.475) was observed in the bats from Central Asia and Oman; an absolutely wide but relatively narrow braincase (x̄ LaNc > 6.5 mm; x̄ LaNc/LCc < 0.475) was observed in the bats from Central Europe; and an absolutely narrow but relatively wide braincase (x̄ LaNc < 6.5 mm; x̄ LaNc/LCc > 0.475) was seen in the bats from the East Mediterranean. Two shape types were found concerning the absolute and relative height of braincase; the bats from Central Europe, the Mediterranean, and Central Asia had a high braincase (x̄ ANc > 4.5 mm; x̄ Nc/LCc > 0.335), and the bats from north-eastern Africa and Oman had a low braincase (x̄ ANc < 4.5 mm; x̄ ANc/LCc < 0.335). An absolutely and relatively large tympanic bulla (LBT > 2.9 mm; LBT/LCc > 0.2) was observed in the bats from Oman, whereas a small bulla (LBT < 2.8 mm; LBT/LCc < 0.2) was seen in the bats from all other geographic sample sets (Fig. 5).

Table 3.

Dental dimensions of the examined sample sets of the Rhinolophus hipposideros group; midas = dimensions of the respective type specimen; for the sample set delimitations and dimension abbreviations see Materials and Methods. Mean values shown in bold

img-z11-2_269.gif

The rostral part of the skull was absolutely and relatively long (x̄ CM3 > 5.4 mm; x̄ CM3/LCc > 0.395) in the bats from Central Asia and Oman (Fig. 5); an absolutely long but relatively short rostrum (x̄ CM3 > 5.4 mm; x̄ CM3/LCc < 0.385) was observed in the bats from Central Europe, whereas an absolutely short but relatively long rostrum (x̄ CM3 < 5.4 mm; x̄ CM3/LCc > 0.385) was found in the bats from the Mediterranean and north-eastern Africa. A relatively narrow rostrum (x̄ LaInf/LCc < 0.262) was found in the bats from Central Europe and the Mediterranean; a relatively very wide rostrum (x̄ LaInf/LCc > 0.264) was observed in the bats from Oman; and a relatively medium-sized rostrum width (x̄ LaInf/LCc 0.262–0.264) was seen in the bats from Central Asia and north-eastern Africa.

Table 4.

Descriptive features of morphotypes of particular sample of the Rhinolophus hipposideros group (average states of absolute metric values are defined)

img-z12-2_269.gif

Fig. 4.

Bivariate plot of skull dimensions of the examined samples of the Rhinolophus hipposideros group: condylocanine length of skull (LCc) against length of the upper tooth-row (CM3); values in mm

img-z12-6_269.jpg

Although the tooth metric characters largely followed the size trends in the skulls, certain shape variability and size trends were detectable in particular teeth. The largest upper canines (Cs; x̄ LaCs > 0.90 mm) were observed in the bats from Oman, the smallest (x̄ LaCs < 0.81 mm) were seen in the bats from the West Mediterranean and north-eastern Africa, and the upper canines (x̄ LaCs 0.82–0.85 mm) in the bats from Central Europe, the East Mediterranean, and Central Asia were medium size (Fig. 6). The small upper premolar (P2) was found to be large (LP2 > 0.52 mm; x̄ LP2×LaP2 > 0.25 mm2) in the bats from Central Europe, small (x̄ LP2 < 0.42 mm; x̄ LP2×LaP2 < 0.20 mm2) in the bats from Oman, and medium-sized (x̄ LP2 0.46–0.50 mm; x̄ LP2×LaP2 0.20–0.25 mm2) in the bats from the Mediterranean, Central Asia, and north-eastern Africa. The large upper premolar (P4) was found to be large (x̄ LP41 > 0.98 mm) in the bats from Central Europe and Oman, small (x̄ LP41 < 0.95 mm) in the bats from the East Mediterranean and northeastern Africa, and medium-sized (x̄ LP41 0.95–0.98 mm) in the bats from the West Mediterranean and Central Asia; P4 was relatively wide (x̄ LaP4/LP41 > 1.6) in the bats from the East Mediterranean and Central Asia, and relatively narrow (x̄ LaP4/LP41 < 1.6) in the bats from the remaining four sample sets; it was relatively long in its medial portion (i.e., with a smallest posterior concavity in the distal margin of talon; x̄ LP42/LaP4 > 0.36) in the bats from north-eastern Africa, short (x̄ LP42/LaP4 < 0.34) in the bats from the East Mediterranean and Central Asia, and medium length (x̄ LP42/LaP4 0.34–035) in the bats from Central Europe, the West Mediterranean, and Oman.

Fig. 5.

Bivariate plot of skull dimensions of the examined samples of the Rhinolophus hipposideros group: relative length of rostrum (CM3/LCc) against relative horizontal length of tympanic bulla (LBT/LCc)

img-z13-4_269.jpg

The first upper molar (M1) was found to be large (x̄ LM1 > 1.4 mm; x̄ LM1×LaM1 > 2.7 mm2) in the bats from Central Europe and Central Asia, small (x̄ LM1 < 1.3 mm; x̄ LM1×LaM1 < 2.5 mm2) in the bats from north-eastern Africa, and medium-sized (x̄ LM1 1.35–1.38 mm; x̄ LM1×LaM1 2.5–2.7 mm2) in the bats from the Mediterranean and Oman; M1 was relatively wide (x̄ LaM1/LM1 > 1.4) in the bats from Central Europe, Central Asia, and northeastern Africa, relatively narrow (x̄ LaM1/LM1 1.36) in the samples from Oman, and medium width (x̄ LaM1/LM1 1.39–1.40) in the bats from the Mediterranean. The third upper molar (M3) was large (x̄ LM3 > 1.1 mm; x̄ LM3×LaM3 > 1.5 mm2) in the bats from Central Asia and Oman and small (x̄ LM3 < 1.1 mm; x̄ LM3×LaM3 < 1.5 mm2) in the bats from Central Europe, the Mediterranean, and north-eastern Africa; M3 was relatively wide (x̄ LaM3/LM3 > 1.4) in the bats from Central Europe, relatively narrow (x̄ LaM3/LM3 <1.3) in the samples from the West Mediterranean, Central Asia, and Oman, and medium width (x̄ LaM3/LM3 1.3–1.4) in the bats from the East Mediterranean and north-eastern Africa. In relation to M1, M3 was found to be large (x̄ LM3×LaM3/LM1×LaM1 > 0.6) in the bats from Oman, small (x̄ LM3×LaM3/LM1 ×LaM1 < 0.5) in the bats from Central Europe, and medium-sized (x̄ LM3×LaM3/LM1×LaM1 0.5–0.6) in the bats from the remaining four sample sets (Fig. 6).

The lower canine (Ci) was observed to be large (x̄ LCi > 0.7 mm) in the bats from Central Europe and Central Asia and small (x̄ LCi < 0.6 mm) in the remaining four sample sets; in relation to the first lower molar (Mi), the Ci was relatively large (x̄ LCi/LMi > 0.5) in the bats from Central Europe, the West Mediterranean, Central Asia, and northeastern Africa, and relatively small (x̄ LCi/LMi < 0.5 mm) in the bats from the East Mediterranean and Oman. The first lower premolar (P2) was large (x̄ LaP2 > 0.52 mm; x̄ LP2×LaP2 > 0.3 mm2) in the bats from Central Europe and the Mediterranean, small (x̄ LaP2 < 0.48 mm; x̄ LP2×LaP2 < 0.27 mm2) in the bats from Oman, and medium-sized (x̄ LaP2 0.48–0.52 mm; x̄ LP2×LaP2 0.28–0.30 mm2) in the bats from Central Asia and north-eastern Africa. In relation to the last lower premolar (P4), P2 was very small (x̄ LP2×LaP2/LP4×LaP4 < 0.53) in the bats from Oman. In all other sample sets, this tooth was found to be large or very large (x̄ LP2×LaP2/LP4 ×LaP4 > 0.58). The last lower premolar (P4) was large (x̄ LaP4 > 0.75 mm; x̄ LP4×LaP4 > 0.48 mm2) in the bats from Central Europe, Central Asia, and Oman and small (x̄ LaP4 < 0.75 mm; x̄ LP4×LaP4 > 0.48 mm2) in the bats from the Mediterranean and north-eastern Africa. The small lower premolar (P3) was found to be large (x̄ LP3 > 0.23 mm) in the Omani samples, but small (x̄ LP3 < 0.23 mm) in all other sample sets. The first lower molar (Mi) and the lower molar-row were large (x̄ LMi > 1.4 mm; (x̄ M1M3 > 3.95 mm) in the bats from Central Asia and Oman and small (x̄ LMi <1.4 mm; x̄ M1M3 < 3.95 mm) in the four remaining sample sets.

Fig. 6.

Bivariate plot of tooth dimensions of the examined samples of the Rhinolophus hipposideros group: relative crown size of third upper molar (LM3×LaM3/LM1×LaM1) against crown size of upper canine (LCs×LaCs)

img-z14-5_269.jpg

In summary, the comparison demonstrated certain characters were unique in four of the six examined sample sets (Tables 24 and  Supplementary Tables S5–S7 (01-AC-24-2-p-269-298-Supplement.pdf); see Appendix II for a review of the state conditions of evaluated metric characters in the particular sample sets). The Omani sample set was shown to be the most distinct among all bats (including the statistical comparison —  Supplementary Table S7 (01-AC-24-2-p-269-298-Supplement.pdf)). Within the matrix of 25 metric characters evaluated above, the Omani bats showed in ten characters a state to be unique in relation to all other sample sets: an absolutely and relatively very narrow skull with a relatively very wide rostrum and an absolutely and relatively very large tympanic bulla, very large upper canine, very small first (small) upper premolar (P2), relatively very narrow (palato-labially short) first upper molar (M1) and a relatively very large third upper molar (M3), an absolutely as well as relatively very small first lower premolar (P2), and a very large small-lower premolar (P3). Six unique characters were found in the samples from Central Europe: a very large skull size with an absolutely wide but relatively narrow braincase and an absolutely long but relatively short rostrum, a very large first (small) upper premolar (P2), and a very small and relatively very wide third upper molar (M3). Only two unique characters were documented in the sample set from north-eastern Africa (large upper premolar (P4) being relatively narrow as a whole but relatively wide in its medial portion, and a very small first upper molar, M1), and one was documented in the East Mediterranean set (an absolutely narrow but relatively wide braincase). No unique character among the evaluated metric traits was observed in the bats from the West Mediterranean and Central Asia.

The examined skulls of holotype specimens of two names of the R. hipposideros group (R. midas and R. h. escalerae) were compared with the above-defined morphotypes (Tables 2 and 3). The type of escalerae from western Morocco conforms in most characters to the Mediterranean populations, namely in the skull size and shape (i.e., the skull width, absolute and relative length and width of the rostrum, absolute and relative width and height of the braincase, and mandible length). The type of midas from southern Iran conforms in most respects to the bats from Oman. Similarities were found in all types of characters, in the skull size and shape and in the sizes and shapes of teeth. The type skull of R. midas is large with large tooth rows, although it is absolutely and relatively narrow; the braincase is very low; the rostrum is rather wide; and the tympanic bullae are very large. Although the Omani skulls are slightly smaller than the type skull of midas is in absolute dimensions, they well agree in the relative dimensions as well as in the absolute and relative size of tympanic bullae (Figs. 4 and 5). Even more pronounced than in the skull dimensions, the similarity of the midas type and Omani bats is apparent in the tooth dimensions. These bats are very similar in the extremely small size of the small upper premolar (P2), large absolute and mainly relative size of the last upper molar (M3), small relative and absolute size of the lower canine (Ci), small absolute and relative size of the first lower premolar (P2), and large absolute size of the smallest lower premolar (P3).

The separate position of the Omani samples plus the type specimen of midas in relation to all other sample sets is also illustrated by the results of the UPGMA cluster analysis (Fig. 7). Both the results calculated from the skull and tooth data showed similar positions of the Omani set together with midas positions in separate clusters, whereas the remaining sample sets from Central Europe, the Mediterranean, Central Asia, and north-eastern Africa in other clusters showed variable inner topology of particular sets. The results of a canonical analysis calculated from nine selected plain skull dimensions (LCc, LaZ, LaI, CM3, CC, M3M3, LBT, ACo, CM3) and seven relative skull dimensions (LaZ/LCc, LaInf/LCc, LaN/LCc, LaM/LCc, ANc/LCc, LBT/LCc, CM3/LCc) conformed to the results of the empirical comparisons and cluster analysis ( Supplementary Fig. S8 (01-AC-24-2-p-269-298-Supplement.pdf) — CV1 46.37% of variance, CV2 32.62%). They clearly separated the Omani bats as the most distinct sample set (CV1 > 1.6; CV2 ≤ 6.5) in CV2 without an overlap with the four other sample sets, which, however, overlapped in both canonical variables with each other. Because only one skull was available from north-eastern Africa and the type series of escalerae and midas are composed only of holotypes, the relative positions of these three samples to other populations were not evaluated by the canonical analysis.

Fig. 7.

Results of the cluster analysis (UPGMA): differences between mean values of morphometric traits among the particular sets of samples of the Rhinolophus hipposideros group calculated from 27 plain and relative dimensions of skull (A), and from 36 plain and relative dimensions of teeth (B). Samples and sample sets: CEU — Central Europe; WMT — West Mediterranean; EMT — East Mediterranean; CAS — Central Asia; NEA — north-eastern Africa; OMA — Oman; mid — type specimen of R. midas Andersen, 1905

img-z15-4_269.jpg

Comparison of Echolocation Call Parameters

The parameters of echolocation calls of R. hipposideros show similar values in the majority of characteristics throughout most of its distribution range, including central and southern Europe and south-western Asia, although the samples are rather small in some populations and the descriptive information level of the associated data could be limited (Table 5). Two exceptions among the population samples were found in the data from Malta and Oman (Mifsud and Vella, 2019; own data); whereas in the Maltese bats, the frequency values (start, end, peak) were reported to be much higher than in all other populations (≥ 115 kHz in all these parameters, without a value overlap with other populations), in the Omani bats, the peak frequency was found to be extremely low (< 101 kHz, without an overlap with the respective values from other populations; see Fig. 8).

Table 5.

Echolocation parameters in various populations of the Rhinolophus hipposideros group; based on published and original data. Abbreviations: SF = start frequency, EF = end frequency, PF = peak frequency (shown in bold), D = pulse duration, IPI = inter-pulse interval

img-z16-8_269.gif

Discussion

Our analysis uncovered the existence of an unexpected diversity within the R. hipposideros group, challenging its existing phylogenetic and taxonomic arrangement as concluded by Koopman (1994), Horáček et al. (2000), Csorba et al. (2003), Simmons (2005), or Burgin (2019). Genetic and morphological examinations of representative sets of specimens showed two main, geographically exclusive phylogenetic lineages within the group that are well delimited by molecular characteristics and possess two distinct morphotypes and two distinct echotypes. The genetic separation of the lineages is deep and detectable in both nuclear and mitochondrial genomes; in the Cyt-b gene, the uncorrected p-distance of 8.9–10.8% was found, which is roughly twice that considered sufficient for a taxonomic split (Baker and Bradley, 2006); this distance is even higher than those reported by Demos et al. (2019) for various species-pairs in Rhinolophus (or than the results for the species-pairs obtained here, e.g., on average 5.80% for capensis-swinnyi, 4.99% for euryale-mehelyi, 4.07% for ferrumequinumclivosus, or 3.65% for willardi-kahuzi). Thus, the degree of genetic separation of the two lineages within the hipposideros group is sufficient to allow us to consider them as two separate species.

Although one lineage/species was detected in the majority of the distribution range of the R. hipposideros group stretching across the whole southwestern Palaearctic (Europe, north-western and north-eastern Africa, north of the Middle East, Afghanistan, and West Turkestan), the other lineage/species was discovered in a very limited area in the north-eastern regions of Oman. The divergence of these two lineages is estimated to have occurred in the interval 4.3–10.0 Ma, when the more realistic (concerning the fossil evidence) estimation model of Stoffberg et al. (2010) is applied. This age approximately corresponds with the late Miocene period or with the Miocene-Pliocene transition (7.0–5.4 Ma — Herbert et al., 2016); that is, with the periods of dramatic environmental changes that could have led to the separation of species lineages. Alternatively, when Dool's et al. (2016) model is used, the estimated divergence occurred in the interval of 2.0–4.9 Ma, which is roughly at the Pliocene-Pleistocene transition (2.6 Ma; Gibbard et al., 2010) and linked with massive environmental changes as well. However, both time estimations mainly correspond to the main splits of species groups in the Afro-Palaearctic clade of the genus Rhinolophus and are associated with much older periods than most of the estimated divergences of crown pairs of species are within this clade (Stoffberg et al., 2010; Dool et al., 2016).

Fig. 8.

Spectrograms of echolocation call examples of the Rhinolophus hipposideros group (original data); A — an individual recorded inside the Aksamitka Cave, Slovakia; B — an individual foraging at Arjank, Iran (cf. Benda et al., 2012); C — a handled individual recorded at Zingrogh, Tajikistan; D — a handled individual recorded at Misfah, Oman. For details see Materials and Methods

img-z17-5_269.jpg

The first, broadly distributed lineage/species can be easily identified with R. hipposideros (André, 1797) s.str. described from Germany, because only this genetic lineage was discovered in Europe. Based on genetic data, this species was confirmed to occur in the prevailing part of the range of the group as described by e.g., Horáček et al. (2000), Csorba et al. (2003), and Burgin (2019), with the exception of the Caucasus region, southern Iran, and southwestern Arabia.

The Omani lineage/species represents a recently discovered population of the lesser horseshoe bat (cf. Harrison and Bates, 1991; Horáček et al., 2000; Benda et al., 2013). It is known from just six localities (including those where only echolocation call recordings were made) in the Al Hajjar Mountains, situated between Sal Alah (26°02′N, 56°22′E) in the north and Tayman (22°31′N, 59°20′E) in the southeast (some 550–600 km in a line along the Al Hajjar range). Only four specimens were available for examination, which represent a morphotype very distinct from all other examined populations of the lesser horseshoe bat, typified by a very narrow skull with a relatively long rostrum, very large tympanic bullae, a very small first (small) upper premolar (P2), very large third upper molar (M3), very small first lower premolar (P2), and very large second (small) lower premolar (P3). Because the identical morphotype was also detected in the holotype specimen of R. midas Andersen, 1905, this name could also be applied for the Omani species/lineage, and the species has two geographic parts: Omani (known from four bats) and Iranian (known from the type specimen).

The type locality of R. midas is Jask, Hormozgan Province, Iran (25°40′N, 57°49′E), on the Iranian side of the Gulf of Oman, just opposite to the Al Hajjar Mountains of Oman. From the biogeographical perspective, the range of R. midas, lying in limited areas on both sides of the Gulf of Oman, is understandable. Similar geographical patterns of distribution range have been documented in other bats endemic to the Middle East (Harrison and Bates, 1991; Benda et al., 2012); namely, Rhinopoma muscatellum Thomas, 1903 (Rhinopomatidae) to a slightly larger geographical extent than in R. midas and Hypsugo arabicus (Harrison, 1979) (Vespertilionidae), with a distribution pattern very similar to R. midas. Besides R. midas, also R. hipposideros is distributed in Iran. However, according to the available records, it occurs in parapatry with R. midas, in uplands of the central and northern parts of the country (Benda et al., 2012). Only one record of the lesser horseshoe bat from Iran can be theoretically attributable to R. midas besides the type specimen, a bat observed in a cave on Qeshm Island in the Strait of Hormuz (Benda et al., 2012), only 106 km north-west of Sal Alah, the northernmost known site of this bat in Oman, and 250 km west–north-west of Jask, the type locality. The closest site of occurrence of R. hipposideros s.str. in Iran, confirmed by the genetic analysis (and the closest site as well), is the Tadovan Cave in the Zagros Mts. (1,190 m a.s.l.; Fars Prov., 28°51′N, 53°20′E — Shahabi et al., 2019), some 330 km NW of Qeshm Island.

Besides the genetic and morphological differences between R. midas and R. hipposideros s.str., the two species also differ in the pattern of their echolocation calls. Whereas in R. hipposideros s.str. the frequency of maximum energy (peak frequency, PF) of the call was detected around 110 kHz in most populations and only occasionally was it documented within the interval of 100–105 kHz (Benda et al., 2010; Győrössy et al., 2020), in R. midas, the PF was recorded in the interval of 94.1–100.6 kHz (x̄ = 98.2 kHz), i.e., at values much lower than in R. hipposideros s.str. Although no important differences in body size were found between R. hipposideros s.str. and R. midas, the latter species is of a similar size as the Mediterranean populations of the former species, a difference between the species could possibly be present in the auricle size as well as the size of the inner ear (i.e., in characteristics linked with the frequency value of the echolocation call — Huihua et al., 2003). The limited available data suggest such type of difference. The ear length in R. midas from Oman was 18.0–19.0 mm (x̄ = 18.5 mm), and in R. hipposideros s.str. from Iran it was 15.8–18.6 mm (x̄ = 17.1 mm — Benda et al., 2012), and in the bats from Lebanon 14.2–18.8 mm (x̄ = 17.4 mm ― Benda et al., 2016), whereas the forearm length in R. midas was 36.8–38.1 mm (x̄ = 37.4 mm), and in R. hipposideros s.str., it was 37.7–40.9 mm (x̄ = 39.0 mm) from Iran and 35.3–39.4 mm (x̄ = 37.7 mm) from Lebanon. Therefore, although the body size in R. midas is on average smaller than or similar to R. hiposideros s.str. from its geographically closest populations, the ear size seems to be on average larger in R. midas than it is in R. hipposideros s.str. (the species name midas also refers to the large ear size; it was selected by Andersen (1905) most probably after King Midas, a character from Greek mythology who had donkey ears). However, the differences in the external characteristics that would allow species identification remain to be found and tested; the currently available number of samples is too small for any conclusion. However, the size and shape of the nose-leaf of R. midas from Oman seems to be of identical parameters to those in R. hipposideros s.str. (see Fig. 9).

Hence, the R. hipposideros group (or the subgenus Phyllorhina Leach, 1816) now comprises two species, R. hipposideros and R. midas, identically as originally suggested by Andersen (1905) when he established the group. This author originally described R. midas as a separate species; later, he included it into the species rank of R. hipposideros (Andersen, 1918), and this unique morphotype, known from a single specimen until now, was for a long time overlooked. The midas morphotype seems to be rather conservative and perhaps more similar to the ancestral one because it exhibits a smaller degree of the reduction of distal molars and tiny premolars than known in the hipposideros s.str. morphotype.

The results of our analysis can also contribute to a revision of the intraspecific taxonomy of R. hipposideros s.str. Traditionally, the systematic reconstructions were based on body and skull size, nose-leaf shape, and several selected skull and tooth characters, an approach that resulted in delimitations of numerous taxa, namely in the Mediterranean area (see Introduction), a conception introduced by Andersen (1905) and accepted by many authors up to today (see Csorba et al., 2003; Simmons, 2005; Burgin, 2019). The molecular genetic analysis and broad evaluation of morphological characters brought a different view of the phylogenetic relationships within this species. The genetic analysis revealed the existence of two main genetic sublineages within the species, the western lineage, comprising most populations of Europe, including the British Isles, Sardinia, Malta, and Crete, and the Maghreb, and the eastern lineage, comprising the populations of Asia, including Eastern Mediterranean islands (Rhodes and Cyprus), and of Crimea. Both mitochondrial and nuclear markers showed the single Ethiopian sample to be a part of the eastern lineage, although without support for the mtDNA results (this could be a consequence of relatively large geographical distance between localities of the samples from Ethiopia and the Levant). However, the limited samples from West Turkestan (Tajikistan) were placed differently in the topology of both marker types, either into the eastern lineage (mtDNA) or into a separate lineage (nDNA) in a sister position to the above grouping. However, the West Turkestani samples are very limited and their localities are geographically extremely distant from the remaining analysed samples (the direct distance between the Tajikistani and central Iranian localities is some 1,800 km, across deserts and high mountains). Thus, the phylogenetic position of the easternmost populations of R. hipposideros s.str. remains to be investigated more elaborately, employing materials from all parts of Iran and West Turkestan, and from Afghanistan and Kashmir.

Fig. 9.

Portraits of Rhinolophus midas Andersen, 1905 from Oman; A, B — ♂ (NMP 93782), Misfah, Ad Dakhiliyah Province, 9 April 2011, lateral and frontal views; C — ♀ (NMP 93994), Sal Alah, Masandam Province, 13 March 2012, lateral view. Photos by A. Reiter

img-z19-1_269.jpg

Although the geographic division to the western and eastern sublineages was not statistically supported by our results, it conforms to the results of previous analyses (Kůs, 2008; Dool et al., 2013), and it is additionally supported by karyological evidence. The geographical boundary between the lineages seems to be localised at the European-Asian transition between the Balkans and Anatolia, and from this location the boundary between the ranges of the 56- and 58-chromosome races is also reported (Zima et al., 1992; Zima, 2004; Arslan and Zima, 2014). Hence, the separation of the two lineages could actually be linked to the phylogenetic history of the species.

However, the morphological evidence did not contribute markedly to the reconstruction of the intraspecific relationships within R. hipposideros s.str. Two main morphological trends could be demonstrated from the data evaluated: (1) the increase of the body and skull size among the populations along the geographical gradient (latitudinal from the south to the north in the western part of the range, longitudinal from the Mediterranean to the continental climatic zone in the east) and (2) a mosaic-like distribution of characters among populations. The two population sets from the Mediterranean Basin (WMT and EMT) are the most similar to each other in the absolute and relative metric characters, although they belong to two separate sublineages. However, the most distinct population of R. hipposideros s.str. in morphometric traits is that of Central Europe. The representatives of the latter population are on average the largest in body size because they originate from the northernmost area of the species occurrence, although they represent a part of the western sublineage. The populations of the western sublineage share identical haplotypes of the mtDNA despite enormous geographical distances between them (e.g., one universal haplotype was found in Ireland, Great Britain, France, Italy, Austria, Slovenia, Slovakia, Bulgaria, and Greece). This haplotype arrangement suggests relatively recent dispersions of populations across the southern part of Europe and thus a relatively fast evolution of very distinct morphotypes.

The size differences among morphotypes of R. hipposideros may correlate with the changes of climatic conditions along a geographical gradient, in accordance with Bergmann's rule (Bergmann, 1847), that are expected to affect also bat populations (Ashton et al., 2000). On a smaller geographic scale, Salinas-Ramos et al. (2021) recently demonstrated a similar size shift in Italy along approximately 1000 km of the south-north gradient; these authors also explained that it aligned with Bergmann's rule. Other environmental influences that could be responsible for the geography-associated shift in body size, such as the character displacement (cf. Grant, 1972), do not seem to be significant in this bat species. All the evaluated populations come from regions where at least three size categories of horseshoe bats can be observed (i.e., Mediterranean, Central Europe, Central Asia, and north-eastern Africa); therefore, no effects from interspecific competition within the genus and no morphometric or other deflections in particular species were observed (see e.g., Andreas et al., 2013). If the character displacement really influenced the morphometry in R. hipposideros s.str., it would be primarily observed in the British Isles, where only two horseshoe bat species live in sympatry (the medium-sized category is missing). However, the body size of R. hipposideros on these islands is smaller than of the bats in Central Europe (Andersen, 1905; Miller, 1912) where three Rhinolophus species occur and where R. hipposideros would be much smaller if the character displacement works there. The medium body size of the British bats (in relation to the Mediterranean and Central European ones) is most likely caused by the islands' milder climate compared to Central Europe and harsher climate compared to the Mediterranean.

The Central European morphotype is the most distinct within R. hipposideros s.str. because of its extremely large skull size with a relatively narrow braincase and short rostrum, very large first upper premolar (P2), and very small and relatively wide third upper molar (M3). However, these differences seem to be a consequence of the allometric size changes of the skull, where the skull is enlarged in length (mainly the braincase), but is not enlarged to the same degree in width and in tooth-row length; the distal molars are enlarged less than the mesial ones and are relatively short (i.e., seem to be more reduced in length) but are not narrow.

The size differences along the geographical latitude from the Mediterranean to Central Europe were first discussed by Andersen (1905, 1907), who distinguished two subspecies at two edges of this gradient: R. h. hipposideros in the north and R. h. minimus in the south. However, this conception was revised when Saint Girons and Caubère (1966) and Felten et al. (1977) demonstrated a cline changes in metric traits, although Miller (1912) had already considered it to be rather dubious. Our results also give no support for such type of taxonomic division. Already Andersen (1918) demonstrated the mosaic-like distribution of morphological characters among populations of R. hipposideros s.str. in Europe and the Mediterranean and suggested the existence of six separate taxa within this species in the area between Morocco and Ireland in the west and Turkey and Cyprus in the east. This character distribution was again evaluated by Felten et al. (1977), who did not support such division and rather suggested only one, nominotypical subspecies existed in the whole area (except for Crete and Sicily). The echolocation data (another type of evidence) also showed a certain character plasticity within R. hipposideros populations in the Mediterranean area; bats living on the islands of Sardinia and Malta exhibited much higher values of call frequencies (up to 117 kHz on average — Russo et al., 2007; Mifsud and Vella, 2019) than the bats on the European continent. However, these insular bats represent an inner part of the western lineage of R. hipposideros s.str. and do not exhibit any substantial genetic differences from other populations of the lineage (Dool et al., 2013).

The documented pattern of morphological and morphometrical variability in R. hipposideros s.str. does not help when evaluating phylogenetic relationships among examined populations and the echolocation data show similar relevance when assessing the intraspecific variations in this bat species. Therefore, the results of the molecular genetic analysis remain the only evidence that support the reconstruction of the phylogenetic relationships within this species enough. Splitting the species content into two sublineages for both the nuclear and the mitochondrial genomes represents a well-detected separation event. Therefore, the sublineages could be co-identified with two subspecies. In both sublineages, similar levels of plasticity in morphological characters and similar character diversities in echolocation parameters were ascertained.

The taxonomic affiliation of the western sublineage that occurs throughout most of Europe and in the Maghreb is clear. The species is described from Germany (André, 1797) and therefore, this sublineage must to be identified with the nominotypical subspecies. The majority of the available names for R. hipposideros were proposed based on specimens from European type localities, situated in the contemporary countries of Spain, England, Germany, France, Corsica (France), Switzerland, Austria, Czech Republic, and Romania (minor Geoffroy, 1803, minutus Montagu, 1808, bihastatus Geoffroy, 1813, bifer de Blainville, 1840, alpinus Koch, 1865, pallidus Koch, 1865, typus Koch, 1865, kisnyiresiensis Daday, 1885, troglophilus Daday, 1887, helvetica Bretscher, 1904, phasma Cabrera, 1904, typicus Andersen, 1905, majori Andersen, 1918, anomalus Söderlund, 1921, intermedius Söderlund, 1921, moravicus Kostroň, 1943). Therefore, all of them should be considered junior synonyms of the nominotypical subspecies, R. hipposideros hipposideros. Two additional names were created based on bats from Morocco: escalerae Andersen, 1918 and vespa Laurent, 1937. Since the Maghrebian populations are a part of the western sublineage, these two names belong among junior synonyms of R. h. hipposideros. As summarised in Introduction, in the distribution range of the western sublineage (= R. h. hipposideros), up to six different subspecies were reported to occur (escalerae, hipposideros, majori, minutus, minimus, vespa — see Andersen, 1918; Ellerman and Morrison-Scott, 1951; Koopman, 1994; Csorba et al., 2003; Simmons, 2005; Burgin, 2019). However, this arrangement is rejected here because we did not find supporting evidence for it in our results, similar to the results by Dool et al. (2013).

The eastern sublineage of R. hipposideros s.str. is distributed in the Asian range of the species, including the Levant, Asia Minor (including adjacent islands), Crimea, and Iran (except for the Persian Gulf coastal areas). The affiliations of the populations from the eastern parts of the species distribution range (i.e., West Turkestan, Afghanistan, Kashmir) to this sublineage has not been fully resolved. Traditional taxonomic views divided this range into two parts according to the body size: the small-sized Levantine and Turkish populations were assigned to the Mediterranean taxon R. h. minimus (Ellerman and Morrison-Scott, 1951; Harrison, 1964; Koopman, 1994; Csorba et al., 2003; Burgin, 2019) or R. h. hipposideros (Felten et al., 1977; Corbet, 1978), whereas the large-sized eastern populations were assigned to R. h. midas (Andersen, 1905, 1918; Ellerman and Morrison-Scott, 1951; Harrison, 1964; Corbet, 1978; DeBlase, 1980; Harrison and Bates, 1991; Koopman, 1994; Horáček et al., 2000; Csorba et al., 2003; Benda et al., 2012; Burgin, 2019; see also Benda et al. (2006) for a more detailed review). However, our results do not support such west-east separation within the eastern sublineage (see also Dool et al., 2013).

As we demonstrated above, the name midas Andersen, 1905 is unavailable for designation of the Middle Eastern populations of R. hipposideros s.str. because this name is assigned to a different species. Interestingly, in contrast to the western part of the species range of R. hipposideros with 19 available names (see above), no synonym of this species name is currently available based on the material from Asia. However, the single Ethiopian sample examined in our analysis was shown to be a part of the eastern sublineage and it originates from the Yohannis Maikudi Church (13°51′N, 39°27′E) at Degum, Tigray State, approximately 240 km south–south-east of Keren, Eritrea, the type locality of R. minimus von Heuglin, 1861. Therefore, our Ethiopian sample could serve as a reference for topotype population of the latter name, which could be used as R. h. minimus for the eastern sublineage. This name was originally attributed to a separate species by von Heuglin (1861), but was rather early included into the species rank of R. hipposideros by Peters (1871). Andersen (1905, 1907, 1918) used this name for the small-sized Mediterranean populations of the species, but this conception was later questioned (Grulich, 1949; Saint Girons and Caubère, 1966; Felten et al., 1977; Corbet, 1978; Palmeirim, 1990; Benda et al., 2012) and is not supported by our results or the results by Dool et al. (2013). Thus, we consider the name minimus von Heuglin, 1861 to be unavailable for the European and/or Maghrebian populations of R. hipposideros s.str., although numerous recent authors applied this name in a way identical to Andersen's (1905) view (Koopman, 1994; Horáček et al., 2000; Roer and Schober, 2001; Csorba et al., 2003; Simmons, 2005; Burgin, 2019).

As already indicated, the populations of R. hipposideros s.str. that occur in the high mountains of the eastern margin of the species distribution range (Tian Shan, Pamir-Alai, Pamir, Hindu Kush, Karakoram) have an unresolved systematic position because only one specimen was examined for both types of genetic markers. These populations could be a part of the eastern sublineage (R. h. minimus), which is supported by the results of our mitochondrial marker analysis (contra Dool et al., 2013). Alternatively, they could pertain to a separate lineage of the species, as the nuclear markers show (again, contra Dool et al., 2013), and could represent a taxon of their own. In that case, no name would be available for such taxon/populations and it remains to be created (cf. Bates and Harrison, 1997; Csorba et al., 2003).

Samples of two populations of R. hipposideros, from the Caucasus region and from the southwestern part of Arabia, which are important from a biogeographical point of view, were not included in our analysis. Harrison and Bates (1991) identified R. h. minimus in the latter region; whereas from the Caucasus, two forms were reported diversely, R. h. hipposideros (Ognev, 1927; Strelkov, 1963; Kuzâkin, 1965; Koopman, 1994; Roer and Schober, 2001) or R. h. midas (Horáček et al., 2000; Csorba et al., 2003; Rahmatulina, 2005). However, based on the available data, we can estimate that both populations are affiliated with the eastern sublineage. The Caucasus region is situated in a space bordered by Crimea in the northwest and Iran in the south-east. In both of these border regions, the eastern sublineage was detected. Similarly, the south-western region of Arabia is bordered by the Levant in the north and the Ethiopian Highlands in the south, where the eastern sublineage was also detected. Therefore, it is most probable that these two populations belong to R. h. minimus.

Two additional names appeared in literature among synonyms of R. hipposideros (see e.g., Corbet, 1978, 1984; Csorba et al., 2003; Simmons, 2005) — eggenhoeffner Fitzinger, 1870 and billanjani DeBlase, 1972. However, the revisions of original sources showed both names unavailable for zoological nomenclature, being manuscript names (see Fitzinger, 1870; Miller, 1912; DeBlase, 1972, 1980; Benda et al., 2012).

To conclude, the revised taxonomic arrangement of the R. hipposideros group differs greatly from the most frequently presented views in recent years (see Csorba et al., 2003; Simmons, 2005; Burgin, 2019). The group consists of two species: R. hipposideros, which is widespread over south-western Eurasia and north-western and north-eastern Africa, and R. midas, which is distributed in a small range around the Strait of Hormuz and Gulf of Oman. These two species differ from each other in their morphological, genetic, and echolocation parameters. The extensive range of R. hipposideros s.str. is at least inhabited by two subspecies: R. h. hipposideros in the Maghreb and in Europe, west of the Dnieper River (cf. Zagorodniuk, 1999), Bosporus, and the Strait of Karpathos, and R. h. minimus east of this boundary, including the populations of Crimea, (Caucasus), the Middle East and north-eastern Africa (Sudan, Eritrea, Djibouti, Ethiopia). Besides genetic traits, these two subspecies also differ from each other in karyotype: 2n = 58 was found in R. h. minimus, and 2n = 54–56 was found in R. h. hipposideros. However, no significant morphological differences were found between the two subspecies of R. hipposideros.

Supplementary Information

Contents: Supplementary Figures:  Fig. S1 (01-AC-24-2-p-269-298-Supplement.pdf). Maximum likelihood tree of reconstructed phylogenetic relationships of the R. hipposideros group based on as complete as possible cytochrome-b dataset (1,103 bp). Branch support values are shown at the nodes;  Fig. S2 (01-AC-24-2-p-269-298-Supplement.pdf). Maximum likelihood tree of the reconstructed phylogenetic relationships of the R. hipposideros group and selected species of the genus Rhinolophus based on ACOX. Branch support values are shown above/below the branches in order SH-aLRT/UFBoot;  Fig. S3 (01-AC-24-2-p-269-298-Supplement.pdf). Maximum likelihood tree of the reconstructed phylogenetic relationships of the R. hipposideros group and selected species of the genus Rhinolophus based on BGN. Branch support values are shown above/below the branches in order SH-aLRT/UFBoot;  Fig. S4 (01-AC-24-2-p-269-298-Supplement.pdf). Maximum likelihood tree of the reconstructed phylogenetic relationships of the R. hipposideros group and selected species of the genus Rhinolophus based on COPS. Branch support values are shown above/below the branches in order SH-aLRT/UFBoot;  Fig. S5 (01-AC-24-2-p-269-298-Supplement.pdf). Maximum likelihood tree of the reconstructed phylogenetic relationships of the R. hipposideros group and selected species of the genus Rhinolophus based on ROGDI. Branch support values are shown above/below the branches in order SH-aLRT/UFBoot;  Fig. S6 (01-AC-24-2-p-269-298-Supplement.pdf). Maximum likelihood tree of the reconstructed phylogenetic relationships of the R. hipposideros group and selected species of the genus Rhinolophus based on STAT. Branch support values are shown above/below the branches in order SH-aLRT/UFBoot;  Fig. S7 (01-AC-24-2-p-269-298-Supplement.pdf). Chronogram of the family Rhinolophidae based on a Bayesian inference of the nuclear dataset (according to the model by Dool et al., 2016). The numbers at nodes show mean divergence time estimates (Ma) and horizontal boxes 95% highest posterior density intervals of these estimates. The asterisk (*) indicates nodes with low branch support, the rest of the nodes were supported (PP ≥ 0.95);  Fig. S8 (01-AC-24-2-p-269-298-Supplement.pdf). Bivariate plot of skull dimensions of the examined samples of the R. hipposideros group: results of the canonical discriminant analysis of selected nine plain and seven relative dimensions (see Results for details).  Supplementary Tables: Table S1. A (01-AC-24-2-p-269-298-Supplement.pdf)) Original sequences and sequences from GenBank used in the molecular genetic analysis; B) Sequences with the total length of 1,103 bp from GenBank used for the R. hipposideros tree, see  Supplementary Fig. S1; Table S2 (01-AC-24-2-p-269-298-Supplement.pdf). Names, sequences, and annealing temperatures of primers used in this study;  Table S3 (01-AC-24-2-p-269-298-Supplement.pdf). Substitution models as identified by ModelFinder for the different partitions used in MrBayes and IQTREE, respectively;  Table S4 (01-AC-24-2-p-269-298-Supplement.pdf). Summary of BPP for the nuclear dataset. Values for BPP species are posterior probabilities (PP) of delimitation from BPP runs under each of four different schemes under two different algorithms (see Table 1 in Demos et al., 2019);  Table S5 (01-AC-24-2-p-269-298-Supplement.pdf). Relative cranial dimensions of the examined sample sets of the R. hipposideros group; midas, escalerae = dimensions of the respective type specimens; for the sample set delimitations and dimension abbreviations see Materials and Methods;  Table S6 (01-AC-24-2-p-269-298-Supplement.pdf). Relative dental dimensions of the examined sample sets of the R. hipposideros group; midas = dimensions of the respective type specimen; for the sample set delimitations and dimension abbreviations see Materials and Methods;  Table S7 (01-AC-24-2-p-269-298-Supplement.pdf). Results of the one-way ANOVA test of skull dimensions between particular sample sets; for the sample set delimitations and abbreviations and for dimension abbreviations see Materials and Methods. Supplementary Information is available exclusively on BioOne.

Acknowledgements

We thank Riyad Sadek (AUB), Paulina Jenkins, Daphne Hill, and Louise Tomsett (BMNH), Ivan Horáček (CUP), Barbara Rzebik-Kowalska and Bronisław W. Wołoszyn (ISEA), Petr Koubek and Jiří Chamr (IVB), Manuel Ruedi (MHNG), Giuliano Doria (MSNG), Maria Mostadius (MZLU), Teodora Ivanova and †Boân Petrov (NMNHS), Friederike Spitzenberger and Frank Zachos (NMW), †Dieter Kock, †Gerhard Storch, Julia Altmann, and Katrin Krohmann (SMF), and Sergei Kruskop (ZMMU), for accessing museum specimens under their care (for institution abbreviations, see Appendix I), and Martin Cel'uch (BAT-MAN, Bardejov, Slovakia) for providing his unpublished field data. We also thank Ivan Horáček (CUP) for fruitful discussions concerning the divergence time estimations. The preparation of this revision was supported by the Ministry of Culture of the Czech Republic (# DKRVO 2019–2023/6.IX.d, 00023252) and by the Scientific Grant Agency VEGA (1/0298/19).

Literature Cited

1.

Allen, G. M. 1939. A checklist of African mammals. Bulletin of the Museum of Comparative Zoology at Harvard College, 83: 1–763. Google Scholar

2.

Andersen, K. 1904. On von Heuglin's, Rüppell's, and Sundevall's types of African Rhinolophi. Annals and Magazine of Natural History, Seventh Series, 14: 451–458. Google Scholar

3.

Andersen, K. 1905. On some bats of the genus Rhinolophus, with remarks on their mutual affinities, and descriptions of twenty-six new forms. Proceedings of the Zoological Society of London, 1905(2): 75–145. Google Scholar

4.

Andersen, K. 1907. On the geographical races of the lesser horseshoe bat (Rhinolophus hipposiderus). Annals and Magazine of Natural History, 7, 20: 384–389. Google Scholar

5.

Andersen, K. 1918. Diagnoses of new bats of the families Rhinolophidæ and Megadermatidæ. Annals and Magazine of Natural History, 9, 2: 374–384. Google Scholar

6.

André, C. C. 1797. Der Zoologe, oder Compendiöse Bibliothek des wissenswürdigsten aus der Thiergeschichte und allgemeinen Naturkunde. Erste Band. Heft 5–8. Johann Jacob Gebauer, Eisenach and Halle, [xiv] + 416 pp. Google Scholar

7.

Andreas, M., A. Reiter, E. Cepáková, and M. Uhrin. 2013. Body size as an important factor determining trophic niche partitioning in three syntopic rhinolophid bat species. Biologia, 65: 170–175. Google Scholar

8.

Anisimova, M., M. Gil, J.-F. Dufayard, C. Dessimoz, and O. Gascuel. 2011. Survey of branch support methods demonstrates accuracy, power, and robustness of fast likelihood-based approximation schemes. Systematic Biology, 60: 685–699. Google Scholar

9.

Arslan, A., and J. Zima. 2014. Karyotypes of the mammals of Turkey and neighbouring regions: a review. Folia Zoologica, 63: 1–62. Google Scholar

10.

Ashton, K. G., M. C. Tracy, and A. De Queiroz. 2000. Is Bergmann's rule valid for mammals. American Naturalist, 156: 390–415. Google Scholar

11.

Baker, R. J., and R. D. Bradley. 2006. Speciation in mammals and the genetic species concept. Journal of Mammalogy, 87: 643–662. Google Scholar

12.

Bates, P. J. J., and D. L. Harrison. 1997. Bats of the Indian subcontinent. Harrison Zoological Museum, Sevenoaks, 268 pp. Google Scholar

13.

Bechstein, J. M. 1799 [1800]. Thomas Pennant's allgemeine Uebersicht der vierfüßigen Thiere. Zweyter Band. Verlage des Industrie-Comptoirs, Weimar, x +[346] pp. Google Scholar

14.

Bechstein, J. M. 1801. Gemeinnützige naturgeschichte Deutschlands nach allen drey Reichen. Ein Handbuch zur deutlichen und vollständigern Selbstbelehrung besonders für Forstmänner, Jugendlehrer und Oekonomen. Siegfried Lebrecht Crusius, Leipzig, xxx + 1370 pp. Google Scholar

15.

Benda, P., and J. Gaisler. 2015. Bats (Mammalia: Chiroptera) of the Eastern Mediterranean and Middle East. Part 12. Bat fauna of Afghanistan: revision of distribution and taxonomy. Acta Societatis Zoologicae Bohemicae, 79: 267–458. Google Scholar

16.

Benda, P., and J. Mlíkovský. 2022. Nomenclatural notes on the lesser horseshoe bat, Rhinolophus hipposideros (Mammalia: Chiroptera). Journal of the National Museum (Prague), Natural History Series, 191: 5–14. Google Scholar

17.

Benda, P., M. Andreas, D. Kock, R. Lučan, P. Munclinger, P. Nová, J. Obuch, K. Ochman, A. Reiter, M. Uhrin, and D. Weinfurtová. 2006. Bats (Mammalia: Chiroptera) of the Eastern Mediterranean. Part 4. Bat fauna of Syria: distribution, systematics, ecology. Acta Societatis Zoologicae Bohemicae, 70: 1–329. Google Scholar

18.

Benda, P., C. Dietz, M. Andreas, J. Hotový, R. K. Lučan, A. Maltby, K. Meakin, J. Truscott, and P. Vallo. 2008. Bats (Mammalia: Chiroptera) of the Eastern Mediterranean and Middle East. Part 6. Bats of Sinai (Egypt) with some taxonomic, ecological and echolocation data on that fauna. Acta Societatis Zoologicae Bohemicae, 72: 1–103. Google Scholar

19.

Benda, P., P. Georgiakakis, C. Dietz, V. Hanák, K. Galanaki, V. Markantonatou, A. Chudárková, P. Hulva, and I. Horáček. 2009. Bats (Mammalia: Chiroptera) of the Eastern Mediterranean and Middle East. Part 7. The bat fauna of Crete, Greece. Acta Societatis Zoologicae Bohemicae, 72: 105–190. Google Scholar

20.

Benda, P., R. K. Lučan, J. Obuch, A. Reiter, M. Andreas, P. Bačkor, T. Bohnenstengel, E. K. Eid, M. Ševčík, P. Vallo, MR. 2010. Bats (Mammalia: Chiroptera) of the Eastern Mediterranean and Middle East. Part 8. Bats of Jordan: fauna, ecology, echolocation, ectoparasites. Acta Societatis Zoologicae Bohemicae, 74: 185–353. Google Scholar

21.

Benda, P., K. Faizolâhi, M. Andreas, J. Obuch, A. Reiter, M. Ševčík, M. Uhrin, P. Vallo, and S. Ashrafi. 2012. Bats (Mammalia: Chiroptera) of the Eastern Mediterranean and Middle East. Part 10. Bat fauna of Iran. Acta Societatis Zoologicae Bohemicae, 76: 163–582. Google Scholar

22.

Benda, P., A. Reiter, and M. Uhrin. 2013. Bat fauna of the Sultanate of Oman. Mammalian Biology, 78S: 6. Google Scholar

23.

Benda, P., M. R. Abi Said, I. Bou Jaoude, R. Karanouh, R. K. Lučan, R. Sadek, M. Ševčík, M. Uhrin, and I. Horáček. 2016. Bats (Mammalia: Chiroptera) of the Eastern Mediterranean and Middle East. Part 13. Review of distribution and ectoparasites of bats in Lebanon. Acta Societatis Zoologicae Bohemicae, 80: 207–316. Google Scholar

24.

Bendjeddou, M. L., I. Bouam, S. Aulagnier, S. E. Abdelaziz, K. Etayeb, A. D. Mihalca, and A. D. Sándor. 2022. First record of the lesser horseshoe bat, Rhinolophus hipposideros (Borkhausen, 1797), in Libya and potential distribution in North Africa. Mammalia, 86: 328–332. Google Scholar

25.

Bergmann, C. 1847. Ueber die Verhältnisse der Wärmeökonomie der Thiere zu ihre Grösse. Göttinger Studien, 3: 595–708. Google Scholar

26.

De Blainville, H.-M. D. 1840. Ostéographie ou description iconographique comparée de squelette et du systéme dentaire des mammifères récents et fossiles pour servir de base a la zoologie et la géologie. Tome premier. Primatès — secundatès. J. B. Baillière et Fills, Paris, ccxxiii + 47 + 52 + 48 + 40 + 68 + 104 +115 pp. Google Scholar

27.

Blasius, J. H. 1857. Naturgeschichte der Säugethiere Deutschlands und der angrenzenden Länder von Mitteleuropa. Friedrich Vieweg und Sohn, Braunschweig, vi + 549 pp. Google Scholar

28.

Bogdanowicz, W. 1992. Phenetic relationships among bats of the family Rhinolophidae. Acta Theriologica, 37: 213–240. Google Scholar

29.

Borkhausen, M. B. 1797. Deutsche Fauna oder kurzgefaßte Naturgeschichte der Thiere Deutschlands. Erster Theil. Säugethiere und Vögel. Varentrapp und Wenner, Frankfurt am Mayn, xxiv + 620 pp. Google Scholar

30.

Bretscher, K. 1904. Rhinolophus euryale in der Mittelschweiz. Vierteljahrsschrift der Naturforchenden Gesellschaft in Zürich, 49: 254–257. Google Scholar

31.

Burgin, C. J. 2019. Genus Rhinolophus Lacepède, 1799. Pp. 280–332, in Handbook of the mammals of the World. 9. Bats ( D. E. Wilson and R. A. Mittermeier, eds.). Lynx Edicions, Barcelona, 1008 pp. Google Scholar

32.

Cabrera Lattore, A. 1904. Ensayo monográfico sobre los quirópteros de España. Memorias de la Real Sociedad Española de Historia Natural, 2: 249–287. Google Scholar

33.

Castresana, J. 2000. Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Molecular Biology and Evolution, 17: 540–552. Google Scholar

34.

Chernomor, O., A. Von Haeseler, and B. Q. Minh. 2016. Terrace aware data structure for phylogenomic inference from supermatrices. Systematic Biology, 65: 997–1008. Google Scholar

35.

Çoraman, E., A. Furman, A. Karataş, and R. Bilgin. 2013. Phylogeographic analysis of Anatolian bats highlights the importance of the region for preserving the chiropteran mitochondrial genetic diversity in the Western Palaearctic. Conservation Genetics, 14: 1205–1216. Google Scholar

36.

Corbet, G. B. 1978. The mammals of the Palaearctic region: a taxonomic review. British Museum (Natural History) and Cornell University Press, London and Ithaca, v + 314 pp. Google Scholar

37.

Corbet, G. B. 1984. The mammals of the Palaearctic region: a taxonomic review. Supplement. British Museum (Natural History), London, 46 pp. Google Scholar

38.

Csorba, G., P. Ujhelyi, and N. Thomas. 2003. Horseshoe bats of the world (Chiroptera: Rhinolophidae). Alana Books, Bishop's Castle, xxxii + 160 pp. Google Scholar

39.

Daday, J. 1885. Jelentés az erd. orsz. muzeum-egylet ígazgatóválasztmányának megbizásából az 1885. év nyarán végzett chiropterologiai gyüjtések eredményeiről és az erd. orsz. muzeum-egylet denevér-gyüjteményének jegyzéke [A report on the result of bat collections made in the summer 1885 on behalf of the Presidium of the Society of Transylvanian Museum and the list of the bat collection of the Society of Transylvanian Museum]. Orvos-Természettudományi Értesitő [Kolozsvárt], 10: 266–276. [In Hungarian]. Google Scholar

40.

Daday, J. 1887. Új adatok Erdély denevér-faunáják ismeretéhez [New data on the bat fauna of Transylvania]. Értekezések a Természettudományok Köréből [Budapest], 16(7): 1–45. [In Hungarian]. Google Scholar

41.

Deblase, A. F. 1972. The Bats of Iran. Ph.D. Thesis, Oklahoma State University, Stillwater, vi+ 240 pp. Google Scholar

42.

Deblase, A. F. 1980. The bats of Iran: systematics, distribution, ecology. Fieldiana: Zoology (N.S.), 4: xvii + 1–424. Google Scholar

43.

Demos, T. C., P. W. Webala, S. M. Goodman, J. C. K. Peter-Hans, M. Bartonjo, and B. D. Patterson. 2019. Molecular phylogenetics of the African horseshoe bats (Chiroptera: Rhinolophidae): expanded geographic and taxonomic sampling of the Afrotropics. BMC Evolutionary Biology, 19: 1–14. Google Scholar

44.

Dobson, G. E. 1876. Monograph of the Asiatic Chiroptera, and catalogue of the species of bats in the collection of the Indian Museum, Calcutta. Trustees of the Indian Museum, London, x + 228 pp. Google Scholar

45.

Dool, S. E, S. J. Puechmaille, C. Dietz, J. Juste, C. Ibáñez, P. Hulva, S. G. Roué, E. J. Petit, G. Jones, D. Russo , et al. 2013. Phylogeography and postglacial recolonization of Europe by Rhinolophus hipposideros: evidence from multiple genetic markers. Molecular Ecology, 22: 4055–4070. Google Scholar

46.

Dool, S. L., S. J. Puechmaille, N. M. Foley, B. Allegrini, A. Bastian, G. L. Mutumi, T. G. Maluleke, L. J. Odendaal, E. C. Teeling, and D. S. Jacobs. 2016. Nuclear introns outperform mitochondrial DNA in inter-specific phylogenetic reconstructions: lessons from horseshoe bats (Rhinolophidae: Chiroptera). Molecular Phylogenetics and Evolution, 97: 196–212. Google Scholar

47.

Downs, N. C., W. J. Cresswell, P. Reason, G. Sutton, D. Wells, and S. Wray. 2016. Sex-specific habitat preferences of foraging and commuting lesser horseshoe bat Rhinolophus hipposideros (Borkhausen, 1797) in lowland England. Acta Chiropterologica, 18: 451–465. Google Scholar

48.

Ellerman, J. R., and T. C. S. Morrison-Scott. 1951. Checklist of Palaearctic and Indian mammals 1758 to 1946. British Museum, London, 810 pp. Google Scholar

49.

Felten, H., F. Spitzenberger, and G. Storch. 1977. Zur Kleinsäugerfauna West-Anatoliens. Teil IIIa. Senckenbergiana Biologica, 58: 1–44. Google Scholar

50.

Fitzinger, L. J. 1870. Kritische Durchsicht der Ordnung der Flatterthiere oder Handflügler (Chiroptera). Familie der Kammnasen (Rhinolophi). II. Abtheilung. Sitzungsberichte der Kaiserlichen Akademie der Wissenschaften, Mathematisch-Naturwissenschaftliche Classe, Erste Abtheilung, 61(2): 123–198. Google Scholar

51.

Flot, J. F. 2010. SeqPHASE: a web tool for interconverting PHASE input/output files and FASTA sequence alignments. Molecular Ecology Resources, 10: 162–166. Google Scholar

52.

Foley, N. M., V. D. Thong, P. Soisook, S. M. Goodman, K. N. Armstrong, D. S. Jacobs, S. J. Puechmaille, and E. C. Teeling. 2015. How and why overcome the impediments to resolution: lessons from rhinolophid and hipposiderid bats. Molecular Biology and Evolution, 32: 313–333. Google Scholar

53.

Gaisler, J. 2013. Rhinolophus hipposideros Lesser horseshoe bat. Pp. 338–340, in Mammals of Africa. Volume IV. Hedgehogs, shrews and bats ( M. Happold, and D. C. D. Happold, eds.). Bloomsbury Publishing, London, 800 pp. Google Scholar

54.

García-Mudarra, J. L., C. Ibáñez, and J. Juste. 2009. The Straits of Gibraltar: barrier or bridge to Ibero-Moroccan bat diversity. Biological Journal of the Linnean Society, 96: 434–450. Google Scholar

55.

Geoffroy Saint-Hilaire, E.]. 1803. Catalogue des mammifères du Muséum national d'histoire naturelle. Muséum National d'Histoire Naturelle, Paris, 272 pp. Google Scholar

56.

Geoffroy Saint-Hilaire, [E.] 1813. Sur un genre de chauvesouris, sous le nom de Rhinolophes. Annales du Muséum d'Histoire Naturelle [Paris], 20: 254–266. Google Scholar

57.

Gernhard, T. 2008. The conditioned reconstructed process. Journal of Theoretical Biology, 253: 769–778. Google Scholar

58.

Gibbard, P. L., M. J. Head, M. J. C. Walker, and SUBCOMMISION ON QUATERNARY STRATIGRAPHY. 2010. Formal ratification of the Quaternary System/Period and the Pleistocene Series/Epoch with a base at 2.58 Ma. Journal of Quaternary Science, 25: 96–102. Google Scholar

59.

Grant, P. R. 1972. Convergent and divergent character displacement. Biological Journal of the Linnean Society, 4: 39–68. Google Scholar

60.

Grulich, I. 1949. Příspěvek k poznání viariability vrápence malého Rhinolophus hipposideros Bechst. [Contribution to the knowledge of the variability of Rhinolophus hipposideros Bechst.]. Práce Moravskoslezské Akademie Věd Přírodních, 21(5): 1–60. [In Czech]. Google Scholar

61.

Guillén Servent, A., C. M. Francis, and R. E. Ricklefs. 2003. Phylogeny and biogeography of the horseshoe bats. Pp. xii–xxiv, in Horseshoe bats of the world (Chiroptera: Rhinolophidae) ( G. Csorba, P. Ujhelyi, and N. Thomas, eds.). Alana Books, Bishop's Castle, xxxii + 160 pp. Google Scholar

62.

Guindon, S., J.-F. Dufayard, V. Lefort, M. Anisimova, W. Hordijk, and O. Gascuel. 2010. New algorithms and methods to estimate maximum likelihood phylogenies: assessing the performance of PhyML 3.0. Systematic Biology, 59: 307–321. Google Scholar

63.

Győrössy, D., K. Győrössy, and P. Estók. 2020. Comparative analysis of the echolocation calls of the lesser horseshoe bat (Rhinolophus hipposideros) and the Mediterranean horseshoe bat (Rhinolophus euryale) in the Carpathian Basin. North-western Journal of Zoology, 16: 204–210. Google Scholar

64.

Hackett, T. D., M. W. Holderied, and C. Korine. 2017. Echolocation call description of 15 species of Middle-Eastern desert dwelling insectivorous bats. Bioacoustics, 26: 217–235. Google Scholar

65.

Harrison, D. L. 1964. The mammals of Arabia. Volume I. Introduction. Insectivora — Chiroptera — Primates. Ernest Benn Limited, London, 192 pp. Google Scholar

66.

Harrison, D. L., and P. J. J. Bates. 1991. The mammals of Arabia. Second edition. Harrison Zoological Museum, Sevenoaks, 354 pp. Google Scholar

67.

Herbert, T. D., K. T. Lawrence, A. Tzanova, L. C. Peterson, R. Caballero-Gill, and D. S. Kelly. 2016. Late Miocene global cooling and the rise of modern ecosystems. Nature Geoscience, 9: 843–847. Google Scholar

68.

Hoang, D. T., O. Chernomor, A. Von Haeseler, B. Q. Minh, and L. S. Vinh. 2018. UFBoot2: Improving the ultrafast bootstrap approximation. Molecular Biology and Evolution, 35: 518–522. Google Scholar

69.

Horáček, I., V. Hanák, and J. Gaisler. 2000. Bats of the Palearctic region: a taxonomic and biogeographic review. Pp. 11–157, in Proceedings of the VIIIth European Bat Research Symposium. Volume I. Approaches to biogeography and ecology of bats ( B. W. Wołoszyn, ed.). Chiropterological Information Center, Institute of Systematics and Evolution of Animals PAS, Kraków, 273 pp. Google Scholar

70.

Huihua, Z., Z. Shuyi, Z. Mingxue, and Z. Jiang. 2003. Correlations between call frequency and ear length in bats belonging to the families Rhinolophidae and Hipposideridae. Journal of Zoology (London), 259: 189–195. Google Scholar

71.

Ibáñez, C., J. L. García-Mudarra, M. Ruedi, B. Stadelmann, and J. Juste. 2006. The Iberian contribution to cryptic diversity in European bats. Acta Chiropterologica, 8: 277–297. Google Scholar

72.

Kacprzyk, J., E. C. Teeling, C. Kelleher, and M. Volleth. 2016. Wing membrane biopsies for bat cytogenetics: finding of 2n=54 in Irish Rhinolophus hipposideros (Rhinolophidae, Chiroptera, Mammalia) supports two geographically separated chromosomal variants in Europe. Cytogenetics and Genome Research, 148: 279–283. Google Scholar

73.

Kalyaanamoorthy, S., B. Q. Minh, T. K. F. Wong, A. Von Haeseler, and L. S. Jermiin. 2017. ModelFinder: Fast model selection for accurate phylogenetic estimates. Nature Methods, 14: 587–589. Google Scholar

74.

Katoh, K., and D. M. Standley. 2013. MAFFT Multiple Sequence Alignment Software Version 7: improvements in performance and usability. Molecular Biology and Evolution, 30: 772–780. Google Scholar

75.

Koch, C. 1865. Das Wesentliche der Chiroptern mit besonderer Beschreibung der in dem Herzogthum Nassau und den angränzenden Landestheilen vorkommenden Fledermäuse. Jahrbücher des Vereins für Naturkunde im Herzoghtum Nassau, 17–18: 261–593. Google Scholar

76.

Kolenati, F. A. 1860. Monographie der europäischen Chiroptern. Jahresheft der Naturwissenschaftlichen Section der K. K. Mähr. Schles. Gesellschaft für Ackerbau, Natur- und Landeskunde, 1859: 1–156. Google Scholar

77.

Koopman, K. F. 1994. Chiroptera: Systematics. Pp. 1–217, in Handbuch der Zoologie. Band VIII. Mammalia. Teilband 60 ( J. Niethammer, H. Schliemann, and D. Starck, eds.). Walter de Gruyter, Berlin, 223 pp. Google Scholar

78.

Kostroň, K. 1943. Rhinolophus moravicus n. sp., nový vrápenec z Moravy [Rhinolophus moravicus n. sp., a new horseshoe bat from Moravia]. Práce Moravské Přírodovědecké Společnosti [Brno], 15(9): 1–40. [In Czech]. Google Scholar

79.

Kožurina, E. I. 2006. Avtorstvo naučnogo nazvaniâ malogo podkovonosa [Authorship of the scientific name of the lesser horseshoe bat]. Plecotus et al., 9: 43–45. [In Russian]. Google Scholar

80.

Kruskop, S. V. 2012. Order Chiroptera Blumenbach, 1779. Pp. 73–126, in The mammals of Russia: a taxonomic and geographic reference ( I. Y. Pavlinov, and A. A. Lisovsky, eds.). KMK Scientific Press Ltd., Moskva, 604 pp. Google Scholar

81.

Kůs, P. 2008. Molekulárně genetická studie vrápenců východního Středomoří (Chiroptera: Rhinolophidae: Rhinolophus) [Molecular genetic study of horseshoe bats of the Eastern Mediterranean (Chiroptera: Rhinolophidae: Rhinolophus)]. M.Sc. Thesis, Department of Zoology, Charles University, Praha, 81 pp. [In Czech]. Google Scholar

82.

Kuzâkin, A. P. 1965. Otrâd rukokrylye. Ordo Chiroptera [Order Bats. Ordo Chiroptera]. Pp. 79–116, in Opredelitel' mlekopitaûŝih SSSR [Key to identification of mammals of the Soviet Union] ( N. A. Bobrinskij, B. A. Kuznecov, and A. P. Kuzâkin, eds.). Prosveŝenie, Moskva, 282 pp. [In Russian]. Google Scholar

83.

Lacépède, B. G. E. 1799. Tableau des divisions, sous-divisions, ordres et genres des mammifères. Pp. 1–18, in Discours de l'ouverture et de clôture du cours d'histoire naturelle Donné dans le Muséum national d'Histoire naturelle, l'an VII de la République, et Tableaux méthodiques des mammifères et des oiseaux. Plassan, Paris, 57 pp. Google Scholar

84.

Laurent, P. 1937. Le petit rhinolophe fer à cheval au Maroc: description d'une forme nouvelle, Rhinolophus hipposideros vespa subsp. nov. Bulletin de la Société d'Histoire Naturelle de l'Afrique du Nord, 28: 157–165. Google Scholar

85.

Lay, D. M. 1967. A study of the mammals of Iran, resulting from the Street Expedition of 1962–63. Fieldiana: Zoology, 54: 1–282. Google Scholar

86.

Leach, W. E. 1816. Systematic catalogue of the specimens of the indigenous mammalia and birds that are preserved at the British Museum: with their localities and authorities. Richard and Arthur Taylor, London, 42 pp. Google Scholar

87.

Leaché, A. D., and M. K. Fujita. 2010. Bayesian species delimitation in West African forest geckos (Hemidactylus fasciatus). Proceedings of the Royal Society, 277B: 3017–3077. Google Scholar

88.

Li, G., G. Jones, S. J. Rossiter, S. F. Chen, S. Parsons, and S. Zhang. 2006. Phylogenetics of small horseshoe bats from East Asia based on mitochondrial DNA sequence variation. Journal of Mammalogy, 87: 1234–1240. Google Scholar

89.

Lino, A., C. Fonseca, U. Goiti, and M. J. R. Pereira. 2014. Prey selection by Rhinolophus hipposideros (Chiroptera, Rhinolophidae) in a modified forest in Southwest Europe. Acta Chiropterologica, 16: 75–83. Google Scholar

90.

Méhely, L. 1900. Magyarorság denevéreinek monographiája [A monograph on the bats of Hungary]. Hornyánszky Viktor es kir. udvári könyvnyomdája, Budapest, xi + 372 pp. [In Hungarian]. Google Scholar

91.

Mifsud, C. M., and A. Vella. 2019. Acoustic characterization of bats from Malta: setting a baseline for monitoring and conservation of bat populations. Bioacoustics, 28: 427–442. Google Scholar

92.

Miller, G. S. 1912. Catalogue of the mammals of western Europe (Europe exclusive of Russia) in the collection of the British Museum. British Museum, London, xv + 1019 pp. Google Scholar

93.

Miller, M. A., W. Pfeiffer, and T. Schwartz. 2010. Creating the CIPRES Science Gateway for inference of large phylogenetic trees. In : Gateway Computing Environments Workshop (GCE), 14 Nov. 2010 (ANONYMOUS, ed.). Curran Associates, Inc., New Orleans, 8 pp. Google Scholar

94.

Montagu, G. 1808. An account of the larger and lesser species of horseshoe bats, proving them to be distinct; together with a description of Vespertilio barbastellus, taken in the south of Devonshire. Transactions of the Linnean Society of London, 9: 162–173. Google Scholar

95.

Nguyen, L., H. A. Schmidt, A. Von Haesler, and B. Q. Minh. 2015. IQ-TREE: a fast and effective stochastic algorithm for estimating maximum likelihood phylogenies. Molecular Biology and Evolution, 32: 268–274. Google Scholar

96.

Ognev, S. I. 1927. A synopsis of the Russian bats. Journal of Mammalogy, 8: 140–157. Google Scholar

97.

Obrist, M. K., R. Boesch, and P. F. Flückiger. 2004. Variability in echolocation call design of 26 Swiss bat species: consequences, limits and options for automated field identification with a synergetic pattern recognition approach. Mammalia, 68: 307–322. Google Scholar

98.

Parsons, S., and G. Jones. 2000. Acoustic identification of twelve species of echolocating bat by discriminant function analysis and artificial neural networks. Journal of Experimental Biology, 203: 2641–2656. Google Scholar

99.

Peters, W. 1871. Über die Gattungen und Arten der Hufeisennasen, Rhinolophi. Monatsberichte der Königlich-Preussischen Akademie der Wissenschaften zu Berlin, 1871: 301–332. Google Scholar

100.

Palmeirim, J. 1990. Bats of Portugal: zoogeography and systematics. Univeristy of Kansas Museum of Natural History, Miscellaneous Publications, 82: 1–53. Google Scholar

101.

Panouse, J. B. 1951. Les chauves-souris du Maroc. Travaux de l'Institut Scientifique Chérifien, 1: 1–120. Google Scholar

102.

Papadatou, E., R. K. Butlin, and J. D. Altringham. 2008. Identification of bat species in Greece from their echolocation calls. Acta Chiropterologica, 10: 127–143. Google Scholar

103.

Puechmaille, S. J., M. Argouilh, P. Piyapan, M. Yokubol, K. M. Mie, P. J. J. Bates, C. Satasook, T. New, B. S. S. Hla, I. J. Mackie , et al. 2011. The evolution of sensory divergence in the context of limited gene flow in the bumblebee bat. Nature Communications, 2 (573): 1–9. Google Scholar

104.

Qumsiyeh, M. B. 1985. The bats of Egypt. Special Publication, The Museum of Texas Tech University, 23: 1–101. Google Scholar

105.

Rahmatulina, I. K. 2005. Rukokrylye Azerbajdžana (fauna, èkologiâ, zoogeografiâ) [Bats of Azerbaijan (fauna, ecology, zoogeography)]. Nacional'naâ Akademiâ Nauk Azerbajdžana, Baku, 475 pp. [In Russian]. Google Scholar

106.

Rambaut, A., M. A. Suchard, D. Xie, and A. Drummond. 2014. Tracer v1.6.  http://beast.bio.ed.ac.uk/TracerGoogle Scholar

107.

Rannala, B., and Z. Yang. 2003. Bayes estimation of species divergence times and ancestral population sizes using DNA sequences from multiple loci. Genetics, 164: 1645–1656. Google Scholar

108.

Roer, H., and W. Schober. 2001. Rhinolophus hipposideros (Bechstein, 1800) — Kleine Hufeisennase. Pp. 39–58, in Handbuch der Säugetiere Europas. Band 4: Fledertiere. Teil I: Chiroptera I. Rhinolophidae, Vespertilionidae 1 ( F. Krapp, ed.). Aula-Verlag, Wiebelsheim, x + 602 pp. Google Scholar

109.

Ronquist, F., and J. P. Huelsenbeck. 2003. MRBAYES 3: Bayesian phylogenetic inference under mixed models. Bioinformatics, 19: 1572–1574. Google Scholar

110.

Russo, D., and G. Jones. 2002. Identification of twenty-two bat species (Mammaia: Chiroptera) from Italy by analysis of time-expanded recordings of echolocation calls. Journal of Zoology (London), 258: 91–103. Google Scholar

111.

Russo, D., M. Mucedda, M. Bello, S. Biscardi, E. Pidinchedda, and G. Jones. 2007. Divergent echolocation frequencies in insular rhinolophids (Chroptera): a case of character displacement? Journal of Biogeography, 34: 2129–2138. Google Scholar

112.

Saint Girons, M.-C., and B. Caubère. 1966. Notes sur les mammiferes de France. V. — Sur la répartition de Rhinolophus hipposideros hipposideros (Bechstein, 1800) et Rhinolophus hipposideros minimus Heuglin, 1861. Mammalia, 30: 308–326. Google Scholar

113.

Salicini, I., C. Ibáñez, and J. Juste. 2011. Multilocus phylogeny and species delimitation within the Natterer's bat species complex in the Western Palearctic. Molecular Phylogenetics and Evolution, 61: 888–898. Google Scholar

114.

Salinas-Ramos, V. B., P. Agnelli, L. Bosso, L. Ancillotto, V. Sánchez-Cordero, and D. Russo. 2021. Body size variation in Italian lesser horseshoe bats Rhinolophus hipposideros over 147 years: exploring the effects of climate change, urbanization and geography. Biology [Basel], 10(16): 1–10. Google Scholar

115.

Shahabi, S., M. Sharifi, and V. Akmali. 2019. Echolocation call frequency and mitochondrial control region variation in the closely related bat species of the genus Rhinolophus (Chiroptera: Rhinolophidae) occurring in Iran: implications for taxonomy and intraspecific phylogeny. Mammal Research, 64: 485–501. Google Scholar

116.

Simmons, N. B. 2005. Order Chiroptera. Pp. 312–529, in Mammal species of the world. A taxonomic and geographic reference,  3rd edition . Volume 1 ( D. E. Wilson, and D. M. Reeder, eds.). Johns Hopkins University Press, Baltimore, xxxv + 743 pp. Google Scholar

117.

Smirnov, D. G., G. S. Džamirzoev, and U. A. Bykov. 2022. Sravnitel'naâ harakteristika èholokacionnyh signalov treh vidov podkovonosov (Chiroptera, Rhinolophidae) Vostočnogo Kavkaza (Dagestan) [Comparative characteristics of the echolocation calls in three species of horseshoe bats (Chiroptera, Rhinolophidae) of the eastern Caucasus (Dagestan)]. Izvestiâ Vysših Učebnyh Zavedenij, Povolžskij Region, Estestvennye Nauki, 2: 36–48. [In Russian]. Google Scholar

118.

Söderlund, F. 1921. Zwei neue Arten der Gattung Rhinolophus, gefunden im Wildbad Gastein. Zoologischer Anzeiger, 52: 122–125. Google Scholar

119.

Stoffberg, S., D. S. Jacobs, I. J. Mackie, and C. A. Matthee. 2010. Molecular phylogenetics and historical biogeography of Rhinolophus bats. Molecular Phylogenetics and Evolution, 54: 1–9. Google Scholar

120.

Strelkov, P. P. 1963. II. Otrâd Chiroptera — Rukokrylye [II. Order Chiroptera — bats]. Pp. 122–218, in Mlekopitaûŝie fauny SSSR. Čast' 1 [Mammals of the fauna of the Soviet Union. Part 1] ( I. I. Sokolov, ed.). Izdatel'stvo Akademii nauk SSSR, Moskva and Leningrad, 638 pp. [In Russian]. Google Scholar

121.

Tamura, K., G. Stecher, and S. Kumar. 2021. MEGA11: Molecular Evolutionary Genetics Analysis Version 11. Molecular Biology and Evolution, 38: 3022–3027. Google Scholar

122.

Taylor, P. J., A. Macdonald, S. M. Goodman, T. Kearney, F. P. Cotterill, S. Stoffberg, A. Monadjem, M. C. Schoeman, J. Guyton, P. Naskrecki, and L. R. Richards. 2018. Integrative taxonomy resolves three new cryptic species of small southern African horseshoe bats (Rhinolophus). Zoological Journal of the Linnean Society, 184: 1249–1276. Google Scholar

123.

Thong, V. D., S. J. Puechamaille, A. Denzinger, P. J. J. Bates, C. Dietz, G. Csorba, P. Soisook, E. C. Teeling, S. Matsumura, N. M. Furey, and H.-U. Schnitzler. 2012. Systematics of the Hipposideros turpis complex and a description of a new subspecies from Vietnam. Mammal Review, 42: 166–192. Google Scholar

124.

Trifinopoulos, J., L.-T. Nguyen, A. Von Haeseler, and B. Q. Minh. 2016. W-IQTREE: a fast online phylogenetic tool for maximum likelihood analysis. Nucleic Acids Research, 44: W232–W235. Google Scholar

125.

Trouessart, E.-L. 1879. Catalogue des mammifères vivants et fossiles. Ordo III. Chiroptera. Revue et Magazin de Zoologie Pure et Appliquée, (3)6: 201–254. Google Scholar

126.

Tupinier, Y. 2001. Historique de la description des espèces européennes de Chiroptères. Le Rhinolophe, 15: 1–140. Google Scholar

127.

Volleth, M., M. Biedermann, W. Schorcht, and K.-G. Heller. 2013. Evidence for two karyotypic variants of the lesser horseshoe bat (Rhinolophus hipposideros, Chiroptera, Mammalia) in Central Europe. Cytogenetic and Genome Research, 14: 55–61. Google Scholar

128.

Von Heuglin, T. 1861. Beiträge zur Fauna der Säugethiere N.O.-Afrika's. Novorum Actorum Academiae Caesareae Leopoldino-Carolinae Germanicae Naturae Curiosorum, 29(8): 1–24. Google Scholar

129.

Yang, Z., and B. Rannala. 2010. Bayesian species delimitation using multilocus sequence data. Proceedings of the National Academy of Sciences of the United States of America, 107: 9264–9269. Google Scholar

130.

Yule, G. U. 1925. II. A mathematical theory of evolution, based on the conclusions of Dr. J. C. Willis, F. R. S. Philosophical Transactions of the Royal Society of London, 213B: 21–87. Google Scholar

131.

Zagorodniuk, I. V. 1999. Taxonomy, biogeography and abundance of the horseshoe bats in eastern Europe. Acta Zoologica Cracoviensia, 42: 407–421. Google Scholar

132.

Zima, J. 2004. Karyotypic variation in mammals of the Balkan Peninsula. Pp. 109–133, in Balkan biodiversity. Pattern and process in the European hotspot ( H. I. Griffiths, B. Kryštufek, and J. M. Reed, eds.). Kluwer Academic Press, Dordrecht, ix + 357 pp. Google Scholar

133.

Zima, J., M. Volleth, I. Horáček, J. Červený, A. Červená, K. Průcha, and M. Macholán. 1992. Comparative karyology of rhinolophid bats (Chiroptera: Rhinolophidae). Pp. 229–236, in Prague studies in mammalogy ( I. Horáček and V. Vohralík, eds.). Charles University Press, Praha, xxi + 246 pp. Google Scholar

Appendices

Appendix I

List of the specimens examined in the morphological analysis; an asterisk (*) denote specimens used also in the molecular genetic analysis. Collection abbrevitations: AUB = American University Beirut, Lebanon; BMNH = Natural History Museum, London, United Kingdom; CUP = Department of Zoology, Charles University, Prague, Czech Republic; ISEA = Institute of Systematics and Evolution of Animals, Polish Academy of Sciences, Kraków, Poland; IVB = Institute of Vertebrate Biology, Academy of Sciences of the Czech Republic, Brno, Czech Republic; MHNG = Natural History Museum, Geneva, Switzerland; MNHN = National Museum of Natural History, Paris, France; MSNG = Civil Natural History Museum Giacomo Doria, Genoa, Italy; MZLU = Museum of Zoology and Entomology, Lund University, Sweden; NMNHS = National Museum of Natural History, Sofia, Bulgaria; NMP = National Museum (Natural History), Prague, Czech Republic; NMW = Natural History Museum, Vienna, Austria; OHC = Otto von Helversen Collection, Erlangen, Germany; SMF = Senckenberg Museum and Research Institute, Frankfurt am Main, Germany; ZMMU = Zoological Museum, Moscow State University, Moscow, Russia

Afghanistan: 1 ♀ (IVB af547 [S+B]), Abdukil at Shigi, cave above the Kunar river, 1 April 1967, leg. J. Gaisler, D. Povolný, Z. Šebek and F. Tenora; — 1 ♀ (SMF 39214 [S+A]), Barg-i-Matal, Konar, 2010 m, 21 July 1964, leg. D. Meyer-Oehme; — 1S+A]), Dahan Ghar, Wardak, Höhle, 2020 m, 12 March 1965, leg. D. Meyer-Oehme; — 1 ♂ (MZLU L58/3277, L58/3321 [S+A]), Grotte Boulan, 9 April 1958, leg. K. Lindberg; — 2 ♂♂ (IVB af1388 [B], af1389 [S+B]), Jalal Abad, hotel, attic, 19 February 1965, leg. D. Povolný and F. Tenora; — 2 ♂♂ (SMF 39217 [S+A], 39218 [A]), Jalalabad, Nangarhar, 650 m, 3 August 1965, leg. D. Meyer-Oehme; — 1 ♂ (IVB af1057 [S+B]), Lalanda, Lalanda cave, 20 km S of Kabul, 12 May 1967, leg. J. Gaisler, D. Povolný, Z. Šebek and F. Tenora; — 1 ♀ (IVB af573 [S+B]), Sarobi, cave above the Sarobi–Kabul road, 5 April 1967, leg. J. Gaisler, D. Povolný, Z. Šebek and F. Tenora; — 1 ♂, 1 ♀ (SMF 39213, 39215 [S+A]), Tscharasiaw, Logar, 1850 m, 23 September 1963, 2 October 1963, leg. D. Meyer-Oehme.

Albania: 1 ♂ (NMP 96541 [S+A*]), Gjirokastër, castle, 27 January 2016, leg. F. Bego, P. Benda and M. Uhrin; — 1 ♂(NMP 96536 [S+A*]), Gollomboç, Hermit Cave, 25 January 2016, leg. F. Bego, P. Benda and M. Uhrin; — 1 ♂ (NMP 96531 [S+A*]), Tren, Treni Cave, 25 January 2016, leg. F. Bego, P. Benda and M. Uhrin; — 1 ♂ (NMP 96551 [A*]), Vithkuq, chapel crypt, 27 June 2016, leg. P. Benda and M. Uhrin.

Algeria: 1 ♂ (ISEA 9586 [S+B]), 20 km NW of Sebdou, 6 November 1981, leg. K. Kowalski and B. Rzebik-Kowalska; — 2 ♂♂ (ISEA 9584, 9585 [S+B]), Brezina, cave, 31 October 1981, leg. K. Kowalski and B. Rzebik-Kowalska; — 1 ♂ (IVB A204 [S+B]), Gorges de Kherrata, tunnel, 15 January 1982, leg. J. Gaisler; — 1 ♂ (ISEA 9587 [S+B]), Misserghin, 14 December 1982, leg. K. Kowalski and B. Rzebik-Kowalska; — 1 ♂, 1 ♀ (ISEA 9588, 9664 [S+B]), Sig, 4 January 1983, 25 January 1983, leg. K. Kowalski and B. Rzebik-Kowalska; — 1 ♂ (IVB A237 [S+B]), Sebdou, 1 May 1982, leg. J. Gaisler.

Azerbaijan: 1 ♀ (NMP 91697 [S+B]), Suçma, Şəki District, 25 April 1976, leg. I. Rakhmatulina.

Bulgaria: 2 ♂♂ (NMP 49788, 49789 [S+A]), Âgodina, Gorna Karanska dupka Cave, 16 August 1978, leg. P. Donát, J. Flegr, J. Janda and V. Vohralík; — 5 ♂♂, 1 ♀ (NMP 49780–49786 [S+A]), Âgodina, Imamova dupka Cave, 15 August 1978, leg. P. Donát, J. Flegr, J. Janda and V. Vohralík; — 1 ♀(NMP 49807 [S+A]), Bačkovo, cave, 30 July 1979, leg. D. Holečková, P. Donát, I. Horáček, J. Jirouš and V. Vohralík; — 2 ♂♂ (NMP 49434, 49435 [S+A]), Bačkovo, Bačkovski Monastery, 14 July 1976, leg. M. Braniš, V. Hanák, I. Horáček, K. Hůrka, J. Jirouš, V. Švihla and V. Vohralík; — 1 ♀ (NMNHS unnum. [S]), Borovo, 19 March 1968, leg. P. Beron; — 2 ♂♂, 3 ♀♀ (NMP 50091–50095 [S+B]), Brestnica, Saeva dupka Cave, 8 February 1965, leg. J. Figala, J. Gaisler, V. Hanák and K. Hůrka; — 1 ♂ (NMP 49433 [S+A]), Čepelare, 13 July 1976, leg. M. Braniš, V. Hanák, I. Horáček, K. Hůrka, J. Jirouš, V. Švihla and V. Vohralík; — 1 ♂ (NMNHS unnum. [S]), Filipovci, 27 February 1967, leg. P. Beron; — 1 ind. (NMNHS unnum. [S]), Ginci, Tošova dupka Cave, 17 February 1968, leg. P. Beron; — 5 ♀♀ (NMP 50027–50031 [S+A]), Gorna Breznica, 24 July 1981, leg. J. Flousek, R. Fuchs and V. Vohralík; — 2 ♂♂, 1 ♀ (NMP 49354, 49758, 49777 [S+A]), Karlukovo, 5 July 1976, 8 August 1978, 9 August 1978, leg. M. Braniš, P. Donát, J. Flegr, V. Hanák, I. Horáček, K. Hůrka, J. Janda, J. Jirouš, V. Švihla and V. Vohralík; — 1 ♂ (NMP 50080 [S+B]), Karlukovo, Bankova peŝera Cave, 7 February 1965, leg. J. Figala, J. Gaisler, V. Hanák and K. Hůrka; — 1 ♂ (NMP 49753 [S+A]), Karlukovo, Temnata dupka Cave, 7 August 1978, leg. P. Donát, J. Flegr, J. Janda and V. Vohralík; — 5 ♂♂(NMP 49793–49797 [S+A]), Kotel, 15 July 1979, leg. D. Holečková, P. Donát, I. Horáček, J. Jirouš and V. Vohralík; — 1 ind. (NMNHS N12 [S]), Kričim, date unlisted, leg. I. Bureš; — 2 ♂♂ (NMP 50136, 50137 [S+B]), Lakatnik, Svinskata peŝera Cave, 19 March 1956, collector unlisted; — 1 ind. (NMP 49813 [S+B]), Lakatnik, Temnata dupka Cave, 3 January 1962, leg. J. Sklenář; — 1 ♂, 4 ♀♀ (NMP 49368–49372 [S+A], Lilânovo, 9 July 1976, leg. M. Braniš, V. Hanák, I. Horáček, K. Hůrka, J. Jirouš, V. Švihla and V. Vohralík; — 5 ♂♂, 2 ♀♀(NMP 49997–50003 [S+A]), Orehovo, 30 August 1980, leg. D. Holečková, J. Jirouš, H. Prágerová and V. Vohralík; — 1 ♀(NMNHS 739 [S]), Pepelina, Orlova čuka Cave, February 1961, leg. I. Ivanov; — 1 ♀ (IVB 398 [S+B]), Peŝera, Lilova skala Cave, 3 February 1965, leg. J. Figala, J. Gaisler, V. Hanák and K. Hůrka; — 1 ♂ (NMP 50072 [S+B]), Peŝera, Nova peŝera Cave, 4 February 1965, leg. J. Figala, J. Gaisler, V. Hanák and K. Hůrka; — 1 ♂ (NMP 50076 [S+B]), Peŝera, Snežânka Cave, 5 February 1965, leg. J. Figala, J. Gaisler, V. Hanák and K. Hůrka; — 1 ♀ (NMP 49347 [S+B]), Ropotamo, 6 June 1957, leg. V. Hanák; — 2 ♂♂ (NMNHS N63, unnum. [S]), Studen Kladenec, mine, 3 May 1996, leg. T. Ivanova; — 1 ♂ (NMNHS unnum. [S]), Treklâno, date and collector unlisted; — 3 ♂♂, 1 ♀ (NMNHS unnum. [S]), Urvič, 8 April 1971, leg. V. Beškov.

Croatia: 1 ♂, 1 ♀ (NMP 96815 [S+A], 96816 [A]), Pokrovnik, Škarin Samograd Cave, 5 September 1977, leg. J. Červený and J. Kučera.

Cyprus: 1 ♂ (NMP 97092 [S+A]), Afendrika, Panagia Hrysiotissa, cave, 21 January 2018, leg. P. Benda and M. Uhrin; — 2 ♀♀ (MSNG 44488 [A]), Akantu (Cipro), 12 January 1899, leg. Cecconi; — 1 ♂ (NMP 97121 [S+A]), Alevkaya, Küpö Cave, 2 October 2018, leg. P. Benda and M. Uhrin; — 2 ♂♂ (NMP 90424, 91269 [S+A*]), Cinarli, Inçirli Cave, 6 April 2005, 17 April 2005, leg. P. Benda, V. Hanák, I. Horáček, P. Hulva and R. Lučan; — 4 ♂♂, 2 ♀♀ (NMP 90923–90928 [S+A*]), Troodos Forest, valley south of Kakopetria, mine, 27 July 2006, leg. P. Benda.

Czech Republic: 1 ♀ (NMP 343/64 [S]), Jílové u Prahy, 30 October 1964, leg. V. Hanák; – 1 ♂ (NMP E7 [S]), Karlštejn, mine, 15 February 1957, leg. V. Hanák; — 1 ♀ (NMP 155/62 [B]), Lednice, 9 June 1962, leg. V. Hanák; — 2 ♂♂, 1 ♀ (NMP 341/58, 347/58, 348/58 [S]), Mníšek pod Brdy, 8 March 1958, leg. J. Sklenář; — 1 ♀ (NMP 422/59 [S]), Svitavy, 4 March 1959, leg. Stach; — 1 ♂ (NMP ZN17 [S]), Vranov nad Dyjí, castle attic, 31 July 1957, leg. V. Hanák; — 1 ♂, 1 ind. (NMP ZB11, ZB12 [S]), Zbraslav, 1 December 1956, leg. V. Hanák; — 2 ♀♀ (NMP ZN26, ZN27 [S]), Znojmo, castle attic, 3 August 1957, leg. V. Hanák.

Ethiopia: 1 ♀ (NMP 95890 [S+A*]), Degum, Yohannis Maikudi Church, 31 October 2012, leg. P. Benda.

Greece: 1 ♂, 6 ♀♀ (NMP 48710–48715, 49028 [S+A*]), Kompotades, bunker, 9 September 1996, 10 September 1996, 31 August 2001, leg. M. Andreas, P. Benda and M. Uhrin; — 1 ♂ (NMP 92303 [A]), Krītī, Avdoy, Agios Fōteinīs Cave, 10 October 2007, leg. P. Benda; — 2 ♂♂, 1 ♀ (NMP 91193, 91194 [S+A*], 92292 [A*]), Krītī, Gerani, Geranioy Cave, 6 October 2006, 8 October 2007, leg. P. Benda, V. Hanák and P. Hulva; — 1 ♂ (NMP 92320 [S+A*]), Krītī, Kritsa, Gaidoyrotrypa Cave, 14 October 2007, leg. P. Benda; — 1 ♂, 1 ♀ (NMP 91197, 91198 [S+A*]), Krītī, Milatos, Milatoy Cave, 7 October 2006, leg. P. Benda, V. Hanák and P. Hulva; — 1 ♂(NMP 92290 [A*]), Krītī, Ploytī, Mikrī Lavyrinthos Cave, 7 October 2007, leg. P. Benda; — 1 ♂ (NMP 92317 [S+A*]), Krītī, Sitanos, Exō Latsidi Cave, 13 October 2007, leg. P. Benda; — 2 ♂♂, 2 ♀♀ (NMP 92297–92300 [S+A*]), Krītī, Theriso, Sarakinas Cave, 8 October 2007, leg. P. Benda; — 1 ♀ (NMP 48643 [S+B]), Marōneia, Kyklōpa Cave, 19 June 1989, leg. R. Chaloupka, V. Hanák and V. Vohralík; — 1 ♂(NMP 96614 [S+A*]), Rodos, Agios Paylos, 16 August 2012, leg. P. Benda; — 3 ♀♀ (NMP 96615, 96616 [S+A*], 96617 [A*]), Rodos, Gadoyra Dam, hut, 17 August 2012, leg. P. Benda.

Iran: 1 ♂ (NMP 94427 [A]), Assalem, 3 October 2002, leg. P. Hulva; — 1 ♀ (MHNG 1905.3 [A]), Bouchir, Brazjan, June 1968, leg. A. Arata; — 3 ♂♂ (NMP 48096, 48097, 48439 [S+A*]), Emamzadeh (Esfahan Prov.), 1 May 1997, 6 April 2000, leg. P. Benda and A. Reiter; – 1 ♀ (BMNH 94.11.16.1 [S], holotype of Rhinolophus midas Andersen, 1905), Jask, Persian Gulf, date and collector unlisted; — 1 ♂ (NMP 39588 [A]), Karaj River valley, 1934, leg. Kargl; — 1 ind. (NMP 93858 [S+Sk]), Moghan Cave, October 1999, leg. K. Faizolahi; — 1 ♀(NMP 48117 [S+A*]), Nosrat Abad, 7 May 1997, leg. P. Benda; — 1 ♂ (NMW 21008 [S+A]), Schiras, 1894, leg. B. Wagschal.

Jordan: 1 ind. (NMP 92842 [S+Sk]), Bait Idis, Jesus' Cave, 15 July 2010, leg. P. Benda and A. Reiter; — 2 ♀♀ (NMP 92409, 92410 [S+A*]), Dibbin, Dibbin Forest, underground corridor, 27 October 2008, leg. P. Benda and J. Obuch; — 1 ♂, 2 ♀♀ (NMP 92508–92510 [S+A*]), Zubiya, Zubiya Cave, 24 May 2009, leg. P. Benda and A. Reiter.

Kirghizstan: 1 ♀ (NMP 58323 [S+A]), Kyzyl-Kiâk, cave, 30 June 1988, leg. J. Červený and J. Obuch; — 1 ♀ (NMP 58324/2 [S+A]), Toâ-Moûn, Kolodec Fersmana mine, 12 July 1988, leg. J. Červený and J. Obuch.

Kosovo: 1 ♂ (NMP 96803 [S+A]), Bubël, cave, 27 October 2001, leg. P. Benda.

Lebanon: 3 ♂♂ (NMP 91806, 93709 [S+A*], 91807 [A*]), Aamchit, Saleh Cave, 28 January 2007, 25 March 2009, leg. T. Bartonička, P. Benda, R. Černý, I. Horáček and R. Lučan; — 1 ♂, 1 ♀ (NMP 93552 [S+A*], 93553 [A*]), Aanjar, Aanjar Cave, 5 June 2010, leg. P. Benda and M. Uhrin; — 1 ♀ (NMP 91782 [S+A]), Afqa Cave, 22 January 2007, leg. P. Benda, R. Černý, I. Horáček and R. Lučan; — 1 ♂ (NMP 91798 [S+A*]), Antelias, Kenaan Cave, 25 January 2007, leg. P. Benda, R. Černý, I. Horáček and R. Lučan; — 1 ♂ (AUB M170 [B]), Beit ed Dine, tunnel under building, 7 September 1960, leg. J. E. Stencel; — 1 ♀ (NMP 93711 [A*]), Dahr El Mghara, Aaonamie Cave, 28 March 2009, leg. T. Bartonička, P. Benda, I. Horáček and R. Lučan; — 1 ♂ (NMP 91775 [S+A*]), Er Roueiss Cave, 22 January 2007, leg. P. Benda, R. Černý, I. Horáček and R. Lučan; — 1 ♂ (NMP 91801 [A*]), Faraya, El Qana Cave, 27 January 2007, leg. P. Benda, R. Černý, I. Horáček and R. Lučan; — 2 ♂♂ (NMP 93537, 93538 [S+A*]), Faraya, Raymond Cave, 2 June 2010, leg. P. Benda and M. Uhrin; — 1 ♂, 1 ♀ (NMP 91769 [A*], 91770 [S+A]), Haqel El Azime, Achou Cave, 21 January 2007, leg. P. Benda, R. Černý, I. Horáček and R. Lučan; — 2 ♂♂ (NMP 91802 [A*], 91906 [S+A*]), Hrajel, Seraaya Cave, 27 January 2007, 20 January 2008, leg. P. Benda, R. Černý, I. Horáček, R. Lučan and M. Uhrin; — 1 ♀ (NMP 95792 [S+A*]), Jezzine, Pont El Khalass, 23 June 2006, leg. I. Horáček, P. Hulva, R. Lučan and P. Němec; — 3 ♂♂, 1 ♀ (NMP 91753–91755 [S+A*], 91756 [A*]), Marjaba, mine, 19 January 2007, leg. P. Benda, R. Černý, I. Horáček and R. Lučan; — 1 ♂ (NMP 91809 [S+A*]), Nabaa Es Safa, mine, 29 January 2007, leg. P. Benda, R. Černý, I. Horáček and R. Lučan; — 1 ♂, 1 ♀ (NMP 91789, 91790 [S+A*]), Qadisha Cave, 23 January 2007, leg. P. Benda, R. Černý, I. Horáček and R. Lučan; — 1 ♀ (NMP 93577 [S+A*]), Seraal, 10 June 2010, leg. P. Benda and M. Uhrin; — 1 ♂ (NMP 91786 [S+A]), Tourzaiya, Mebaaj Cave, 23 January 2007, leg. P. Benda, R. Černý, I. Horáček and R. Lučan; — 1 ♀ (NMP 93706 [S+A*]), Wadi Jilo, 22 March 2009, leg. T. Bartonička, P. Benda, I. Horáček and R. Lučan.

Morocco: 1 ♀ (NMP 93602 [S+A*]), Gorges du Dadès, Aït-Ali, 7 October 2010, leg. P. Benda, A. Reiter, M. Ševčík and M. Uhrin; — 1 ind. (BMNH 10.11.24.2. [S], holotype of Rhinolophus hipposideros escalerae Andersen, 1918), Mogador, date and collector unlisted; — 2 ♀♀ (NMP 94519, 94520 [S+A*]), Takoumit, small cave, 26 April 2008, leg. P. Benda, J. Červený, A. Konečný and P. Vallo.

North Macedonia: 1 ♂ (NMP 96847 [S+A]), north-eastern bank of the Ohrid Lake, 10 July 1977, leg. V. Tauber.

Oman: 2 ♀♀ (NMP 93717 [S+A*], 93718 [A*]), Bani Habib, house, 28 March 2011, leg. P. Benda, A. Reiter and M. Uhrin; — 1 ♂ (NMP 93782 [S+A*]), Misfah, mosque, 9 April 2011, leg. P. Benda, A. Reiter and M. Uhrin; — 1 ♀ (NMP 93994 [S+A*]), Sal Alah, Birkat Khaldiyah, cistern, 13 March 2012, leg. P. Benda, A. Reiter and M. Uhrin.

Serbia: 1 ♂ (NMP 38955 [S+B]), Petnica, 23 May 1969, leg. J. Hanzák; — 1 ind. (NMP 96856 [S+B]), Serbia (undef.), May 1969, leg. J. Hanzák.

Slovakia: 5 ♂♂ (NMP 118/58, 121–123/58, 125/58, 130/58 [S]), Ardovo, Ardovská Cave, 5 February 1958, leg. V. Hanák; — 2 ♂♂ (NMP 84/63 [S+B], 85/63 [S]), Červený Kláštor, Aksamitka, 2 March 1963, leg. V. Hanák; — 1 ♂, 1 ♀ (NMP 101/58, 102/58 [S]), Domica, Čertova diera Cave, 5 February 1958, leg. V. Hanák; — 1 ♀ (NMP 7712/1957 [S+B]), Domica, Domica Cave, 24 August 1957, leg. J. Hanzák; — 5 ♂♂, 1 ♀(NMP 109/58–114/58 [S]), Domica, Liščia diera Cave, 5 February 1958, leg. V. Hanák; — 1 ♂ (NMP 154/58 [S]), Drienovec, cave, 6 February 1958, leg. V. Hanák; — 4 ♂♂, 2 ♀♀ (NMP J209–J213, J215 [S]), Gombasek, Ludmila Cave, 20 November 1955, 6 December 1956, 11 December 1956, leg. V. Hanák; — 1 ♂ (NMP 172/58 [S]), Hačava, Hačavská Cave, 7 February 1958, leg. V. Hanák; — 1 ♂ (NMP 7/69 [B]), Jasov, Jasovská Cave, 14 February 1969, leg. J. Gaisler; — 1 ♂, 2 ♀♀(NMP J185–J187 [S]), Kečovo, mine, 10 December 1956, leg. V. Hanák; — 3 ♂♂, 10 ♀♀ (NMP 160/61, 163/61, 181/61, 193/61, 194/61, 198/61, 200/61, 202/61, 204–206/61, 211/61, 212/61 [S]), Tisovec, Jaskyňa Netopierov Cave, 15 May 1961, 16 February 1961, leg. V. Hanák.

Sudan: 1 ind. (BMNH 47.5.27.48 [S]), Sennar, date and collector unlisted.

Syria: 3 ♀♀ (NMP 48054 [S+A*], 48055, 48056 [A*]), Qala'at Salah Ad Din, ruins, 30 June 1998, leg. M. Andreas and M. Uhrin; — 1 ♀ (NMP 48979 [S+A]), Qanawat, house, 27 April 2001, leg. P. Munclinger and P. Nová.

Tajikistan: 1 ♀ (NMP 95742 [S+A*]), Zingrogh, small cave, 12 May 2016, leg. P. Benda, A. Reiter and M. Uhrin.

Turkey: 1 ♀ (NMW 11731 [S+B]), 5 km W Igneada, Vil. Kirklareli, 15 May 1967, leg. F. Spitzenberger; — 1 ♂ (NMW 24585 [S+B]), Apollohöhle 2 km W Ahmetbeyli, Vil. Izmir, 16 February 1969, leg. F. Spitzenberger; — 2 ♀♀ (NMW 34330, 34331 [S+B]), Efes, Vil. Izmir, 2 August 1984, leg. A. Mayer, F. Spitzenberger and E. Weiss; — 1 ♂, 1 ♀ (NMW 22236, 22237 [S+B]), Ephesus, Westküste, 12 August 1976, leg. P. Wolff; — 1 ♂ (NMW 24587 [S+B]), Höhle Icme Pinari bei Arak, Vil. Isparta, 1 March 1969, leg. F. Spitzenberger; — 1 ind. (SMF 92191 [S]), Höhle Karain (Schwarze Höhle) und Höhle Oküzini (Ochsenhöhle), 450 m, 37.08N, 30.20E, Rand des Taurus-Gebirge an der Ebene von Antalya, 30 km NWN von Antalya, Vil. Antalya, 1990–1994, leg. P. Lacroix; — 2 ♂♂(NMW 24586, 24588 [S+B]), Höhlen NE Bornova, Vil. Izmir, 6 April 1969, leg. F. Spitzenberger; — 1 ♂ (NMW 13299 [S+A]), Maden köy, Vil. Nigde, 1 August 1970, leg. F. Spitzenberger; — 1 ♀ (CUP T93/63 [S+A]), Narlikuyu, 29 October 1993, leg. P. Benda and I. Horáček; — 3 ♀♀ (NMW 19313–19315 [S+A]), Nestorianische Kirche, Vil. Hakkari, 16 August 1973, leg. F. Spitzenberger; — 2 ♀♀ (NMP 90488, 90489 [S+A*]), Posyagbasan nr. Adana, 15 June 2003, leg. J. Hájek and J. Hotový; — 1 ♀ (NMW 20510 [B]), Rize, Vil. Rize, 14 July 1961, leg. M. Caglar; — 4 ♂♂ (MHNG 967.48, 967.49 [A], 967.50, 967.51 [S+A]), Satzmal magarasi, ouest de Sile, 29 April 1955, leg. H. Coiffat and P. Strinati; — 1 ♂ (MSNG 44534 [A]), Smirne, 1870, leg. G. Gonzenbach; — 1 ♂, 2 ♀♀ (CUP T93/65, T93/67, T93/68 [S+A]), Yalan Dünya Mağara Cave, 30 October 1993, leg. P. Benda and I. Horáček.

Turkmenistan: 1 ♂ (ZMMU S-169662 [A]), Aj-Derse, Kara-Kalpakskij District, May 1982, collector unlisted.

Ukraine: 1 ♀ (NMP pb4360 [S+A*]), Krym, General'skoe, 18 September 2009, leg. P. Benda, S. Gazarân and M. Uhrin; — 2 ♀♀ (NMP pb4287, pb4289 [S+A*]), Krym, Kujbyševo, 12 September 2009, leg. P. Benda, S. Gazarân and M. Uhrin; — 1 ♀ (NMP pb4342 [S+A*]), Krym, Partizanskoe, 16 September 2009, leg. P. Benda, S. Gazarân and M. Uhrin.

Uzbekistan: 1 ♂ (ZMMU S-13789 [S+B]), Nuratau, Pariš, 26 May 1934, leg. R. Meklenburcev.

Appendix II

Description of morphotypes (review of particular state conditions found in the examined sample sets)

Central Europe: body: large; skull: large in size, absolutely and relatively wide; braincase absolutely wide but relatively narrow, and absolutely and relatively high; tympanic bulla absolutely and relatively small; rostrum absolutely long but relatively short and narrow; teeth: upper canine (Cs) medium-sized; small upper premolar (P2) large; large upper premolar (P4) large, relatively narrow, and relatively medium-wide in its medial portion; first upper molar (M1) large and relatively wide; third upper molar (M3) small and relatively wide, very small in relation to M1; lower canine (Ci) is large, large in relation to the first lower molar (Mi); first lower premolar (P2) large, large in relation to the last lower premolar (P4); last lower premolar (P4) large; first lower molar (Mi) and the lower molar-row small.

West Mediterranean: body: small; skull: medium-sized in size, absolutely and relatively wide; braincase absolutely and relatively wide, and absolutely and relatively high; tympanic bulla absolutely and relatively small; rostrum absolutely short but relatively long, and relatively narrow; teeth: upper canine (Cs) small; small upper premolar (P2) medium-sized; large upper premolar (P4) medium-sized, relatively narrow, and relatively medium-wide in its medial portion; first upper molar (M1) medium-sized and relatively medium-wide; third upper molar (M3) small and relatively narrow, medium-sized in relation to M1; lower canine (Ci) small, large in relation to the first lower molar (Mi); first lower premolar (P2) large, large in relation to the last lower premolar (P4); last lower premolar (P4) small; first lower molar (Mi) and the lower molar-row small.

East Mediterranean: body: small; skull: small in size, absolutely narrow but relatively wide; braincase absolutely narrow but relatively wide, and absolutely and relatively high; tympanic bulla absolutely and relatively small; rostrum absolutely short but relatively long, and relatively narrow; teeth: upper canine (Cs) medium-sized; small upper premolar (P2) medium-sized; large upper premolar (P4) small and relatively wide, and relatively narrow in its medial portion; first upper molar (M1) medium-sized and relatively medium-wide; third upper molar (M3) small and relatively medium-wide, medium-sized in relation to M1; lower canine (Ci) small, small in relation to the first lower molar (Mi); first lower premolar (P2) large, large in relation to the last lower premolar (P4); last lower premolar (P4) small; first lower molar (M1) and the lower molar-row small.

Central Asia: body: large; skull: medium-sized in size, absolutely and relatively wide; braincase absolutely and relatively narrow, and absolutely and relatively high; tympanic bulla absolutely and relatively small; rostrum absolutely and relatively long, and relatively medium-sized in width; teeth: upper canine (Cs) medium-sized; small upper premolar (P2) medium-sized; large upper premolar (P4) medium-sized and relatively wide, relatively narrow in its medial portion; first upper molar (M1) large and relatively wide; third upper molar (M3) large and relatively narrow, medium-sized in relation to M1; lower canine (Ci) large, large in relation to the first lower molar; first lower premolar (P2) medium-sized, large in relation to the last lower premolar (P4); last lower premolar (P4) large; first lower molar (Mi) and the lower molar-row large.

North-eastern Africa: body: medium-sized; skull: small in size, absolutely narrow but relatively wide; braincase absolutely and relatively wide, and absolutely and relatively low; tympanic bulla absolutely and relatively small; rostrum absolutely short but relatively long, and relatively medium-sized in width; teeth: upper canine (Cs) small; small upper premolar (P2) medium-sized; large upper premolar (P4) small and relatively narrow, relatively wide in its medial portion; first upper molar (M1) small and relatively wide; third upper molar (M3) small and relatively medium-wide, medium-sized in relation to M1; lower canine (Ci) small, large in relation to the first lower molar (Mi); first lower premolar (P2) medium-sized, large in relation to the last lower premolar (P4); last lower premolar (P4) small; first lower molar (Mi) and the lower molar-row small.

Oman: body: small; skull: small in size, absolutely and relatively narrow; braincase absolutely and relatively narrow, and absolutely and relatively low; tympanic bulla absolutely and relatively large; rostrum absolutely and relatively long, and relatively very wide; teeth: upper canine (Cs) large; small upper premolar (P2) small; large upper premolar (P4) large and relatively narrow, relatively medium-wide in its medial portion; first upper molar (M1) medium-sized and relatively narrow; third upper molar (M3) large and relatively narrow, very large in relation to M1; lower canine (Ci) small, small in relation to the first lower molar (Mi); first lower premolar (P2) small, very small in relation to the last lower premolar (P4); last lower premolar (P4) large; first lower molar (Mi) and the lower molar-row are large.

© Museum and Institute of Zoology PAS
Petr Benda, Marek Uvizl, Peter Vallo, Antonín Reiter, and Marcel Uhrin "A Revision of the Rhinolophus hipposideros group (Chiroptera: Rhinolophidae) with Definition of an Additional Species from the Middle East," Acta Chiropterologica 24(2), 269-298, (2 February 2023). https://doi.org/10.3161/15081109ACC2022.24.2.001
Received: 8 September 2022; Accepted: 29 November 2022; Published: 2 February 2023
KEYWORDS
echolocation data
molecular analysis
Morphometrics
Rhinolophus
taxonomy
Back to Top