Next Article in Journal
Floral Development Stage-Specific Transcriptomic Analysis Reveals the Formation Mechanism of Different Shapes of Ray Florets in Chrysanthemum
Next Article in Special Issue
Being a Dog: A Review of the Domestication Process
Previous Article in Journal
Next-Generation Sequencing (NGS) Analysis Illustrates the Phenotypic Variability of Collagen Type IV Nephropathies
Previous Article in Special Issue
Polymorphisms and mRNA Expression Levels of IGF-1, FGF5, and KAP 1.4 in Tibetan Cashmere Goats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Genetic Characterization and Insular Habitat Enveloping of Endangered Leaf-Nosed Bat, Hipposideros nicobarulae (Mammalia: Chiroptera) in India: Phylogenetic Inference and Conservation Implication

1
Department of Marine Biology, Pukyong National University, Busan 48513, Republic of Korea
2
Western Ghat Regional Centre, Zoological Survey of India, Kozhikode 673006, India
3
Agricultural and Ecological Research Unit, Indian Statistical Institute, Kolkata 700108, India
4
Zoological Survey of India, M Block, New Alipore, Kolkata 700053, India
5
Marine Integrated Biomedical Technology Center, National Key Research Institutes in Universities, Pukyong National University, Busan 48513, Republic of Korea
*
Author to whom correspondence should be addressed.
Genes 2023, 14(3), 765; https://doi.org/10.3390/genes14030765
Submission received: 28 February 2023 / Revised: 16 March 2023 / Accepted: 20 March 2023 / Published: 21 March 2023

Abstract

:

Simple Summary

The present study provides the first genetic data on Hipposideros nicobarulae, an endangered and endemic leaf-nosed bat in the Nicobar Islands, India. Preliminary data sheds light on the phylogenetic relationship of H. nicobarulae and other closely related species (H. cf. antricola and H. cf. einnaythu) from South and Southeast Asia. Additionally, the distribution modelling provides information on suitable habitats for H. nicobarulae in the insular biogeography of Nicobar Island. Further integrated studies are needed to clarify the evolution, population genetics, distribution, and diversification of the Hipposideros species in the world.

Abstract

The Nicobar leaf-nosed Bat (Hipposideros nicobarulae) was described in the early 20th century; however, its systematic classification has been debated for over 100 years. This endangered and endemic species has achieved species status through morphological data in the last 10 years. However, the genetic information and phylogenetic relationships of H. nicobarulae remain neglected. The generated mitochondrial cytochrome b gene (mtCytb) sequences (438 bp) of H. nicobarulae contains 53.42–53.65% AT composition and 1.82% variable sites. The studied species, H. nicobarulae maintains an 8.1% to 22.6% genetic distance from other Hipposideros species. The genetic divergence estimated in this study is congruent with the concept of gene speciation in bats. The Bayesian and Maximum-Likelihood phylogenies clearly discriminated all Hipposideros species and showed a sister relationship between H. nicobarulae and H. cf. antricola. Current mtCytb-based investigations of H. nicobarulae have confirmed the species status at the molecular level. Further, the MaxEnt-based species distribution modelling illustrates the most suitable habitat of H. nicobarulae (294 km2), of which the majority (171 km2) is located on Great Nicobar Island. The present study suggests rigorous sampling across the range, taxonomic coverage, the generation of multiple molecular markers (mitochondrial and nuclear), as well as more ecological information, which will help in understanding population genetic structure, habitat suitability, and the implementation of appropriate conservation action plans for H. nicobarulae and other Hipposideros species.

1. Introduction

Bats (order Chiroptera) are the second most diverse group of mammals after Rodentia and currently contain approximately 1456 species worldwide [1]. Among them, more than 130 bat species under nine families have been reported from the Indian mainland and islands. The Palaeotropical genus Hipposideros consists of about 74 species worldwide, and 12 species are distributed in the Indian subcontinent [2]. Among them, four species (Hipposideros diadema, Hipposideros fulvus, Hipposideros grandis, and Hipposideros larvatus) are distributed in the Andaman group of islands, whereas three species (H. diadema, Hipposideros gentilis, and Hipposideros nicobarulae) are reported from the Nicobar group of islands, India [3,4,5,6,7]. To date, H. nicobarulae is only known from the type locality, Nicobar Island in the Andaman and Nicobar archipelago in India [3,4]. The Nicobar Leaf-nosed Bat, H. nicobarulae, was originally described in the early 20th century [8], and for over 100 years this bat species (or “this form”) was considered a subspecies of Hippisideros ater as H. ater nicobarulae [9,10]. However, based on its body size and external and dental morphology, this bat species (or ‘this form’) was elevated to a distinct species in the recent past [3]. Due to the lack of relevant material, the species status of H. nicobarulae is still controversial and has not yet been tested using genetic data. Hippisideros nicobarulae lives in a mixed-sex colony in caves, abandoned buildings, and tunnels as well [3,4]. Due to habitat destruction by tsunamis and other anthropogenic activities, the population size of H. nicobarulae is decreasing day by day [4]. Hence, the species is classified as “endangered” by the Bat Specialist Group of the IUCN Red List of Threatened Species [11].
With the rapid advancement of molecular tools, integrated methods incorporating partial mitochondrial and nuclear genes, complete genomes, and transcriptome data were used in an iterative manner to identify and elucidate the diversity of bats worldwide [12,13,14,15]. A number of molecular approaches have been published in Hipposideros bats and related taxa in various contexts: description of new species or subspecies [16], detection of cryptic species [17,18], phylogenetic and evolutionary significance [19,20], and phylogeographic studies [21,22]. Additionally, being an indicator species, many physiological studies on bats have also been accomplished around the world [23,24]. The assessment of the mitochondrial Cytochrome b (mtCytb) gene has proven useful for identification and genetic diversity estimation in mammals, including bats [25,26,27]. Nevertheless, the genetic information of H. nicobarulae is still not available to the scientific community. In this milieu, we examined bat specimens collected from the Nicobar Islands through an integrated approach to improve our understanding of the systematic position of this charismatic taxon. This preliminary data will enrich the global genetic library and encourage further phylogeographic research on endangered H. nicobarulae and other closely related species.
Additionally, apart from existing threats to H. nicobarulae, the insular island habitat automatically increases the vulnerability of this species. To protect any threatened species, it is important to know how different taxa respond to land use change and how habitat and climate change affect their persistence [28,29]. Several studies have already been conducted to achieve habitat distribution modelling of bats globally [30,31,32] and in India [33]. Due to the lack of sufficient knowledge on habitat fitness, the present study also attempts to perform species distribution modelling of H. nicobarulae to identify a suitable home range in the Nicobar Islands, which can be used for monitoring and strategic conservation of this species in the near future.

2. Materials and Methods

2.1. Ethics Statement and Sampling

A total of six individuals (four males and two females) of H. nicobarulae were collected from three different localities at Great Nicobar Island (6.83 N 93.86 E, 6.85 N 93.81 E, and 6.94 N 93.87 E) by using a mist net (Figure 1). The morphological examination was carried out at the Western Ghat Regional Centre (WGRC) of the Zoological Survey of India (ZSI), Kerala, India. The species was identified by the morphological characters described earlier [3]. A photograph of the live specimen was taken by M. Kamalakannan (Figure 1). The specimens were vouchered in 10% formaldehyde for long-term preservation. The voucher specimens were deposited in the National Zoological Collections of the WGRC, ZSI, under the museum catalogue numbers (ZSIWGRC-3630 to ZSIWGRC-3635). A meager amount of tissue sample was collected from the toe region of each sample for molecular experiments. The experimental procedures were approved by the Zoological Survey of India. No specific permission was required for sampling the targeted taxa.

2.2. Morphological Identification

The external morphology and cranial characteristics were measured for each sample of H. nicobarulae by using vernier calipers to the nearest 0.01 mm. To explain the species level identity and distinct specific status, the morphometrics data of closely related (H. ater and H. cf. einnaythu) and sympatric (H. diadema, H. fulvus, H. grandis, H. larvatus, and H. gentilis) species were acquired from the previous literature and compared [3,34,35,36] (Table 1). We could not obtain the taxonomic measurements of H. cf. antricola from any sources; hence, not incorporated here. The external measurements were performed according to the Bates and Harrison, 1997 [34]; length of head and body (HBL): from the tip of the snout to base of the tail on dorsal side, length of forearm (FA): from the extremity of the elbow to the extremity of the carpus with the wings folded, ear length (EL): from the lower border of the externally auditory meatus to the tip of the pinna, length of hindfoot (HF): from the extremity of the heel behind the Os calcis to the extremity of the longest digit, length of tibia (TIB): from the knee joint to the ankle, length of tail (TL): from the tip of the tail to its base adjacent to the body.

2.3. Mitochondrial DNA Extraction and Sequencing

The present study targeted the amplification of only a small fragment (>400 bp) of the mitochondrial Cytb gene, as it has been shown to be suitable for mammalian species identification [27,37]. Total genomic DNA was extracted by the standard phenol-chloroform isoamyl alcohol method and visualized in 1% agarose gel electrophoresis [38]. To amplify the partial mtCytb gene, mcb 398: 5′-TACCATGAGGACAAATATCATTCTG-3′ and mcb 869: 5′-CCTCCTAGTTTGTTAGGGATTGATCG-3′) primer pairs were used [37]. The 30 mL PCR mixture contains 10 pmol of forward and reverse primer, 20 ng of template DNA, 1X PCR buffer, 1.0–1.5 mM of MgCl2, 0.25 mM of each dNTPs, and 1U of Taq DNA polymerase (Invitrogen). The PCR was executed in a Veriti Thermal Cycler (Applied Biosystems) with the standard thermal settings. The PCR products were purified by using a QIAquick Gel Extraction Kit (QIAGEN). The cycle sequencing was performed using the BigDye Terminator ver. 3.1 Cycle Sequencing Kit (ABI) and 3.2 pmol of each primer. The PCR products were purified by the BigDye X-terminator kit (ABI) and bidirectionally sequenced by the genetic analyser housed at the Rajiv Gandhi Centre for Biotechnology (RGCB), Thiruvananthapuram, Kerala, India.

2.4. Sequence Annotation and Dataset Construction

To avoid the pseudogenes in the raw sequences, the noisy parts of each chromatogram were brought down at both ends through the SeqScanner Version 1.0 (Applied Biosystems). To validate the generated sequence, both nucleotide BLAST (https://blast.ncbi.nlm.nih.gov, accessed on 10 March 2023) and ORF finder (https://www.ncbi.nlm.nih.gov/orffinder/, accessed on 10 March 2023) were used to check the insertion-deletion and amino acid array for the vertebrate mitochondrial gene. The annotated sequences were submitted to GenBank’s global database through Bankit (https://www.ncbi.nlm.nih.gov/WebSub/, accessed on 10 March 2023). To estimate the genetic distances and perform the phylogenetic analysis, a total of 98 mtCytb sequences of Hipposideros species were acquired from NCBI GenBank. The mtCytb sequence of Rhinolophus rouxii (family Rhinolophidae) (accession no. JQ316214) was used as an out-group in the present phylogenetic analyses. The generated and database sequences were aligned through ClustalX software to build the final dataset [39].

2.5. Genetic Divergence and Phylogenetic Analyses

The Kimura-2-parameter (K2P) genetic divergence was estimated by MEGAX [40]. The best-fit model was estimated through JModelTest v2 with the lowest BIC (Bayesian Information Criterion) value [41]. The Bayesian (BA) topology was constructed in Mr. Bayes 3.1.2 by choosing nst = six for the GTR + G + I with one cold and three hot chains of metropolis-coupled Markov Chain Monte Carlo (MCMC); it was run for 10,000,000 generations with 25% burn-in and trees saving at every 100 generations [42]. The maximum-likelihood (ML) tree was constructed through the IQ-Tree web server with the GTR + G + I model and 1000 bootstrap samples [43]. The final topologies were illustrated in the web-based iTOL tool (https://itol.embl.de/, accessed on 10 March 2023) and edited with Adobe Photoshop CS 8.0 [44].

2.6. Species Occurrence Data

The occurrence records of H. nicobarulae were collected from available literature [3,4], and the GBIF online data repository (https://doi.org/10.15468/dl.d8ph4a, accessed on 10 March 2023) [45]. We started with a set of 21 habitat variables grouped into three types: climatic, land cover and land use (LULC), and topographic (Table S1). The climatic variables were represented by 19 bioclimatic variables from Worldclim, Ver: 2.0 (https://www.worldclim.org/, accessed on 10 March 2023) [46]. To examine the influence of individual LULC classes, land use and land cover derived from Copernicus Global Land Service (https://lcviewer.vito.be/download, accessed on 10 March 2023) were used. The elevation profile for the study landscape was generated using the 90-m Shuttle Radar Topography Mission (SRTM) data (http://srtm.csi.cgiar.org/srtmdata/, accessed on 10 March 2023). Finally, all predictors were resampled at 30 arc sec spatial resolution, and spatial multicollinearity among the variables were eliminated by using Pearson’s correlation with r > 0.8 in SDM Toolbox v2.4.

2.7. Model Building and Evaluation

For identifying suitable habitats for H. nicobarulae in the study landscape, we have attempted maximum entropy-based modelling using the software MaxEnt Ver. 3.4.4. The adopted method provides the probability of occurrence of a given species, ranked from 0 (least likely occurrence) to 1 (most likely occurrence). Our result generates a distribution probability as a probability raster surface ranging from a 0 to 1 value, where 0 is the most unsuitable zone and 1 is the most suitable zone. The relative influence of the selected variables was estimated by the Jackknife test of developed regularized training gain [47]. For model evaluation, the area under the curve statistics (AUC) of the receiver operating characteristic (ROC) curves with values ranging from 0 to 1, where a score of less than 0.5 is worse than random, followed by a score of 0.5 as considered a random prediction. A score between 0.7 and 0.8 is regarded as an acceptable model result, followed by a score between 0.8 and 0.9 as excellent, and <0.9 is regarded as an exceptional model [48]. Finally, we have prepared the binary maps from the continuous suitability raster based on an equal test sensitivity and specificity (SES) threshold.

3. Results

3.1. Taxonomic Identification

The collected specimens were identified as H. nicobarulae by the external morphometric characteristics: head and body length (HBL) ranges 41.6–47.9 mm; forearm length (FA): 38.1–42.2 mm; ear length 15.50–15.94 mm; hindfoot length: 7.67–8.05 mm; tibia length: 17.23–18.44 mm; tail length: 24.7–32.5 mm; and craniodental measurements: greatest length of the skull (GTL) ranges 16.68–17.23 mm: condylobasal length (CBL): 16.05–16.38 mm; condylocanine length (CCL): 15.07–15.42 mm; zygomatic breadth (ZB): 8.28–8.62 mm; breadth of the braincase (BB): 7.26–7.50 mm; maxillary toothrow length (C–M3): 5.69–5.74 mm; mandibular toothrow length (C–M3): 6.00–6.29 mm; posterior palatal width (M3–M3): 4.22–4.29 mm; anterior palatal width (C1–C1): 3.39–3.53 mm; mandibular length (M): 10.05–10.13 mm. The external and cranial measurements are comparatively larger than the H. ater (HBL: 38–48 mm; FA: 34–38 mm; GTL: 15.4–16.5 mm; CBL: 13.6–14.6 mm; CCL: 13.2–14.2 mm; M: 9.2–10.2 mm); the ears are large; the nose-leaf is slightly elongated and narrowed; there are no supplementary lateral leaflets; and the postero-lateral chambers of the rostrum are inflated but are significantly smaller than the H. ater (Figure 1 and Figure 2). The external morphological measurements of other closely related Hipposideros species, namely, H. einnaythu, H. fulvus, H. grandis, H. larvatus, and H. gentilis, overlap with H. nicobarulae except for H. diadema, H. grandis, H. gentilis, and H. larvatus, where these four species are significantly larger than H. nicobarulae (Table 1).

3.2. Molecular Identification

The partial cytochrome b analyses included 104 mtCytb sequences from 40 putative species. The generated sequences revealed 92.41% to 92.63% similarity with H. ater sequences (MT149731-MT149733) generated from the Philippines in the global nucleotide BLAST search. The mtCytb sequence (438 bp) contains an average of 53.55% AT composition in all Hipposideros species, with 53.42–53.65% in H. nicobarulae. A total of 49.54% variable sites were observed in the partial mtCytb sequences among all Hipposideros bats, while 1.82% variable sites were found in H. nicobarulae. The present dataset showed a 15.9% overall K2P mean genetic distance in all Hipposideros species. The interspecies genetic distances ranged from 4.1% (between H. grandis and H. larvatus) to 31% (H. grandis and H. fulvus). The studied species, H. nicobarulae maintains 8.1% to 22.6% genetic distance from other species in the present dataset (Table S2). Due to a lack of mtCytb data, genetic comparisons of H. nicobarulae with H. ater distributed in mainland India and Sri Lanka were not possible in this present analysis. However, H. nicobarulae showed 8.1% distance with H. ater (H. cf. antricola) generated from the Philippines, 13.5% distance with H. ater (H. cf. einnaythu) generated from Malaysia, and 12.4% distance with H. ater (hitherto unreported species) from Queensland, Australia. Besides, H. nicobarulae also maintained high genetic distances with other sympatric species (11.9% with H. gentilis, 16.8% with H. diadema, 18.2% with H. larvatus, 19.3% with H. grandis, and 22.6% with H. fulvus) distributed in the Andaman and Nicobar Islands (Table S2). In BA phylogeny, all Hipposideros species were clustered together (Figure 3). The ML tree also showed similar topology of the studied Hipposideros species (Figure S1). The studied species, H. nicobarulae, cohesively cladded with H. cf. antricola (distributed in the Philippines) and was recovered as a sister species. Further, the H. cf. einnaythu (distributed in Malaysia) showed close clustering with the H. nicobarulae + H. cf. antricola clades in the present BA phylogeny (Figure 3). As anticipated, the Hipposideros species were clearly discriminated by this fast-evolving mtCytb gene. We knew that the short sequence of a single mitochondrial gene does not clearly affect the understanding of phylogenetic relationships, but we were able to confirm the species-level genetic information of this endangered bat species, which will help for future phylogenetic, evolutionary, and population genetic studies.

3.3. Model Performance and Habitat Suitability

Our result has precisely predicted the suitable habitats within its extent distribution range. The average training AUC score for replicate runs (n = 50) for the model was found to be 0.861 ± 0.021 (SD) (Figure 4). Out of the total distribution range extent (1841 km2), only 294 km2 (15.96%) was found suitable for H. nicobarulae. The most suitable areas are within the southern range of Grater Nicobar (171 km2). Further, within the smaller islands, the most suitable and continuous habitat patches (40 km2) were distributed in the Car Nicobar. The result suggests that the suitable distribution for H. nicobarulae was greatly influenced by the precipitation of the wettest month, with a relative contribution of 25.80%, followed by the contribution of elevation at 15.80% (Figure 4). Further, the distribution probability concerning the elevation was found within the range of 0–300 m, after which the probability values were found to reduce substantially.

4. Discussion

The Hipposideros bats are distributed in the Palearctic, Afrotropical, Indo-Malayan, and Australasian regions and certainly contain many unresolved systematic issues. Some of them concern the taxonomic position of certain species that were not previously included in the relevant studies for various reasons (rareness in collections and in nature, limited range in areas inaccessible to researchers, etc.). Clarification of the taxonomic position is especially important for narrow-range endemics since it affects the development and implementation of conservation measures. Thus, taking into account the uncertainties with the taxonomy of H. nicobarulae, the present study elucidates an interesting systematic doubt both morphologically and genetically, as well as in terms of their habitat suitability.
The Great Nicobar Island is located about 170 km northwest of Sumatra, 350 km southeast of the Little Andaman Islands, 640 km off the southwest coast of Thailand, and 1306 km off the east coast of Sri Lanka. These islands are evidenced to be a discrete biogeographic unit and hold many magnetic faunal components, which act as a significant model for evolutionary studies [49]. During the Miocene–Pliocene, volcanic eruptions formed many isolated islands, and their sporadic connectivity was directly linked to changes in sea level during the Pleistocene, which accorded both geographical and temporal processes of species diversification in South and Southeast Asia [50]. The bathymetric study unwrapped the invisible connectivity of the Andaman and Nicobar archipelago and Sumatra by a well-developed seamount on the seafloor [51]. Moreover, the unparalleled biogeography of Southeast Asian oceanic islands provides a suitable habitat for many Hipposideros bats [52,53]. Looking at the biogeographic pattern, it is obvious that the species diversity of the Andaman and Nicobar archipelago would be more alike to the Indo-Malayan and Sundaic realms.
Considering the molecular-based species delimitation performed in the present mtCytb dataset, H. nicobarulae is genetically distinct and evolutionary close to its postulated sister species, H. cf. antricola, compared with other mainland species (Hipposideros pomona and Hipposideros durgadasi) of India. Previous studies have suggested that ≥8% genetic distance in the mitochondrial gene is pretty significant for inferring the genetic species concept in bats [26,54]. Similar genetic diversity has also been identified in Neotropical and Palearctic bats [55,56]. The targeted taxa (H. nicobarulae) maintain high genetic divergence (8.1% to 22.6%) with other Hipposideros bats, congruent with previous bat genetic speciation concepts. Further, the paraphyletic cladding of Hipposideros taxa from the African and Asian continents eliminates the biogeographical paradox and provides an interesting example of convergent evolution [57,58]. Moreover, the present integrated approach confirms the evolutionary placement of H. nicobarulae and is congruent with the prevailing hypothesis of Hipposideros classification. This cladding pattern and unique biogeography indicate that the endemic H. nicobarulae is not the result of local radiation, as might initially be assumed by H. ater sensu lato, but appears to have arisen from multiple independent colonization events. However, a more comprehensive study is required, including the generation of more molecular data (mitochondrial and nuclear) supporting many of the deeper nodes of phylogeny and their echolocation divergence, to understand their astonishing evolutionary pattern and diversification that may shed light on the speciation mechanisms of Hipposideros bats in their range distribution.
The mammalian fauna of the Great Nicobar Island comprises mostly bats and a few non-volant mammals. A total of 25 species of bats under 13 genera are found in the Andaman and Nicobar archipelago [4,34]. Among them, seven species under six genera were reported from the Great Nicobar Island (Cynopterus brachyotis, Hipposideros diadema, H. gentilis, H. nicobarulae, Myotis horsfieldii, Murina cyclotis, Pteropus melanotus, and Taphozous melanopogon). Apart from H. nicobarulae (endangered) and P. melanotus (vulnerable), other species are “least concern” in this insular habitat. It is proven that climate change and consequent natural calamities like tsunamis and earthquakes are gradually increasing the sea level of the Indian Ocean, which is greatly affecting the ecosystems of the Andaman and Nicobar Islands [59,60]. It is therefore imperative to select and conserve a minimum land area for biodiversity conservation that will also help conserve endemic and threatened species within an ecosystem [61,62]. Therefore, the conservation of the estimated suitable habitat (294 km2) is crucial for the conservation of the endemic H. nicobarulae population within this isolated insular habitat. Our predictions about habitat suitability will help governments and conservation organizations when formulating conservation policies for these endangered species. We also recommend further ecology studies of this species within its total range (1841 km2) to understand its present status and implement upcoming action plans for proper conservation.

5. Conclusions

The population of H. nicobarulae has declined significantly on Nicobar Island due to climate change and habitat degradation. No proper conservation action has been aimed to protect this species as its taxonomic status is still controversial in the scientific community. Although morphological data overlapped with some Hipposideros species, genetic data clearly discriminated most of the species and showed a matrilineal relationship with H. nicobarulae and H. cf. antricola in the present dataset. The study also delineated the suitable habitats and conservation priority areas of H. nicobarulae on Nicobar Island. We recommend an integrated approach, prioritizing extensive sampling, taxonomic coverage, multiple loci enrichment, and ecological modelling, to gain evolutionary knowledge and conserve this threatened bat species in the wild.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/genes14030765/s1, Figure S1. The mtCytb based ML phylogeny of the studied Hipposideros species; Table S1. Primary environmental and topographical variables used for ensemble modelling; Table S2. The K2P genetic divergences of the analysed of the studied Hipposideros species.

Author Contributions

Conceptualization: S.K. and H.-W.K.; Data curation: S.K., M.K. and T.M.; Formal analysis: S.K., M.K. and T.M.; Funding acquisition: D.B. and H.-W.K.; Investigation: S.K. and M.K.; Methodology: S.K., M.K. and T.M.; Project administration: D.B. and H.-W.K.; Resources: D.B. and H.-W.K.; Software: S.K., M.K. and T.M.; Supervision: D.B. and H.-W.K.; Validation: S.K., M.K. and T.M.; Visualization: S.K., D.B. and H.-W.K.; Writing—original draft: S.K., M.K. and T.M.; Writing—review & editing: S.K., M.K., T.M., D.B. and H.-W.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (2021R1A6A1A03039211), Korea Institute of Marine Science & Technology Promotion (KIMST) funded by the Ministry of Oceans and Fisheries (20220558), and core funding of Zoological Survey of India, Ministry of Environment, Forest and Climate Change (MoEF&CC), Govt. of India. The funders had no role in study design, data collection and analysis or preparation of the manuscript.

Institutional Review Board Statement

Prior permission was acquired from the Principal Chief Conservator of Forests (Wildlife), Andaman and Nicobar Islands, Port Blair (Letter no. CWLW/WL/24/339). The molecular data generation and analyses were approved by the host institutions (Pukyong National University, Zoological Survey of India, and Indian Statistical Institute).

Informed Consent Statement

Not applicable.

Data Availability Statement

The nucleotide sequence data that support the findings of this study are openly available in GenBank of NCBI at [https://www.ncbi.nlm.nih.gov, accessed on 10 March 2023] under the accession no. OP129713-OP129718.

Acknowledgments

The first author (S.K.) acknowledges the Global Postdoc Program fellowship grant received from the Pukyong National University, Republic of Korea. The second author (M.K.) thanks, V.D. Hegde, the Officer-in-charge of Western Ghat Regional Centre, Zoological Survey of India, Kozhikode, India for all the necessary facilities and encouragements. The third author (T.M) thanks, the Department of Science & Technology INSPIRE Faculty Award and Fellowship (Sanction no: DST/INSPIRE/04/2021/001149) of the Government of India. We are thankful to Pipat Soisook, Researcher of Princess Maha Chakri Sirindhorn Natural History Museum, Prince of Songkla University, Thailand for his comments and suggestions during manuscript preparation. We also thankful to C. Sivaperuman, the Officer-in-charge of Andaman and Nicobar Regional Centre, Zoological Survey of India, Port Blair, India and other team members for all the necessary facilities during the Great Nicobar expedition.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Simmons, N.B.; Cirranello, A.L. Bat Species of the World: A Taxonomic and Geographic Database. Available online: https://batnames.org (accessed on 5 February 2023).
  2. Burgin, C.J.; Colella, J.P.; Kahn, P.L.; Upham, N.S. How many species of mammals are there? J. Mammal. 2018, 99, 1–14. [Google Scholar] [CrossRef] [Green Version]
  3. Douangboubpha, B.; Bumrungsri, S.; Satasook, C.; Soisook, P.; HlaBu, S.S.; Aul, B.; Harrison, D.L.; Pearch, M.J.; Thomas, N.M.; Bates, P.J.J. A New Species of Small Hipposideros (Chiroptera: Hipposideridae) from Myanmar and a Revaluation of the Taxon H. nicobarulae Miller, 1902 from the Nicobar Islands. Acta Chiropterologica 2011, 13, 61–78. [Google Scholar] [CrossRef]
  4. Aul, B.; Bates, P.J.J.; Harrison, D.L.; Marimuthu, G. Diversity, distribution and status of bats on the Andaman and Nicobar Islands, India. Oryx 2014, 48, 204–214. [Google Scholar] [CrossRef] [Green Version]
  5. Srinivasulu, B.; Srinivasulu, A.; Srinivasulu, C.; Dar, T.H.; Gopi, A.; Jones, G. First record of the diadem leaf-nosed bat Hipposideros diadema (E. Geoffroy, 1813) (Chiroptera, Hipposideridae) from the Andaman Islands, India with the possible occurrence of a hitherto unreported subspecies. J. Threat. Taxa 2016, 8, 9316–9321. [Google Scholar] [CrossRef]
  6. Srinivasulu, C.; Srinivasulu, A.; Srinivasulu, B.; Gopi, A.; Dar, T.H.; Bates, P.J.J.; Rossiter, S.J.; Jones, G. Recent surveys of bats from the Andaman Islands, India: Diversity, distribution, and echolocation characteristics. Acta Chiropterologica 2017, 19, 419–437. [Google Scholar] [CrossRef]
  7. Srinivasulu, B.; Srinivasulu, C. In plain sight: Bacular and noseleaf morphology supports distinct specific status of Roundleaf Bats Hipposideros pomona Andersen, 1918 and Hipposideros gentilis Andersen, 1918 (Chiroptera: Hipposideridae). J. Threat. Taxa 2018, 10, 12018–12026. [Google Scholar] [CrossRef]
  8. Miller, G.S. The mammals of the Andaman and Nicobar Islands. Proc. U. S. Natl. Mus. 1902, 24, 751–795. [Google Scholar] [CrossRef] [Green Version]
  9. Hill, J.E. A revision of the genus Hipposideros. Bull. Br. Mus. Nat. Hist. Zool. 1963, 2, 1–129. [Google Scholar] [CrossRef]
  10. Simmons, N.B. Order Chiroptera. In Mammal Species of the World: A Taxonomic and Geographic Reference; Wilson, D.E., Reeder, D.M., Eds.; The Johns Hopkins University Press: Baltimore, MD, USA, 2005; pp. 312–529. [Google Scholar]
  11. Srinivasulu, B.; Srinivasulu, C.; Waldien, D.L. Hipposideros nicobarulae. In The IUCN Red List of Threatened Species 2020: E.T80458824A95642215; Cambridge, UK, 2020; Available online: https://doi.org/10.2305/IUCN.UK.2020-3.RLTS.T80458824A95642215.en (accessed on 10 March 2023).
  12. Teeling, E.C.; Springer, M.S.; Madsen, O.; Bates, P.; O’brien, S.J.; Murphy, W.J. A molecular phylogeny of bats illuminates biogeography and fossil record. Science 2005, 307, 580–584. [Google Scholar] [CrossRef]
  13. Esselstyn, J.A.; Evans, B.J.; Sedlock, J.L.; Anwarali Khan, F.A.; Heaney, L.R. Single-locus species delimitation: A test of the mixed Yule-coalescent model, with an empirical application to Philippine round-leaf bats. Proc. Biol. Sci. 2012, 279, 3678–3686. [Google Scholar] [CrossRef]
  14. Hawkins, J.A.; Kaczmarek, M.E.; Müller, M.A.; Drosten, C.; Press, W.H.; Sawyer, S.L. A metaanalysis of bat phylogenetics and positive selection based on genomes and transcriptomes from 18 species. Proc. Natl. Acad. Sci. USA 2019, 116, 11351–11360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Jebb, D.; Huang, Z.; Pippel, M.; Hughes, G.M.; Lavrichenko, K.; Devanna, P.; Winkler, S.; Jermiin, L.S.; Skirmuntt, E.C.; Katzourakis, A.; et al. Six reference-quality genomes reveal evolution of bat adaptations. Nature 2020, 583, 578–584. [Google Scholar] [CrossRef]
  16. Thong, V.D.; Puechmaille, S.J.; Denzinger, A.; Dietz, C.; Csorba, G.; Bates, P.J.J.; Teeling, E.C.; Schnitzler, H.-U. A new species of Hipposideros (Chiroptera: Hipposideridae) from Vietnam. J. Mammal. 2012, 93, 1–11. [Google Scholar] [CrossRef] [Green Version]
  17. Murray, S.W.; Campbell, P.; Kingston, T.; Zubaid, A.; Francis, C.M.; Kunz, T.H. Molecular phylogeny of hipposiderid bats from Southeast Asia and evidence of cryptic diversity. Mol. Phylogenet. Evol. 2012, 62, 597–611. [Google Scholar] [CrossRef]
  18. Rakotoarivelo, A.R.; Willows-Munro, S.; Schoeman, M.C.; Lamb, J.M.; Goodman, S.M. Cryptic diversity in Hipposideros commersoni sensu stricto (Chiroptera: Hipposideridae) in the western portion of Madagascar. BMC Evol. Biol. 2015, 15, 235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Lavery, T.H.; Leung, L.K.P.; Seddon, J.M. Molecular phylogeny of hipposiderid bats (Chiroptera: Hipposideridae) from Solomon Islands and Cape York Peninsula, Australia. Zool. Scr. 2013, 43, 429–442. [Google Scholar]
  20. Foley, N.; Thong, V.D.; Soisook, P.; Goodman, S.; Armstrong, K.; Jacobs, D.; Puechmaille, S.J.; Teeling, E. How and why overcome the impediments to resolution: Lessons from rhinolophid and hipposiderid bats. Mol. Biol. Evol. 2014, 33, 313–333. [Google Scholar] [CrossRef] [Green Version]
  21. Xu, L.; He, C.; Shen, C.; Jiang, T.; Shi, L.; Sun, K.; Berquist, S.W.; Feng, J. Phylogeography and population genetic structure of the great leaf-nosed bat (Hipposideros armiger) in China. J. Hered. 2010, 101, 562–572. [Google Scholar] [CrossRef] [Green Version]
  22. Lin, A.Q.; Csorba, G.; Li, L.F.; Jiang, T.L.; Lu, G.J.; Thong, V.D.; Soisook, P.; Sun, K.P.; Feng, J. Phylogeography of Hipposideros armiger (Chiroptera: Hipposideridae) in the Oriental Region: The contribution of multiple Pleistocene glacial refugia and intrinsic factors to contemporary population genetic structure. J. Biogeogr. 2013, 41, 317–327. [Google Scholar] [CrossRef]
  23. Puechmaille, S.J.; Gouilh, M.A.; Piyapan, P.; Yokubol, M.; Mie, K.M.; Bates, P.J.; Satasook, C.; New, T.; Bu, S.S.; Mackie, I.J.; et al. The evolution of sensory divergence in the context of limited gene flow in the bumblebee bat. Nat. Commun. 2011, 2, 573. [Google Scholar] [CrossRef] [Green Version]
  24. Teeling, E.C.; Scally, M.; Kao, D.J.; Romagnoli, M.L.; Springer, M.S.; Stanhope, M.J. Molecular evidence regarding the origin of echolocation and flight in bats. Nature 2000, 403, 188–192. [Google Scholar] [CrossRef]
  25. Johns, G.C.; Avise, J.C. A comparative summary of genetic distances in the vertebrates from the mitochondrial cytochrome b gene. Mol. Biol. Evol. 1998, 15, 1481–1490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Bradley, R.D.; Baker, R.J. A Test of the Genetic Species Concept: Cytochrome-b Sequences and Mammals. J. Mammal. 2001, 82, 960–973. [Google Scholar] [CrossRef]
  27. Tobe, S.S.; Kitchener, A.C.; Linacre, A.M.T. Reconstructing Mammalian Phylogenies: A Detailed Comparison of the Cytochrome b and Cytochrome Oxidase Subunit I Mitochondrial Genes. PLoS ONE 2010, 5, e14156. [Google Scholar] [CrossRef] [Green Version]
  28. Walther, G.R.; Post, E.; Convey, P.; Menzel, A.; Parmesan, C.; Beebee, T.J.C.; Fromentin, J.M.; Hoegh-Guldberg, O.; Bairlein, F. Ecological responses to recent climate change. Nature 2002, 416, 389–395. [Google Scholar] [CrossRef]
  29. Hughes, A.C.; Satasook, C.; Bates, P.J.; Bumrungsri, S.; Jones, G. The projected effects of climatic and vegetation changes on the distribution and diversity of Southeast Asian bats. Glob. Chang. Biol. 2012, 18, 1854–1865. [Google Scholar] [CrossRef]
  30. Wordley, C.F.R.; Sankaran, M.; Mudappa, D.; Altringham, J.D. Landscape scale habitat suitability modelling of bats in the Western Ghats of India: Bats like something in their tea. Biol. Conserv. 2015, 191, 529–536. [Google Scholar] [CrossRef] [Green Version]
  31. Herkt, K.M.B.; Barnikel, G.; Skidmore, A.K.; Fahr, J. A high-resolution model of bat diversity and endemism for continental Africa. Ecol. Model. 2016, 320, 9–28. [Google Scholar] [CrossRef]
  32. Bandara, A.P.M.J.; Madurapperuma, B.D.; Edirisinghe, G.; Gabadage, D.; Botejue, M.; Surasinghe, T.D. Bioclimatic Envelopes for Two Bat Species from a Tropical Island: Insights on Current and Future Distribution from Ecological Niche Modeling. Diversity 2022, 14, 506. [Google Scholar] [CrossRef]
  33. Raman, S.; Shameer, T.T.; Pooja, U.; Hughes, A.C. Identifying priority areas for bat conservation in the Western Ghats Mountain range, peninsular India. J. Mammal. 2022, gyac060. [Google Scholar] [CrossRef]
  34. Bates, P.J.J.; Harrison, D.L. Bats of Indian Sub-Continent; Harrison Zoological Museum: Sevenoaks, UK, 1997; p. xv-258. [Google Scholar]
  35. Sinha, Y.P. Contribution to the Knowledge of Bats of North East Hills India; Occasional Paper No. 174; Zoological Survey of India: Kolakata, India, 1999; pp. 1–52. [Google Scholar]
  36. Thong, V.D.; Denzinger, A.; Sang, N.V.; Huyen, N.T.T.; Thanh, H.T.; Loi, D.N.; Nha, P.V.; Viet, N.V.; Tien, P.D.; Tuanmu, M.-N.; et al. Bat Diversity in Cat Ba Biosphere Reserve, Northeastern Vietnam: A Review with New Records from Mangrove Ecosystem. Diversity 2021, 13, 376. [Google Scholar] [CrossRef]
  37. Verma, S.K.; Singh, L. Novel universal primers establish identity of an enormous number of animal species for forensic application. Mol. Ecol. Notes 2002, 3, 28–31. [Google Scholar] [CrossRef]
  38. Sambrook, J.; Russell, D.W. Molecular Cloning: A Laboratory Manual, 3rd ed.; Cold Spring Harbor Laboratory Press: New York, NY, USA, 2001; Volume 1. [Google Scholar]
  39. Thompson, J.D.; Gibson, T.J.; Plewniak, F.; Jeanmougin, F.; Higgins, D.G. The CLUSTAL_X windows interface: Flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Res. 1997, 25, 4876–4882. [Google Scholar] [CrossRef] [Green Version]
  40. Kumar, S.; Stecher, G.; Li, M.; Knyaz, C.; Tamura, K. MEGA X: Molecular Evolutionary Genetics Analysis across Computing Platforms. Mol. Biol. Evol. 2018, 35, 1547–1549. [Google Scholar] [CrossRef] [PubMed]
  41. Darriba, D.; Taboada, G.L.; Doallo, R.; Posada, D. JModelTest 2: More models, new heuristics and parallel computing. Nat. Methods 2012, 9, 772. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Ronquist, F.; Huelsenbeck, J.P. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 2003, 19, 1572–1574. [Google Scholar] [CrossRef] [Green Version]
  43. Trifinopoulos, J.; Nguyen, L.-T.; von Haeseler, A.; Minh, B.Q. W-IQ-TREE: A fast online phylogenetic tool for maximum likelihood analysis. Nucleic Acids Res. 2016, 44, W232–W235. [Google Scholar] [CrossRef] [Green Version]
  44. Letunic, I.; Bork, P. Interactive tree of life (iTOL) v4: Recent updates and new developments. Nucleic Acids Res. 2019, 47, W256–W259. [Google Scholar] [CrossRef] [Green Version]
  45. GBIF. Free and Open Access to Biodiversity Data. Available online: https://doi.org/10.15468/dl.d8ph4a (accessed on 5 February 2023).
  46. Fick, S.E.; Hijmans, R.J. Worldclim 2: New 1-km spatial resolution climate surfaces for global land areas. Int. J. Climatol. 2017, 37, 4302–4315. [Google Scholar] [CrossRef]
  47. Phillips, S.J.; Dudık, M. Modeling of species distributions with MaxEnt: New extensions and a comprehensive evaluation. Ecography 2008, 31, 161–175. [Google Scholar] [CrossRef]
  48. Halvorsen, R.; Mazzoni, S.; Dirksen, J.W.; Næsset, E.; Gobakken, T.; Ohlson, M. How important are choice of model selection method and spatial autocorrelation of presence data for distribution modelling by MaxEnt? Ecol. Model. 2016, 328, 108–118. [Google Scholar] [CrossRef]
  49. Kamalakannan, M.; Sivaperuman, C.; Kundu, S.; Gokulakrishnan, G.; Venkatraman, C.; Chandra, K. Discovery of a new mammal species (Soricidae: Eulipotyphla) from Narcondam volcanic island, India. Sci. Rep. 2021, 11, 9416. [Google Scholar] [CrossRef] [PubMed]
  50. Voris, H.K. Maps of Pleistocene Sea levels in Southeast Asia: Shorelines, river systems and time durations. J. Biogeogr. 2000, 27, 1153–1167. [Google Scholar] [CrossRef] [Green Version]
  51. Tripathi, S.K.; Nagendran, G.; Karthikeyan, M.; Vasu, P.; Tripathi, S.K.; Varghese, S.; Raghav, S.; Resmi, S. Morphology of submarine volcanic seamounts from inner volcanic arc of Andaman Sea. Indian J. Geosci. 2017, 71, 451–470. [Google Scholar]
  52. Robinson, M.F.; Jenkins, P.D.; Francis, C.M.; Fulford, A.J.C. A new species of the Hipposideros pratti group (Chiroptera, Hipposideridae) from Lao PDR and Vietnam. Acta Chiropterologica 2003, 5, 31–48. [Google Scholar] [CrossRef]
  53. Bates, P.J.J.; Rossiter, S.J.; Suyanto, A.; Kingston, T. A new species of Hipposideros (Chiroptera: Hipposideridae) from Sulawesi. Acta Chiropterologica 2007, 9, 13–26. [Google Scholar] [CrossRef]
  54. Baker, R.J.; Bradley, R.D. Speciation in Mammals and the Genetic Species Concept. J. Mammal. 2006, 87, 643–662. [Google Scholar] [CrossRef]
  55. Clare, E.L.; Lim, B.K.; Fenton, M.B.; Hebert, P.D.N. Neotropical Bats: Estimating Species Diversity with DNA Barcodes. PLoS ONE 2011, 6, e22648. [Google Scholar] [CrossRef] [PubMed]
  56. Li, H.; Li, Y.; Motokawa, M.; Wu, Y.; Harada, M.; Li, Y. The Effectiveness of Molecular, Karyotype and Morphological Methods in the Identification of Morphologically Conservative Sibling Species: An Integrative Taxonomic Case of the Crocidura attenuata Species Complex in Mainland China. Animals 2023, 13, 643. [Google Scholar] [CrossRef] [PubMed]
  57. Teeling, E.C.; Madsen, O.; Van den Bussche, R.A.; de Jong, W.W.; Stanhope, M.J.; Springer, M.S. Microbat paraphyly and the convergent evolution of a key innovation in Old World rhinolophoid microbats. Proc. Natl. Acad. Sci. USA 2002, 99, 1432–1436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Guillén-Servent, A.; Francis, C.M. A new species of bat of the Hipposideros bicolor group (Chiroptera: Hipposideridae) from Central Laos, with evidence of convergent evolution with Sundaic taxa. Acta Chiropterologica 2006, 8, 39–61. [Google Scholar] [CrossRef]
  59. Malik, J.N.; Johnson, F.C.; Khan, A.; Sahoo, S.; Irshad, R.; Paul, D.; Arora, S.; Baghel, P.K.; Chopra, S. Tsunami records of the last 8000 years in the Andaman Island, India, from mega and large earthquakes: Insights on recurrence interval. Sci Rep. 2019, 9, 18463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Mageswaran, T.; Sachithanandam, V.; Sridhar, R.; Mahapatra, M.; Purvaja, R.; Ramesh, R. Impact of sea level rise and shoreline changes in the tropical island ecosystem of Andaman and Nicobar region, India. Nat. Hazards 2021, 109, 1717–1741. [Google Scholar] [CrossRef]
  61. Brooks, T.M.; Mittermeier, R.A.; da Fonseca, G.A.; Gerlach, J.; Hoffmann, M.; Lamoreux, J.F.; Mittermeier, C.G.; Pilgrim, J.D.; Rodrigues, A.S. Global biodiversity conservation priorities. Science 2006, 313, 58–61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Allan, J.R.; Possingham, H.P.; Atkinson, S.C.; Waldron, A.; Di Marco, M.; Butchart, S.H.M.; Adams, V.M.; Kissling, W.D.; Worsdell, T.; Sandbrook, C.; et al. The minimum land area requiring conservation attention to safeguard biodiversity. Science 2022, 376, 1094–1101. [Google Scholar] [CrossRef]
Figure 1. Map showing the collection localities of H. nicobarulae in the Great Nicobar Island (marked by a red pin). (A) A live specimen of H. nicobarulae captured by M. Kamalakannan during the field survey in the Andaman and Nicobar archipelago; (B) a frontal view of the ears and nose leaves of H. nicobarulae; (C,D) the habitat of H. nicobarulae on the Great Nicobar Island, India.
Figure 1. Map showing the collection localities of H. nicobarulae in the Great Nicobar Island (marked by a red pin). (A) A live specimen of H. nicobarulae captured by M. Kamalakannan during the field survey in the Andaman and Nicobar archipelago; (B) a frontal view of the ears and nose leaves of H. nicobarulae; (C,D) the habitat of H. nicobarulae on the Great Nicobar Island, India.
Genes 14 00765 g001
Figure 2. Craniodental characters (skull and mandible) of H. nicobarulae: (i) lateral view of maxilla; (ii) lateral-ventral view of mandible; (iii) dorsal view of maxilla; (iv) ventral view of maxilla; and (v) occlusal view of mandible.
Figure 2. Craniodental characters (skull and mandible) of H. nicobarulae: (i) lateral view of maxilla; (ii) lateral-ventral view of mandible; (iii) dorsal view of maxilla; (iv) ventral view of maxilla; and (v) occlusal view of mandible.
Genes 14 00765 g002
Figure 3. The BA phylogeny based on the mtCytb gene clearly discriminates all Hipposideros species in this present study. The Genebank accession number is indicated in parentheses along with the species name. The posterior probability support values are mentioned with each node. The relationship between H. nicobarulae and closely related species is mentioned in different colors.
Figure 3. The BA phylogeny based on the mtCytb gene clearly discriminates all Hipposideros species in this present study. The Genebank accession number is indicated in parentheses along with the species name. The posterior probability support values are mentioned with each node. The relationship between H. nicobarulae and closely related species is mentioned in different colors.
Genes 14 00765 g003
Figure 4. (A) The distribution pattern of H. nicobarulae and other closely related species (H. cf. einnaythu, and H. cf. antricola), elevated as a distinct species from H. ater, (B) Representing the probability of suitable habitats of H. nicobarulae in its extent distribution range, (C) Representing the distribution probability in the northern range within Car Nicobar, Terresa, Kamorta, Katchal and Nancowry Islands, (D) Representing the distribution probability in the southern range within Little Nicobar and Great Nicobar. (E) Jackknife test for all the eleven variables (F) Response curves of the important variables for habitat suitability of H. nicobarulae, (G) Percentage contribution is represented by a column graph (color ramp = percentage contribution), and permutation importance is represented by the circular plot (size = permutation importance). (H) The average training ROC for the final model replicates.
Figure 4. (A) The distribution pattern of H. nicobarulae and other closely related species (H. cf. einnaythu, and H. cf. antricola), elevated as a distinct species from H. ater, (B) Representing the probability of suitable habitats of H. nicobarulae in its extent distribution range, (C) Representing the distribution probability in the northern range within Car Nicobar, Terresa, Kamorta, Katchal and Nancowry Islands, (D) Representing the distribution probability in the southern range within Little Nicobar and Great Nicobar. (E) Jackknife test for all the eleven variables (F) Response curves of the important variables for habitat suitability of H. nicobarulae, (G) Percentage contribution is represented by a column graph (color ramp = percentage contribution), and permutation importance is represented by the circular plot (size = permutation importance). (H) The average training ROC for the final model replicates.
Genes 14 00765 g004
Table 1. External morphological measurements (mm) of Hipposideros spp. [3,34,35,36]. The measurements are represented in range (minimum and maximum). HBL: head and body length; FA: forearm length; EL: ear length; HF: hind foot length; Tib: tibia length; T: tail length.
Table 1. External morphological measurements (mm) of Hipposideros spp. [3,34,35,36]. The measurements are represented in range (minimum and maximum). HBL: head and body length; FA: forearm length; EL: ear length; HF: hind foot length; Tib: tibia length; T: tail length.
SpeciesHBLFAELHFTibT
H. nicobarulae (n = 6)41.6–47.938.1–42.215.5–15.97.67–8.0517.2–18.424.7–32.5
H. ater (n = 11)38.0–48.034.0–38.014.8–20.05.30–7.2015.2–17.820.0–30.0
H. einnaythu (n = 2)43.3, 49.139.5, 40.316.6, 16.76.5, 7.016.2, 17.824.7, 28.7
H. diadema (n = 6)-80.2–90.223.3–30.813.67–18.129.7–37.844.3–57.2
H. fulvus (n = 35)40.0–50.038.4–44.019.0–26.06.0–9.816.5–20.724.0–35.0
H. grandis (n = 12)70.0–79.060.0–64.021.5–24.012.0–13.521.0–25.734.0–39.0
H. larvatus (n = 8)-48.7–59.014.4–23.53.9–6.619.0–20.124.8–35.5
H. gentilis (n = 9)36.0–52.039.5–43.222.0–25.06.3–8.518.2–19.128.0–35.0
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kundu, S.; Kamalakannan, M.; Mukherjee, T.; Banerjee, D.; Kim, H.-W. Genetic Characterization and Insular Habitat Enveloping of Endangered Leaf-Nosed Bat, Hipposideros nicobarulae (Mammalia: Chiroptera) in India: Phylogenetic Inference and Conservation Implication. Genes 2023, 14, 765. https://doi.org/10.3390/genes14030765

AMA Style

Kundu S, Kamalakannan M, Mukherjee T, Banerjee D, Kim H-W. Genetic Characterization and Insular Habitat Enveloping of Endangered Leaf-Nosed Bat, Hipposideros nicobarulae (Mammalia: Chiroptera) in India: Phylogenetic Inference and Conservation Implication. Genes. 2023; 14(3):765. https://doi.org/10.3390/genes14030765

Chicago/Turabian Style

Kundu, Shantanu, Manokaran Kamalakannan, Tanoy Mukherjee, Dhriti Banerjee, and Hyun-Woo Kim. 2023. "Genetic Characterization and Insular Habitat Enveloping of Endangered Leaf-Nosed Bat, Hipposideros nicobarulae (Mammalia: Chiroptera) in India: Phylogenetic Inference and Conservation Implication" Genes 14, no. 3: 765. https://doi.org/10.3390/genes14030765

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

Article Metrics

Back to TopTop