Skip to main content

The natural and human-mediated expansion of a human-commensal lizard into the fringes of Southeast Asia

Abstract

Background

Human-commensal species often display deep ancestral genetic structure within their native range and founder-effects and/or evidence of multiple introductions and admixture in newly established areas. We investigated the phylogeography of Eutropis multifasciata, an abundant human-commensal scincid lizard that occurs across Southeast Asia, to determine the extent of its native range and to assess the sources and signatures of human introduction outside of the native range. We sequenced over 350 samples of E. multifasciata for the mitochondrial ND2 gene and reanalyzed a previous RADseq population genetic dataset in a phylogenetic framework.

Results

Nuclear and mitochondrial trees are concordant and show that E. multifasciata has retained high levels of genetic structure across Southeast Asia despite being frequently moved by humans. Lineage boundaries in the native range roughly correspond to several major biogeographic barriers, including Wallace’s Line and the Isthmus of Kra. Islands at the outer fringe of the range show evidence of founder-effects and multiple introductions.

Conclusions

Most of enormous range of E. multifasciata across Southeast Asia is native and it only displays signs of human-introduction or recent expansion along the eastern and northern fringe of its range. There were at least three events of human-introductions to Taiwan and offshore islands, and several oceanic islands in eastern Indonesia show a similar pattern. In Myanmar and Hainan, there is a founder-effect consistent with post-warming expansion after the last glacial maxima or human introduction.

Peer Review reports

Introduction

Determining the extent of a species’ native versus introduced range is often difficult due to lack of records and extreme human modification to landscapes and transportation [1, 2]. Of these so-called cryptogenic species, widespread human-commensals that are good at dispersing both naturally and through human-movements are particularly difficult to decipher [3]. As the introduction of these species can lead to cascading ecological consequences, understanding the native range of species is important to conservation [4]. Phylogeographic and population genetic methods are critical tools to assessing if populations are native versus introduced, but they require comprehensive sampling of range-wide genetic diversity, and different invasion scenarios can lead to very different resulting patterns [5,6,7]. While single introductions to new areas show a classical founder-effect of extremely low genetic diversity in the new populations, multiple introductions are likely to instead produce high genetic diversity with divergent mitochondrial haplotypes in close proximity and admixture in nuclear DNA [5, 8,9,10,11]. Gene flow between introduced divergent lineages can restore genetic diversity to bottlenecked populations and provide the basis for adaptation that can worsen invasions [12,13,14]. Inside the native range, however, genetic structure is often preserved due to a combination of evolutionary and ecological processes [15,16,17]. Together these patterns can be used to understand the geographic extent of the native range and important biogeographic forces within it, and to pinpoint introduced populations.

The Many-lined Sun Skink, Eutropis multifasciata (Kuhl, 1820), is perhaps the most-commonly encountered diurnal lizard in Southeast Asia and possesses an incredibly large range from Bangladesh [18] through Indonesia [19]. It is an abundant inhabitant of urban areas, human-settlements, and forest edges, and is found from sea level to 1800 m in tropical areas [20] and 500 m at its northern extent in Taiwan [21]. Due to its abundance and relatively simple maintenance in a lab, it has become a model system for studying niche plasticity and thermal physiology, including its relationship with viviparity [21,22,23,24,25,26,27]. There are documented introductions to Florida [28, 29], Australia [30], New Guinea [31], and Taiwan (including nearby Ludao and Lanyu) [32, 33] due to frequent human-transportation including reports on airport luggage [34], ferries (J. McGuire, pers. obsv.), and international cargo shipments [35]. On Taiwan, Ludao, and Lanyu, E. multifasciata rapidly invaded and poses a significant conservation threat [32, 36, 37]. We used a geographically comprehensive genetic dataset of E. multifasciata to investigate its biogeographic history, the extent of its invasion, and the processes causing phylogeographic structure to be maintained or eroded. Together, these results can be utilized for conservation management in areas that have been recently invaded.

The phylogeography of E. multifasciata has been studied to some extent but still remains poorly understood. Barley et al. [38] investigated the population genomics of E. multifasciata with RADseq data, demonstrating that E. multifasciata has moderate levels of genome-wide genetic variation. Barley et al. observed higher genetic differentiation on islands relative to the mainland, recovering three Philippine lineages (southern; central/west; northern) and two weakly-differentiated continental populations (Myanmar; remainder of Indochina). They included only a few samples from Sundaland and Wallacea, limiting the scope of the phylogeographic analysis. The Philippine populations could be allied with Borneo, Sulawesi, or Indochina, or divided indicating multiple invasions of the Philippines [39]. Due to its random nature, RADseq data may encompass signatures of local adaptation, and this may be relevant when introduced populations have divergent environmental conditions to the source. For our study, the use of mitochondrial data may provide unique power to elucidate spatial patterns. The genome-wide data provides an interesting backdrop for our study, allowing us to understand the fine-scale mitochondrial structure we uncover given rates of nuclear gene flow estimated by Barley et al.

Since E. multifasciata is found in four major biogeographic regions: Indochina, Sundaland, the Philippines, and Wallacea [40], it is an interesting system to investigate how it is impacted by barriers that restrict many other species. We investigate if these regions are part of the native range, and if so, is there genetic structuring associated with the common biogeographic barriers that separate them? Mabuyine skinks are excellent dispersers that have crossed expansive marine barriers to distant islands [41,42,43], leaving the relative roles of natural vs. human-mediated dispersal in E. multifasciata an open question. An early phylogenetic study that included a few E. multifasciata samples recovered two major clades: one from Borneo, Myanmar, and China, and another from Java, Bali, Seram, and the Philippines [44], providing some indication that clades in E. multifasciata may not follow standard biogeographic area predictions (i.e. the Sunda Shelf not forming a clade). In another study, genetic structure across the Lesser Sunda Archipelago was substantial enough to reflect natural dispersal, and a clade on Bali and Timor was divergent from the remaining Lesser Sundas [19].

With comprehensive geographic sampling from over 350 samples, we investigated the biogeographic history of E. multifasciata to understand the extent of its native range and assessed the sources and signatures of human introduction outside the native range. We found it possesses a large native range where it has retained substantial genetic structure and test if this structure matches common biogeographic barriers or if it was eroded by human impacts. The well-documented invasion of E. multifasciata into Taiwan and neighboring islands provides a unique case-study to understand the signatures of human-mediated dispersal and assess whether expansion into other areas has been natural or human-mediated. We reanalyzed the RADseq data of Barley et al. [38] in this new context to better understand how ancestral mitochondrial structure can hold up to long periods of gene flow.

Materials and methods

Phylogenetic analyses

We used two datasets for phylogenetic analysis: 1) a mitochondrial ND2 dataset of 1023 bp with 361 samples of Eutropis multifasciata (343 were newly sequenced for this study; Supplemental Table S1) and 34 outgroups of Eutropis; and (2) a concatenated RADseq dataset comprised of 182,266 bp from 61 individuals (primarily from the Philippines). Only 18 individuals are common to both datasets, but geographically similar representatives are included from many of the clades. All newly generated sequences are deposited on GenBank (Supplemental Table S1).

For the ND2 dataset, trees were constructed using Bayesian Inference (BI) in MrBayes v3.2.1 [45], Maximum Likelihood (ML) in RAxML v8.1.15 [46], and divergence times estimated in BEAST v1.10.4 [47]. The appropriate partitioning scheme for each analysis was determined using PartitionFinder v2.1.1 [48] based on the Bayesian Information Criterion and greedy algorithm [49] and a starting tree using PhyML [50]. For each analysis, PartitionFinder specified a separate partition and substitution model for each codon position of ND2. For the ML analysis, a GTRCAT model was applied to each codon position separately and run with 1000 rapid bootstrap replicates. For MrBayes, the Markov Chain Monte Carlo (MCMC) was run for 100 million generations, with 4 chains, with the first codon position under GTR + I + G, the second under HKY + I + G, and the third under GTR + G. For all Bayesian analyses, we used Tracer v1.7.1 [51] to assess adequate convergence of the Markov chains and ESS values, and discarded the first 25% of trees as burn-in.

We reanalyzed the RADseq dataset of Barley et al. [38] in a phylogenetic framework, as it had only been previously analyzed specifically to investigate population and landscape genetics. We reduced the dataset to one individual per population for a total of 61 individuals to reduce the amount of missing data. Sequences were aligned using ipyrad v0.7.28, largely using default parameters except for clust_threshold = 0.94 and min_samples_locus = 46. The concatenated alignment consisted of 182,266 bp of sequence data containing 2928 parsimony informative sites. We estimated the phylogeny using four different programs: RAxML, MrBayes, BEAST, and SVDquartets [52] each producing a nearly identical topology. RAxML was run using the concatenated, unpartitioned alignment under the GTRCAT model with automatic rapid bootstrap termination. Two independent runs of MrBayes used the same alignment and were run under the GTR + G model with four chains and a temperature of 0.05 for 10 million generations sampling every 1000. The runs were combined and the first 25% of the trees were discarded as burnin. For SVDquartets, we randomly took one informative SNP per site for a total of 2392 SNPs. SVDquartets was run using PAUP v4.0a164 [53] with 100 bootstrap replicates to assess support.

Divergence times

For the BEAST mitochondrial analysis, we removed the outgroups and analyzed the resulting 361 taxa alignment using a coalescent constant population size tree prior. The HKY + G model was chosen for the first and second codon positions and GTR + G was chosen for the third. We calibrated an uncorrelated lognormal molecular clock with a normally distributed prior with a mean of 0.00895 and standard deviation of 0.0025 [54]. The MCMC was run for 100 million generations and the first 25% of trees were discarded as burn-in.

Estimates of divergence times for phylogenetic analysis of RADseq data are difficult to accurately determine without appropriate fossil or secondary calibrations [55]. Still, we wanted to generate approximate divergence time estimates for the nuclear data for comparison with the mitochondrial estimates. We used BEAST v1.10.4 to estimate divergence times using a fixed, strict molecular clock set at 0.001 substitutions per site per million years, which was empirically derived from calibrations in other skink lizards [56]. We used a coalescent, constant population size tree prior, and modeled evolution of the concatenated matrix under the HKY + I + G model. The MCMC was run for 100 million generations sampling every 10,000 with convergence assessed as above. This molecular clock calibration is unlikely to be an accurate measure of the genome-wide substation rates across RADseq loci, however we use it to give an approximation of divergence times and to compare with the mitochondrial calibration.

Species distribution model

We built a species distribution model (SDM) projected onto the last glacial maximum (LGM) to assess the extent that Pleistocene climatic fluctuations may have influenced relatively recent range expansions. This strategy assumes conservatism in the fundamental niche across climatic fluctuations, which may not have been the case especially in high latitude or high elevation regions and may underestimate the LGM distribution. We constructed the SDM with MaxEnt v.3.4.4 [57]. We combined our specimen localities with verified occurrence records from GBIF (accessed 13 Feb 2019; including museum records and iNaturalist research-grade observations). We reduced GBIF localities to those with relatively high precision and only took one record per unique latitude/longitude resulting in 1300 unique localities. To further reduce spatial auto-correlation, we reduced these localities to one point per 100 km using the machu.occ.rarefy command within the machuruku R package [58], resulting in 234 unique points. We downloaded 19 bioclimatic variables at 30 arcsecond resolution from the PaleoClim database [59] for the present and last glacial maximum (LGM) and cropped them to an extent from − 12° to 28° latitude and 88° to 141° longitude to include the entire range of E. multifasciata. We removed highly correlated variables by extracting pixel values at the occurrence points using the extract command in the Raster package [60] in R and testing for correlations. We reduced the set of variables to eight: Bio2 — Mean Diurnal Range, Bio3 — Isothermality, Bio4 — Temperature Seasonality, Bio5 — Max Temperature of the Warmest Month, Bio6 — Min Temperature of Coldest Month, Bio15 — Precipitation Seasonality, Bio16 — Precipitation of Wettest Quarter, Bio17 — Precipitation of Driest Quarter. We tested for the optimal set of feature classes and regularization multipliers using the ENMevaluate package in R, allowing for any combination of linear, quadratic, hinge, and product feature classes and regularization multipliers from 0.5 to 5 (by 0.5), 10, or 20. We used both linear and quadratic feature classes and selected a regularization multiplier of 3.5 based on the AUC. We ran the MaxEnt model using cross-validation for 10 replicates using a random 50% of the rarefied points for training and the remainder for testing. The median result of the replicates was summarized and plotted using QGIS [61] by selecting output raster values greater than 0.2.

Results

Phylogeography

We recovered four major mitochondrial clades within E. multifasciata in each analysis (Figs. 1 and 2) that are almost entirely geographically non-overlapping. (1) Western Clade — Borneo, Sumatra, Peninsular Malaysia, Indochina, Hainan, the central and western Philippines, Lanyu, and Ludao. (2) Eastern Clade — Java, Sulawesi, the Lesser Sunda Archipelago, the northern and southern Philippines, Taiwan, and Ludao. (3) BTS Clade — Bali, Timor, and Seram. (4) Enggano — Enggano Island. The mitochondrial and nuclear trees, though containing different sets of samples, showed nearly identical topologies with respect to the predominant subclades within the Western and Eastern Clades. The RADseq data only included representatives from the mitochondrial Western and Eastern Clades and not the BTS and Enggano Clades. Two samples showed mito-nuclear discordance between subclades separated by the Isthmus of Kra (Fig. 2; more details below).

Fig. 1
figure 1

A Map displaying phylogeographic structure within Eutropis multifasciata, with the inset showing the mitochondrial BEAST timetree. Triangular tips in the tree have been collapsed and indicate more than one sample within a clade. The colors of geographic sampling points correspond to the tip colors in the tree and are nested according to the larger colored shapes corresponding to the encircled colored clades in the tree. Subclades are labeled corresponding to the tree tip labels. Nodes in the timetree with posterior probability greater than 0.95 are shown with a black dot, and timescale is in units of million years. Relevant biogeographic features discussed in the text are labeled. The dotted black lines indicate the geographic extent of the major numbered clades. Wallace’s Line and Huxley’s modification are shown in light blue for reference. The underlying map is shaded according to a digital elevation model (SRTM). The sea level bathymetry displays the 40 m (darker grey) and 120 m (lighter grey) depth contours (GEBCO), and indicate the areas exposed during periods of reduced sea levels. Location of inset maps shown by boxes. B Inset map of Sulawesi. White-rimmed points within the green T3 clade display additional phylogenetic substructure not indicated on the collapsed timetree. C Inset map of Taiwan, Ludao, and Lanyu

Fig. 2
figure 2

Comparison between the simplified ND2 BEAST tree (left) and the concatenated RADseq BEAST tree (right). Clades recovered in both datasets are indicated by colored boxes. Solid lines between the boxes represent samples that are present in both trees, whereas dashed lines represent clades that are presumed to be equivalent based on geographic location and tree position. The two magenta lines refer to instances of mismatch between the two trees and possible introgression. Black dots at nodes represent posterior probabilities greater than 0.95, or are labeled otherwise, and bars at nodes represent the 95% confidence interval on node height

The Western and Eastern Clades displayed different patterns: Western Clade subclades did not overlap geographically while several Eastern Clade subclades overlapped on individual islands. The boundary between Western and Eastern Clades in both the mitochondrial and nuclear data roughly follows Huxley’s modification of Wallace’s Line (see Fig. 1); however, Java and Bali on the Sunda Shelf are recovered within the Eastern and BTS clades and the central Philippines are recovered within the Western Clade.

We observed discordance with respect to the root position across the mitochondrial phylogenetic analyses. Using RAxML and MrBayes, we consistently recovered a poorly-supported topology of the BTS Clade sister to all remaining E. multifasciata, and Enggano sister to the Eastern Clade, whereas the BEAST molecular-clock rooting recovered the Western Clade sister to the remaining three clades with strong support (Fig. 1). Given this uncertainty, we do not assume any particular root position in the following phylogeographic discussion and instead focus on patterns within the major divergent clades themselves. See Supplementary Text for further details.

Western clade

We recovered four primary subclades within the Western Clade that were geographically separated, and a fifth divergent outlier lineage. The boundaries between these subclades correspond to the Isthmus of Kra, between north and south Sumatra, and between north Borneo and the rest of Borneo. Interestingly, two samples from southern Myanmar (Tanintharyi) were recovered in discordant mitochondrial and nuclear clades across the Isthmus of Kra (see Fig. 2). One sample was recovered in the primarily west Malaysian mitochondrial clade (S1), but the nuclear Indochina clade, and the other sample shows the opposite pattern (C7). Within Indochina, we observed moderate genetic structure with more than eight well-supported slightly diverged mitochondrial lineages but low support for their relative relationships (C1–8). Peninsular Malaysia and offshore islands were recovered in a clade with central Sumatra and islands off the west coast of Sumatra, though with low support for their respective relationships (Fig. 1A, S1–6). Interestingly, samples from Hainan Island are recovered within this clade (S2), closest to southern Thailand and southern Myanmar (S1 and S2). In both RADseq and mtDNA data, northern Borneo (F3) was most closely related to the central Philippines (F1 and F2). A single sample from Bukit Kana in Borneo (I) was recovered as a divergent sister lineage to the entire remaining Western Clade. This single sample also possesses a divergent haplotype in four nuclear genes [62] (results not shown). See Supplementary Text for more details.

Eastern clade

We recovered three primary subclades within the Eastern Clade: (1) A widely-distributed clade in the Lesser Sundas, Luzon, the Eastern Peninsula of Sulawesi, Seram, and Halmahera (U1–5). This clade also includes samples from Taiwan and Ludao; (2) The Northern Peninsula of Sulawesi, Mindanao, and Tawi Tawi in the Sulu Archipelago of the Philippines (L1–2); (3) A clade spanning almost all of Sulawesi but also with representatives on Java, the Zamboanga peninsula of Mindanao, and Ambon (T1–3). Sulawesi was found to contain representatives of all three of the above subclades, and had substructure within each of them (Fig. 1B). The eastern Lesser Sundas allied with Seram, Halmahera and Sulawesi. Luzon (U3) was sister to the two primarily Lesser Sunda clades (U4 and U5). The predominant lineage on Taiwan was identical to haplotypes on Luzon (U3), and was also found on Ludao (Fig. 1C). See Supplementary Text for more details.

BTS and Enggano clades

The BTS Clade was comprised of three divergent lineages spanning across Wallace’s Line. A sample from Timor (M1) was recovered with strong support as sister to a sample from Seram (M2). Several samples from Bali (M3) were recovered as sister to Timor and Seram together. The BTS Clade was 5–9% divergent (raw-pairwise divergence) from all remaining E. multifasciata. The Enggano Clade was comprised solely of samples from remote Enggano island. This clade was also 5–9% divergent from all remaining E. multifasciata.

Divergence times

The mtDNA and RADseq divergence time estimates differed substantially (median crown age of mtDNA: 9.0 Ma; RADseq: 3.1 Ma; see Fig. 2). We emphasize that these molecular clock calibrations should be interpreted with significant caution. We chose to include the divergence time estimates only to show a broad range of possible ages of these clades, rather than to test specific dispersal timing hypotheses. It is difficult to accurately estimate divergence times using RADseq data without fossil or secondary calibrations [55] that are unavailable for Eutropis [41]. In the RADseq data, coalescence times between the Western vs. Eastern clades are only about 7500 generations old and migration rates (in units of 2Nem) range from 0.05–4 per generation [38]. This highlights that moderate amounts of gene flow are present across the range of E. multifasciata, and that a bifurcating tree is an inaccurate representation of the reticulating intraspecific genealogy. Our estimated nuclear divergence times therefore only represent a rough indication of overall genetic divergence at our assumed clock rate and are likely biased towards older divergence estimates. Despite these limitations, the overall topological concordance between the mitochondrial and nuclear clades indicates that the mitochondrial tree is still likely to be a good indicator of the original dispersal pathways.

Species distribution model

The SDM predicted areas occupied by E. multifasciata (Fig. 3). The ten replicates had an average area under curve of 0.766 and standard deviation is 0.034. Bio2 — Mean Diurnal Range had the highest percent contribution (41.9%), followed by Bio6 — Min Temperature of Coldest Month (37.9%). Bio2 also had the highest permutation importance (65.6%), followed by Bio4 — Temperature Seasonality (26.1%). The projected LGM SDM suggests the range of E. multifasciata contracted substantially, with much less suitable habitat especially in the northern part of its range. Hainan and all but the southern tip of Taiwan are predicted to have been uninhabitable during the LGM. Some interior regions of the exposed Sunda Shelf were unsuitable, though suitable corridors are predicted. All islands in the Philippines and Wallacea are predicted to have been suitable.

Fig. 3
figure 3

Simplified species distribution model for Eutropis multifasciata under the present climate (orange) and projected on the climate of the last glacial maximum (LGM; Purple). Much of its northern range is likely to have been uninhabitable during the LGM. Back points indicate all georeferenced localities of E. multifasciata, and blue points indicate the rarefied localities used to train and test the species distribution model to avoid spatial autocorrelation. Note that all of Hainan and most of Taiwan are predicted to have been uninhabitable during the LGM

Discussion

Phylogeography

Despite being frequently transported by humans, human-commensal species most often retain ancestral genetic structure across their native range whereas newly introduced populations show classical founder-effects or evidence of multiple introductions [5, 8, 9, 11, 63,64,65]. In this study we find evidence that E. multifasciata naturally colonized most of its enormous range across Southeast Asia where it retains substantial mitochondrial and nuclear genetic structuring, and that only the northern and eastern fringes were the result of recent human introduction or post-warming expansion.

Western clade — mitochondrial breaks correspond to expected biogeographic barriers

Within the Western Clade, several of the major genetic breaks correspond to commonly observed biogeographic barriers in other taxa, suggesting it is all or nearly all part of the native range. The major break within E. multifasciata in mainland southeast Asia and the Sunda Shelf occurs near the narrow Isthmus of Kra (Clades C and S). The Isthmus of Kra is a major biogeographic barrier separating the biogeographic regions of Sundaland and Indochina [66]. It is common for population or species turnover to occur north or south of the narrowest part of the isthmus [67]. The observed discordance across the isthmus between mitochondrial and nuclear data (Fig. 2) indicates that there is gene flow between the subclades, even at this substantial barrier for most taxa. This is not surprising given the dispersal capability of E. multifasciata and suggests this is a porous barrier. Gene flow can produce unusual artefactual phylogenetic patterns [68,69,70], so a careful assessment of these lineage boundaries beyond the mitochondrial patterns is warranted.

The genetic breaks we found on the Sunda Shelf also correspond to several frequently observed biogeographic barriers. Peninsular Malaysia and central Sumatra were recovered together (S4), matching a commonly observed pattern as they are only separated by a narrow channel and have been frequently connected by land during Pleistocene sea-level fluctuations [71,72,73]. The connection between southern Sumatra and northwestern Borneo (though poorly supported in the analyses) also has been previously reported and likely reflects their previous connection by a raised land-corridor [13, 71, 74]. The main split in northern Borneo (between clades A and F) is also a commonly observed pattern. In several bird species, populations in northwestern Borneo are more closely related to those in eastern Sumatra or Peninsular Malaysia than north Borneo [75,76,77,78,79]. The LGM projection of the species distribution model is consistent with the existence of dispersal corridors for E. multifasciata across the exposed Sunda Shelf (Fig. 3). The exposed Sunda Shelf is predicted to have been a more xeric, savannah-like environment during these climatic fluctuations [80,81,82,83], so it is possible that species such as E. multifasciata that are adapted to open and disturbed habitats were more suited to crossing it than rainforest-adapted species, such as the congener Eutropis rugifera, that faced barriers [84].

Across northern Borneo, where we have the finest-scale sampling, our results support the large Rejang and Baram rivers as barriers separating the mitochondrial lineages. The split on Sumatra is also concordant with the location of the large Batang Hari river, though sampling is not sufficient to confirm the specific location. It is counter-intuitive that freshwater rivers would act as such strong genetic barriers after considering how the species was able to cross much more substantial marine barriers. Trans-riverine dispersal should be more frequent and bidirectional while trans-marine should be rarer and generally unidirectional. Under the assumption that there exists a genetic priority effect preventing or reducing gene flow between clades [16], then these barriers likely reflect a geographic weakness or habitat break where a tension zone can settle on rather than a strong isolating force [85].

Further north, we observed a genetic break between east and west lineages across Indochina in the nuclear data that was poorly supported by the mtDNA. These lineages diverged recently (435 generations) and show strong unidirectional gene flow from east to west (3.99 effective migrants per generation) [38]. The Shan Plateau and Dawna ranges between Myanmar and Thailand may be difficult for E. multifasciata to cross, potentially reinforcing the break in this area. Furthermore, our LGM SDM (Fig. 3) suggests that E. multifasciata in Indochina was split along Indochina’s eastern and western coasts during global cooling events (though this may be an underestimate), and the post-warming expansion northwards could have brought these moderately divergent clades back into contact along a natural barrier. India, Bangladesh, and much of Myanmar were likely colonized during a Pleistocene post-warming range expansion as these populations show very low genetic diversity (also see [18]). In Thailand, Cambodia, Laos, and Vietnam, we do not observe mtDNA phylogeographic structuring associated with clear geographic barriers, but the clades are generally nonoverlapping indicating that some geographic structuring is being maintained in this region.

The mtDNA and nuclear data both split the Philippines into three separate clades: Mindanao, Luzon, and the Central Philippines (Fig. 2; also see [38]). These results suggest that Palawan was a stepping-stone pathway for dispersal into the central Philippines, as has been proposed for several taxa [39, 86]. Divergence times between Borneo and the central Philippines occurred after the Philippines was near its present geographic conformation [39]. There is a moderate amount of gene flow between the Central Philippines and both Mindanao and Luzon (1.7–2.1 effective migrants per generation) [38], though clearly not enough to erode the phylogeographic structure we observed. It is unclear why the Philippine clades are divided as such, but it could be through replacement of one clade by another and/or selective processes leading some populations to be better adapted to small versus large islands [17]. In any case, given the high levels of genetic diversity concordant with geography, the Philippines are very likely a part of the native range.

The Bukit kana outlier

It is unclear what process could lead a single sample from Bukit Kana (I) in Sarawak on Borneo to be recovered, on its own, sister to all remaining Western Clade samples. The first author sequenced four nuclear genes for this sample for his Master’s thesis, and found it held a divergent haplotype for two genes, but was very similar to Sarawak samples in the other two genes [62]. This positioning could indicate hybridization between representatives of the Eastern and Western clades [13, 68,69,70], though there was no evidence of increased heterozygosity in these markers, suggesting it is not a recent hybrid. What is puzzling about the placement in the mtDNA tree is that we should not observe an intermediate hybridization tree placement effect in mtDNA as it is uniparental. This raises the possibility that this haplotype represents a relict mitochondrial lineage that has held up to subsequent nuclear gene flow. It will be interesting to test these alternatives in future work with increased sampling in this region.

Java does not follow Wallace’s line

The only location where Clade 2 crosses to the west side of Wallace’s Line is Java, where most samples were nested in a clade comprised predominately of Sulawesi samples (T3), but we also observe another haplotype present (T2). Despite Java’s proximity to Sumatra and Borneo, species and populations on Java are often found to be divergent from the remaining Sunda Shelf and closely related to Wallacean counterparts [87]. However, since Java has the highest human population density of any island in the world, it is conceivable that on Java the ancestral genetic structure may have been overridden by human-introduction from the increased travel and shipping from neighboring areas. In an early study with just a handful of samples and different genetic markers, Java was recovered with Bali, Seram, and the Philippines [44], so the BTS clade could be an ancestral mitochondrial lineage on Java.

Eastern clade — complex phylogeography and signatures of multiple introductions

Patterns in the Eastern Clade contrast starkly with those in the Western Clade, with many of the major subclades overlapping geographically, which is likely the result of multiple human introductions. Multiple introductions will often lead to haplotypes from multiple sources in a single new population leaving behind a potentially misleading signature of high genetic diversity [5, 9, 88]. Sulawesi stands out in holding five divergent mitochondrial lineages. Mindanao holds three, Seram two, and Taiwan two. We consider at least some of these cases to be the result of human-mediated dispersal (and especially after considering Taiwan, discussed later), as these clades would be unlikely to come into sympatry otherwise.

Some of the structure on Sulawesi is likely the result of the biogeographic history of the island, though it appears to have been eroded to some extent. Even as recent as 2 Ma, Sulawesi was divided into seven or more paleoislands [89] on which lineages of many species have been shown to diverge substantially [90]. Our results are consistent with isolated lineages of E. multifasciata on these paleoislands coming into secondary contact as these paleo-islands fused to form the present-day Sulawesi. For example, lineages L1 and L2 may have diverged on the paleoislands making up the northern Peninsula, and subsequently come into contact (Fig. 1B). It is intriguing that priority effects that seem to be strong in other parts of the range do not appear to be keeping these lineages from becoming sympatric. Perhaps it reflects the divergence that accumulated between populations, which may not have been substantial enough to prevent nuclear gene flow and segregation of mitochondrial haplotypes. Nevertheless, our results are consistent with Sulawesi as a source of lineage diversity that possibly expanded from there into other areas, including the Philippines, Java, and eastern Wallacea.

BTS clade — a relict or human-mediated?

The BTS Clade has been previously recovered [19], but without broader geographic sampling to assess its evolutionary history. It is unusual to see an association between these islands and not the other intermediate Lesser Sunda islands [19, 91, 92], and there is no clear biogeographic explanation for this result. It does show some similarity to Cyrtodactylus which had multiple waves of dispersal through the Lesser Sundas [93]. One possible scenario is that the BTS Clade was more widely prevalent across the Lesser Sunda Archipelago but was later replaced and relictualized by the Eastern Clade. The substantial molecular divergences in this clade reject human-mediated dispersal. This clade is divergent but should not necessarily be taxonomically recognized. A Bali subspecies (E. m. balinensis) was previously recognized, but was synonymized after a morphological comparison found no significant differences [94].

Enggano clade — an ancient, isolated lineage

We recovered a divergent outlying lineage only including specimens from Enggano Island. Enggano is an isolated deep-water island 120 km off the coast of Sumatra that hosts several endemic lizards including two geckos and a flying lizard, and future molecular studies are likely to uncover new endemics [95]. Within the range of E. multifasciata, Enggano is one of the most isolated islands in terms of distance from other possible sources, yet for a species with such high dispersal capability it is still surprising to find such a divergent lineage here and not on several of the other isolated islands it occupies. Further investigation of morphological and nuclear genetic divergence will assess if this population warrants taxonomic recognition as a subspecies or species.

Recent natural and human-mediated expansion at the fringe

The well-documented invasion of E. multifasciata into Taiwan, Ludao, and Lanyu provides a unique opportunity to understand the signature of human-mediated dispersal and allows us to assess other candidate cases by looking for similar patterns. In 1992, E. multifasciata was first documented in Taiwan near the Kaoshiung international trading port [33], presumably from a cargo or timber shipment, and subsequently spread across much of the south and central parts of the island [32]. Today E. multifasciata has become very successful on Taiwan and only seems to be restricted from areas with cold winter temperatures [21]. Repeated Pleistocene cooling events expanded this unsuitable area to nearly all of Taiwan (Fig. 3), and this is likely why E. multifasciata did not establish a persistent natural population despite movement of early humans between Taiwan and the Philippines [96]. In 2008, E. multifasciata was discovered on Ludao near the international airport with a biased sex ratio indicating it recently colonized the island [37]. Since then it has remained isolated to an area about 10 ha around the airport and has been the subject of a significant eradication effort [36, 37]. The arrival on Lanyu is less clear, and it has been questioned whether it was natural or human-mediated [32, 97].

Our results suggest multiple human introductions to Taiwan and offshore islands, some of which have clear sources. Most of Taiwan shares a haplotype with Luzon (U3). Luzon is the closest natural population of E. multifasciata to Taiwan with a major port, and it could have easily been carried by a cargo or timber ship. This haplotype is also represented by one sample from Ludao, indicating a potential introduction from mainland Taiwan or separately from Luzon to the island. A second introduction is from a unique but closely related clade represented by two samples from southwest Taiwan (U1). The second source is less clear, as the samples are resolved sister a population on Sulawesi that could be natural or introduced (U2). The third unique haplotype is present on Lanyu and Ludao and is shared with Ayeyarwady, Myanmar (C8). All samples from Lanyu possess this Myanmar haplotype, whereas only one sample from Ludao possesses it (all others from Ludao possess the Luzon haplotype, T. Lin, unpublished data). The tiny island of Ludao is the only location in the entire range where the Western and Eastern Clades occur in sympatry, a clear signature of two human introductions. Hybridization of these divergent lineages could provide genetic variation for which it could evolve traits for better invasion success, allowing it to spread to new areas [14].

We assessed other possible cases of human introduction by looking either for divergent haplotypes in sympatry or a founder-effect, and observed that all the possible cases occur at the northern and eastern fringes of what we expect is the native range. There are only a few islands other than Taiwan and Ludao with multiple divergent clades: Mindanao, Sulawesi, and Seram. Mindanao is represented by three divergent haplotypes, one nested within the clade in northern Borneo (A2), the second nested within the clade on the northern peninsula of Sulawesi (L2), and the third (T1) sister to the clade including Java and most of Sulawesi (T2 and T3). The first (A2) is likely human-introduced as it is nested within a northern Borneo clade and collected near an airport outside the highly urbanized city of Cagayan de Oro. The second lineage is also likely human-introduced as it shares a haplotype with Sulawesi samples and was collected in Davao City, the largest city on Mindanao. The third lineage (T1) could correspond to a natural dispersal, as it is a divergent mitochondrial haplotype and is shown to be more widespread across Mindanao from the RADseq data (Fig. 2). This is consistent with natural dispersal between Mindanao and Sulawesi along the Sangihe-Talaud Archipelago [39, 98], though the direction is unclear. Mindanao is a large and complex island that harbors seven species of Eutropis [84, 99] so multiple lineages may be more able to persist on different parts of this island than on others.

At the southeast edge of the range, the Lesser Sunda lineages likely represent a recent natural expansion [19], but one of the lineages (U5) may have been introduced to Seram, central Sulawesi, and Halmahera as these all share nearly identical haplotypes—evidence of a founder-effect. On Seram, we also observe multiple introductions, but since E. multifasciata has been on Seram for at least 100 years [100] a more thorough investigation is warranted to confirm human introduction for both haplotypes. Seram and Halmahera are at the very edge of the range, and close to New Guinea where E. multifasciata recently established another new population [31].

The observed low genetic diversity in at the northern edge of the range could possibly be the result of either post-warming expansion or human-introduction. The LGM SDM is concordant with Taiwan being mostly uninhabitable during that time, but Hainan and much of Indochina are also predicted as uninhabitable (Fig. 3). This is likely due to colder winter temperatures as this currently limits its elevational and latitudinal range in Taiwan [21] and is a top predictive variable in the SDM. Many refugia likely persisted across Indochina that could explain the observed high genetic diversity; however, northeastern India and northern Myanmar show a possible founder-effect consistent with post-warming expansion [15]. Similarly, all samples from Hainan share an identical haplotype, but interestingly it matches southern Thailand south of the Isthmus of Kra (S2) and this haplotype is closely related to nearby southern Myanmar (S1 and S2). If the Hainan population dispersed naturally, then this relationship is strange as we would expect the dispersal to come from a nearby landbridged Indochina clade. Further, if E. multifasciata had been present on Hainan since before the LGM and was able to persist in refugia (e.g., along the changing coastline), why then is it not recovered as a divergent lineage but is instead identical to the sample in Myanmar? The earliest museum specimen of E. multifasciata from Hainan was collected ca. 1909 (MCZ:HERP:R-7384, exact collection year not specified) and several more specimens were collected in 1923 at multiple localities (MCZ:HERP:R-39307–8; FMNH-7252–61), so the timing of potential human introduction would have been quite old. Still, Hainan’s treaty port at Haikou had been active since 1858 [101], so E. multifasciata could easily have been introduced from a trading ship in the late-1800’s. This hypothesis should be tested with nuclear genetic data from additional localities on Hainan to rule-out more recent mitochondrial introgression or natural expansion and to assess population demography across the island.

Conclusions

Based on our results from Taiwan, Ludao, and Lanyu, and other instances of indisputable [29, 31, 34] and potential (Sulawesi, Mindanao, Seram, Halmahera) human-mediated dispersal, it is clear that E. multifasciata is very frequently transported around Southeast Asia. In the native range, as evidenced by genetic structure matching common biogeographic barriers, ancestral structure has held up to this frequent human-mediated dispersal. At the fringes, however, we observe divergent haplotypes in sympatry or founder-effects indicative of recent range expansion. Our results support E. multifasciata holding an enormous native range across much of Southeast Asia, and even as it became highly abundant in human-disturbed areas the observed ancient genetic structure remained intact. This work provides an approach that can be applied to other species to investigate their native vs. introduced ranges.

In the native range, we hypothesize that genetic structure is maintained through a combination of high-density blocking and/or local fitness advantages that could be the underlying basis of a genetic priority effect. If new populations become locally-adapted to their unique environment [9], which can evolve relatively rapidly from standing variation, they can hold back incoming gene flow from other areas [16, 102]. Local-adaptation may also buffer against climatic shifts, and could have led to underestimates of our LGM SDM and impacted our RADseq results. High-density blocking prevents new migrants from establishing and introgression from successfully occurring simply due to raw numbers of individuals and corresponding allele frequency differences [15, 103]. Combined, it may be very difficult for gene flow to erode existing phylogeographic structure. Contact between lineages may not initially coincide with biogeographic barriers, but may move and settle on even narrow habitat breaks such as those we observed across rivers, mountain ranges, and isthmuses [85]. Conversely, at the fringes of the native range, this pattern breaks down due to near-simultaneous introduction of skinks by humans to previously uninhabited islands before these processes are established.

Overall, the phylogeography of E. multifasciata is complex owing to its incredibly large natural range in one of the most biogeographically complex and dynamic regions of the world [40]. Within the native range, it showcases several classical biogeographic patterns (e.g., Isthmus of Kra) while also not conforming to others (e.g., Philippines). As such, deciphering the exact extent of its native range is difficult, but it is still clear that recent expansion has been primarily limited to the northern and eastern edges of its range. E. multifasciata has established small populations around the world through human introduction, but only in Taiwan and offshore islands at the edge of its native range have populations grown explosively. This contrasts with other invasive skinks, such as Lampropholis delicata and Carlia ailanpalai, which also display substantial phylogeographic structuring in their native ranges in Australia and New Guinea, but have become invasive across the Pacific and do not appear to undergo the same edge-restricted expansion [65, 88, 104]. Future work following-up on new populations with known or approximate dates of establishment will be an interesting comparison with our results.

Availability of data and materials

All sequences are available on Genbank (Table S1) and full phylogenetic trees for all analyses can be found in the Supplemental Material.

References

  1. Carlton JT. Biological invasions and cryptogenic species. Ecology. 1996;77:1653–5.

    Article  Google Scholar 

  2. Saltonstall K. Cryptic invasion by a non-native genotype of the common reed, Phragmites australis, into North America. Proc Natl Acad Sci USA. 2002;99:2445–9.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  3. Yuan ML, Frederick JH, McGuire JA, Bell RC, Smith SR, Fenton C, et al. Endemism, invasion, and overseas dispersal: the phylogeographic history of the lesser Antillean frog, Eleutherodactylus johnstonei. Biol Invasions. 2022;24:2707–22.

    Article  Google Scholar 

  4. Clavero M, García-Berthou E. Invasive species are a leading cause of animal extinctions. Trends Ecol Evol. 2005;20:110.

    Article  PubMed  Google Scholar 

  5. Kolbe JJ, Glor RE, Schettino LR, Lara AC, Larson A, Losos JB. Multiple sources, admixture, and genetic variation in introduced Anolis lizard populations. Conserv Biol. 2007;21:1612–25.

    Article  PubMed  Google Scholar 

  6. Sakai AK, Allendorf FW, Holt JS, Lodge DM, Molofsky J, With KA, et al. The Population Biology of Invasive Species https://doi.org/10.1146/annurev.ecolsys32081501114037. 2003;32:305–32.

    Google Scholar 

  7. Cristescu ME. Genetic reconstructions of invasion history. Mol Ecol. 2015;24:2212–25.

    Article  PubMed  Google Scholar 

  8. Velo-Antón G, Burrowes PA, Joglar RL, Martínez-Solano I, Beard KH, Parra-Olea G. Phylogenetic study of Eleutherodactylus coqui (Anura: Leptodactylidae) reveals deep genetic fragmentation in Puerto Rico and pinpoints origins of Hawaiian populations. Mol Phylogenet Evol. 2007;45:716–28.

    Article  PubMed  Google Scholar 

  9. Reynolds RG, Kolbe JJ, Glor RE, López-Darias M, Gómez Pourroy CV, Harrison AS, et al. Phylogeographic and phenotypic outcomes of brown anole colonization across the Caribbean provide insight into the beginning stages of an adaptive radiation. J Evol Biol. 2020;33:468–94.

    Article  PubMed  Google Scholar 

  10. Tsutsui ND, Case TJ. Population genetics and colony structure of the argentine ant (Linepithema humile) in its native and introduced ranges. Evolution (N Y). 2001;55:976–85.

    CAS  Google Scholar 

  11. Puckett EE, Munshi-South J. Brown rat demography reveals pre-commensal structure in eastern Asia prior to expansion into Southeast Asia. Genome Res. 2019;29:gr.235754.118.

    Article  Google Scholar 

  12. Bock DG, Baeckens S, Pita-Aquino JN, Chejanovski ZA, Michaelides SN, Muralidhar P, et al. Changes in selection pressure can facilitate hybridization during biological invasion in a Cuban lizard. Proc Natl Acad Sci USA. 2021;118

  13. Chan KO, Hutter CR, Wood PL, Grismer LL, Das I, Brown RM. Gene flow creates a mirage of cryptic species in a southeast Asian spotted stream frog complex. Mol Ecol. 2020;29:3970–87.

    Article  CAS  PubMed  Google Scholar 

  14. Blumenfeld AJ, Eyer PA, Husseneder C, Mo J, LNL J, Wang C, et al. Bridgehead effect and multiple introductions shape the global invasion history of a termite. Commun Biol. 2021;4:1–12.

    Article  Google Scholar 

  15. Waters JM, Fraser CI, Hewitt GM. Founder takes all: density-dependent processes structure biodiversity. Trends Ecol Evol. 2013;28:78–85.

    Article  PubMed  Google Scholar 

  16. De Meester L, Vanoverbeke J, Kilsdonk LJ, Urban MC. Evolving perspectives on monopolization and priority effects. Trends Ecol Evol. 2016;31:136–46.

    Article  PubMed  Google Scholar 

  17. Farkas TE, Hendry AP, Nosil P, Beckerman AP. How maladaptation can structure biodiversity: eco-evolutionary island biogeography. Trends Ecol Evol. 2015;30:154–60.

    Article  PubMed  Google Scholar 

  18. Hossain M, Jing C. Molecular phylogeny of common sun skink, Eutropis (Squamata: Scincidae), of northwestern part of Bangladesh. Bangladesh J Zool. 2019;47:13–25.

    Article  Google Scholar 

  19. Reilly SB, Stubbs AL, Karin BR, Arida E, Iskandar DT, McGuire JA. Recent colonization and expansion through the lesser Sundas by seven amphibian and reptile species. Zool Scr. 2019:48.

  20. Das I. A field guide to the reptiles of South-East Asia. London: Bloomsbury Publishing Plc; 2010.

    Google Scholar 

  21. Lin TE, Chen TY, Wei HL, Richard R, Huang SP. Low cold tolerance of the invasive lizard Eutropis multifasciata constrains its potential elevation distribution in Taiwan. J Therm Biol 2019;82 September 2018:115–122.

  22. Chen T-Y, Richard R, Lin T-E, Huang S-P. Landscape forest impacts the potential activity time of an invasive lizard and its possibilities for range expansion in Taiwan under climate warming. J Therm Biol. 2021; 98 April:102948

  23. Ji X, Lin LH, Lin CX, Qiu QB, Du Y. Sexual dimorphism and female reproduction in the many-lined sun skink (Mabuya multifasciata) from China. J Herpetol. 2006;40:351–7.

    Article  Google Scholar 

  24. Ji X, Lin CX, Lin LH, Qiu QB, Du Y. Evolution of viviparity in warm-climate lizards: an experimental test of the maternal manipulation hypothesis. J Evol Biol. 2007;20:1037–45.

    Article  CAS  PubMed  Google Scholar 

  25. Li H, Mao L, Shi L, Ji X. Many-lined sun skinks (Mabuya multifasciata) shift their thermal preferences downwards when fasted. Asian Herpetol Res. 2010;1:36–9.

    Article  Google Scholar 

  26. Lin CX, Zhang L, Ji X. Influence of pregnancy on locomotor and feeding performances of the skink, Mabuya multifasciata: why do females shift thermal preferences when pregnant? Zoology. 2008;111:188–95.

    Article  PubMed  Google Scholar 

  27. Metallinou M, Weinell JL, Karin BR, Conradie W, Wagner P, Schmitz A, et al. A single origin of extreme matrotrophy in African mabuyine skinks. Biol Lett. 2016;12

  28. Meshaka WE, Butterfield BP, Hauge JB. The exotic amphibians and reptiles of Florida. Malabar, Fla: Krieger Pub. Co.; 2004.

    Google Scholar 

  29. Rochford MR, Ketterlin JK, Vinci J, Alexander DT, Mazzotti FJ. The many-lined sun skink, Eutropis multifasciata (Kuhl 1820) (Mabuyidae), a new population in Broward County, Florida. IRCF Reptil Amphib. 2018;25:217–9.

    Article  Google Scholar 

  30. Ingram G. Does the skink, Mabuya multifasciata, occur in Australia? North Territ Nat. 1987;10:11–2.

    Google Scholar 

  31. O’Shea M, Kusuma KI, Kaiser H. First record of the island wolfsnake, Lycodon capucinus (H. Boie in F. Boie 1827), From New Guinea, with comments on its widespread distribution and confused taxonomy, and a new record for the common sun skink, Eutropis multifasciata (Kuhl 1). ICRF. Reptil Amphib Conserv Nat Hist. 2018;25:70–84.

    Google Scholar 

  32. Lee KH, Chen TH, Shang G, Clulow S, Yang YJ, Lin SM. A check list and population trends of invasive amphibians and reptiles in Taiwan. Zookeys. 2019;2019:85–130.

    Article  Google Scholar 

  33. Ota H, Chang H, Liu K, Hikida T. A new record of the viviparous skink, Mabuya multifasciata (Kuhl, 1820) (Squamata: Reptilia), from Taiwan. Zool Stud. 1994;33:86–9.

    Google Scholar 

  34. Wiles GJ. Recent record of reptiles and amphibians accidentally transported to Guam, Mariana Islands. Micronesica. 2000;32:285–7.

    Google Scholar 

  35. Krysko KL, Somma LA, Gillette CR, Cueva D, Wasilewski JA, Enge KM, et al. New verified nonindigenous amphibians and reptiles in Florida through 2015, with a summary of over 152 years of introductions. IRCF Reptil Amphib. 2016;23:110–43.

    Article  Google Scholar 

  36. Chao RF. Using transformative learning theory to explore the mechanisms of citizen participation for environmental education on the removal of invasive species: the case of Green Island, Taiwan Eurasia. J Math Sci Technol Educ. 2017;13:2665–82.

    Google Scholar 

  37. Chao RF, Lin TE. Effect of citizen action on suppression of invasive alien lizard population: a case of the removal of Eutropis multifasciata on Green Island, Taiwan. Appl Ecol Environ Res. 2017;15:1–13.

    Article  CAS  Google Scholar 

  38. Barley AJ, Monnahan PJ, Thomson RC, Grismer LL, Brown RM. Sun skink landscape genomics: assessing the roles of micro-evolutionary processes in shaping genetic and phenotypic diversity across a heterogeneous and fragmented landscape. Mol Ecol. 2015;24:1696–712.

    Article  PubMed  Google Scholar 

  39. Brown RM, Siler CD, Oliveros CH, Esselstyn JA, Diesmos AC, Hosner PA, et al. Evolutionary processes of diversification in a model island archipelago. Annu Rev Ecol Evol Syst. 2013;44:411–35.

    Article  Google Scholar 

  40. Lohman DJ, De BM, Page T, Von RK, Hall R, Ng PKL, et al. Biogeography of the indo-Australian archipelago. Annu Rev Ecol Evol Syst. 2011;42:205–26.

    Article  Google Scholar 

  41. Karin BR, Metallinou M, Weinell JL, Jackman TR, Bauer AM. Resolving the higher-order phylogenetic relationships of the circumtropical Mabuya group (Squamata: Scincidae): an out-of-Asia diversification. Mol Phylogenet Evol. 2016;102:220–32.

    Article  PubMed  Google Scholar 

  42. Carranza S, Arnold EN. Investigating the origin of thansoceanic distributions: mtDNA shows Mabuya lizards (Reptilia, Scincidae) crossed the Atlantic twice. Syst Biodivers. 2003;1:275–82.

    Article  Google Scholar 

  43. Pinto-Sánchez NR, Calderón-Espinosa ML, Miralles A, Crawford AJ, Ramírez-Pinilla MP. Molecular phylogenetics and biogeography of the Neotropical skink genus Mabuya Fitzinger (Squamata: Scincidae) with emphasis on Colombian populations. Mol Phylogenet Evol. 2015;93:188–211.

    Article  PubMed  Google Scholar 

  44. Mausfeld P, Schmitz A. Molecular phylogeography, intraspecific variation and speciation of the Asian scincid lizard genus Eutropis Fitzinger, 1843 (Squamata: Reptilia: Scincidae): taxonomic and biogeographic implications. Org Divers Evol. 2003;3:161–71.

    Article  Google Scholar 

  45. Ronquist F, Huelsenbeck JP. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics. 2003;19:1572–4.

    Article  CAS  PubMed  Google Scholar 

  46. Stamatakis A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics. 2014;30:1312–3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Suchard MA, Lemey P, Baele G, Ayres DL, Drummond AJ, Rambaut A. Bayesian phylogenetic and phylodynamic data integration using BEAST 1.10. Virus Evol. 2018;4

  48. Lanfear R, Frandsen PB, Wright AM, Senfeld T, Calcott B. Partitionfinder 2: new methods for selecting partitioned models of evolution for molecular and morphological phylogenetic analyses. Mol Biol Evol. 2017;34:772–3.

    CAS  PubMed  Google Scholar 

  49. Lanfear R, Calcott B, Ho SYW, Guindon S. PartitionFinder: combined selection of partitioning schemes and substitution models for phylogenetic analyses. Mol Biol Evol. 2012;29:1695–701.

    Article  CAS  PubMed  Google Scholar 

  50. Guindon S, Dufayard J-FF, Lefort V, Anisimova M, Hordijk W, Gascuel O. New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0. Syst Biol. 2010;59:307–21.

    Article  CAS  PubMed  Google Scholar 

  51. Rambaut A, Drummond AJ. Tracer v1.6. Available from: http://tree.bio.ed.ac.uk/software/tracer/. 2013.

  52. Chifman J, Kubatko L. Quartet inference from SNP data under the coalescent model. Bioinformatics. 2014;30:3317–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Swofford DL. PAUP*. Phylogenetic analysis using parsimony (*and other methods) Version 4; 2002.

    Google Scholar 

  54. Barley AJ, Datta-Roy A, Karanth KP, Brown RM. Sun skink diversification across the Indian-southeast Asian biogeographical interface. J Biogeogr. 2015;42:292–304.

    Article  Google Scholar 

  55. Stange M, Sánchez-Villagra MR, Salzburger W, Matschiner M. Bayesian divergence-time estimation with genome-wide single-nucleotide polymorphism data of sea catfishes (Ariidae) supports Miocene closure of the Panamanian isthmus. Syst Biol. 2018;67:681–99.

    Article  PubMed  PubMed Central  Google Scholar 

  56. Brandley MC, Wang Y, Guo X, Montes De Oca AN, Fería-Ortíz M, Hikida T, et al. Accommodating heterogenous rates of evolution in molecular divergence dating methods: an example using intercontinental dispersal of plestiodon (Eumeces) lizards. Syst Biol. 2011;60:3–15.

    Article  CAS  PubMed  Google Scholar 

  57. Phillips SJ, Anderson RP, Schapire RE. Maximum entropy modeling of species geographic distributions. Ecol Model. 2006;190:231–59.

    Article  Google Scholar 

  58. Guillory WX, Brown JL. A new method for integrating ecological niche modeling with phylogenetics to estimate ancestral distributions. Syst Biol. 2021;70:1033–45.

    Article  PubMed  Google Scholar 

  59. Brown JL, Hill DJ, Dolan AM, Carnaval AC, Haywood AM. PaleoClim, high spatial resolution paleoclimate surfaces for global land areas. Sci Data. 2018;5:1–9.

    Article  ADS  Google Scholar 

  60. Hijmans RJ. The “raster” package (Version 1.2–6). 2021;:49.

  61. QGIS development team. QGIS geographic information system. Open source Geospatial Foundation project. 2021.

    Google Scholar 

  62. Karin BR. Phylogeography and evolution of three species of Eutropis from Borneo. Masters Thesis. Department of Biology, Villanova University; 2016.

    Google Scholar 

  63. Agarwal I, Ceríaco LMP, Metallinou M, Jackman TR, Bauer AM. How the African house gecko (Hemidactylus mabouia) conquered the world. R Soc Open Sci. 2021:8.

  64. Wogan GOU, Stuart BL, Iskandar DT, McGuire JA. Deep genetic structure and ecological divergence in a widespread human commensal toad. Biol Lett. 2016;12:20150807.

    Article  PubMed  PubMed Central  Google Scholar 

  65. Chapple DG, Hoskin CJ, Chapple SNJ, Thompson MB. Phylogeographic divergence in the widespread delicate skink (Lampropholis delicata) corresponds to dry habitat barriers in eastern Australia. BMC Evol Biol. 2011;11:1–18.

    Article  Google Scholar 

  66. de Bruyn M, Stelbrink B, Morley RJ, Hall R, Carvalho GR, Cannon CH, et al. Borneo and Indochina are major evolutionary hotspots for southeast Asian biodiversity. Syst Biol. 2014;63:879–901.

    Article  PubMed  Google Scholar 

  67. Parnell J. The biogeography of the isthmus of Kra region: a review. Nord J Bot. 2013;31:001–15.

    Article  Google Scholar 

  68. Pyron RA, O’Connell KA, Lemmon EM, Lemmon AR, Beamer DA. Candidate-species delimitation in Desmognathus salamanders reveals gene flow across lineage boundaries, confounding phylogenetic estimation and clarifying hybrid zones. Ecol Evol. 2022;12:e8574.

    Article  PubMed  PubMed Central  Google Scholar 

  69. Dolinay M, Nečas T, Zimkus BM, Schmitz A, Fokam EB, Lemmon EM, et al. Gene flow in phylogenomics: sequence capture resolves species limits and biogeography of Afromontane forest endemic frogs from the Cameroon highlands. Mol Phylogenet Evol. 2021;163

  70. Chan KO, Hutter CR, Wood PL, Su YC, Brown RM. Gene flow increases phylogenetic structure and inflates cryptic species estimations: a case study on widespread Philippine puddle frogs (Occidozyga laevis). Syst Biol. 2022;71:40–57.

    Article  CAS  Google Scholar 

  71. Voris HK. Maps of Pleistocene Sea levels in Southeast Asia: shorelines, river systems and time durations. J Biogeogr. 2000;27:1153–67.

    Article  Google Scholar 

  72. Sarr AC, Husson L, Sepulchre P, Pastier AM, Pedoja K, Elliot M, et al. Subsiding Sundaland. Geology. 2019;47:119–22.

    Article  ADS  Google Scholar 

  73. Husson L, Boucher FC, Sarr AC, Sepulchre P, Cahyarini SY. Evidence of Sundaland’s subsidence requires revisiting its biogeography. J Biogeogr. 2020;47:843–53.

    Article  Google Scholar 

  74. Woodruff DS, Turner LM. The Indochinese-Sundaic zoogeographic transition: a description and analysis of terrestrial mammal species distributions. J Biogeogr. 2009;36:803–21.

    Article  Google Scholar 

  75. Moyle RG, Schilthuizen M, Rahman MA, Sheldon FH. Molecular phylogenetic analysis of the white-crowned forktail Enicurus leschenaulti in Borneo. J Avian Biol. 2005;36:96–101.

    Article  Google Scholar 

  76. Sheldon FH, Lohman DJ, Lim HC, Zou F, Goodman SM, Prawiradilaga DM, et al. Phylogeography of the magpie-robin species complex (Aves: Turdidae: Copsychus) reveals a Philippine species, an interesting isolating barrier and unusual dispersal patterns in the Indian Ocean and Southeast Asia. J Biogeogr. 2009;36:1070–83.

    Article  Google Scholar 

  77. Lim HC, Sheldon FH. Multilocus analysis of the evolutionary dynamics of rainforest bird populations in Southeast Asia. Mol Ecol. 2011;20:3414–38.

    Article  PubMed  Google Scholar 

  78. Moyle RG, Taylor SS, Oliveros CH, Lim HC, Haines CL, Rahman M, et al. Diversification of an endemeic southeast Asian genus: phylogenetic relationships of the spiderhunters (Nectariniidae: Arachnothera). Auk. 2011;128:777–88.

    Article  Google Scholar 

  79. den Tex RJ, Leonard JA. A molecular phylogeny of Asian barbets: speciation and extinction in the tropics. Mol Phylogenet Evol. 2013;68:1–13.

    Article  Google Scholar 

  80. Cannon CH, Morley RJ, Bush ABG. The current refugial rainforests of Sundaland are unrepresentative of their biogeographic past and highly vulnerable to disturbance. Proc Natl Acad Sci USA. 2009;106:11188–93.

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  81. Gathorne-Hardy FJ, Syaukani DRG, Eggelton P, Jones DT. Quaternary rainforest refugia in Southeast Asia: using termites (Isoptera) as indicators. Biol J Linn Soc. 2002;75:453–66.

    Article  Google Scholar 

  82. Bird MI, Taylor D, Hunt C. Palaeoenvironments of insular Southeast Asia during the last glacial period: a savanna corridor in Sundaland? Quat Sci Rev. 2005;24:2228–42.

    Article  ADS  Google Scholar 

  83. Louys J, Meijaard E. Palaeoecology of southeast Asian megafauna-bearing sites from the Pleistocene and a review of environmental changes in the region. J Biogeogr. 2010;37:1432–49.

    Article  Google Scholar 

  84. Karin BR, Das I, Jackman TR, Bauer AM. Ancient divergence time estimates in Eutropis rugifera support the existence of Pleistocene barriers on the exposed Sunda shelf. PeerJ. 2017;2017:e3762.

    Article  Google Scholar 

  85. Barton NH, Hewitt GM. Adaptation, speciation and hybrid zones. Nature. 1989;341:497–503.

    Article  ADS  CAS  PubMed  Google Scholar 

  86. Brown RM, Siler CD. Spotted stream frog diversification at the Australasian faunal zone interface, mainland versus island comparisons, and a test of the Philippine “dual-umbilicus” hypothesis. J Biogeogr. 2014;41:182–95.

    Article  Google Scholar 

  87. Stelbrink B, Albrecht C, Hall R, Von RT. The biogeography of Sulawesi revisited: is there evidence for a vicariant origin of taxa on Wallace’s “anomalous island”? Evolution (N Y). 2012;66:2252–71.

    Google Scholar 

  88. Chapple DG, Miller KA, Kraus F, Thompson MB. Divergent introduction histories among invasive populations of the delicate skink (Lampropholis delicata): has the importance of genetic admixture in the success of biological invasions been overemphasized? Divers Distrib. 2013;19:134–46.

    Article  Google Scholar 

  89. Nugraha AMS, Hall R. Late Cenozoic palaeogeography of Sulawesi, Indonesia. Palaeogeogr Palaeoclimatol Palaeoecol. 2018;490:191–209.

    Article  Google Scholar 

  90. Evans BJ, Supriatna J, Andayani N, Setiadi MI, Cannatella DC, Melnick DJ. Monkeys and toads define areas of endemism on Sulawesi. Evolution (N Y). 2003;57:1436–43.

    Google Scholar 

  91. Reilly SB, Stubbs AL, Karin BR, Bi K, Arida E, Iskandar DT, et al. Leap-frog dispersal and mitochondrial introgression: Phylogenomics and biogeography of Limnonectes fanged frogs in the lesser Sundas archipelago of Wallacea. J Biogeogr. 2019;46:757–69.

    Article  Google Scholar 

  92. Reilly SB, Stubbs AL, Arida E, Karin BR, Arifin U, Kaiser H, et al. Phylogenomic analysis reveals dispersal-driven speciation and divergence with gene flow in lesser Sunda flying lizards (genus Draco). Syst Biol. 2022;71:221–41.

    Article  CAS  Google Scholar 

  93. Reilly SB, Stubbs AL, Karin BR, Arida E, Arifin U, Hamidy A, et al. Bewildering biogeography: waves of dispersal and diversification across southern Wallacea by bent-toed geckos (genus: Cyrtodactylus). Mol Phylogenet Evol. 2023;186

  94. Amarasinghe AAT, Thammachoti P, Campbell PD, Hallermann J, Henkanaththegedara SM, Suranjan Karunarathna DM, et al. Systematic composition of the Eutropis multifasciata (Kuhl 1820) species complex (Squamata: Scincidae) and designation of a neotype. Herpetologica. 2018;74:342–54.

    Article  Google Scholar 

  95. Grismer LL, Riyanto A, Iskandar DT, Mcguire JA. A new species of Hemiphyllodactylus Bleeker, 1860 (Squamata: Gekkonidae) from Pulau Enggano, southwestern Sumatra. Indonesia Zootaxa. 2014;3821:485–95.

    Article  PubMed  Google Scholar 

  96. Bellwood P, Fox JJ, Tryon D, editors. The Austronesians: historical and comparative perspectives. Canberra: Australian National University E Press; 2006.

    Google Scholar 

  97. Hsu MH, Lin JW, Liao CP, Hsu JY, Huang WS. Trans-marine dispersal inferred from the saltwater tolerance of lizards from Taiwan. Res Sq. 2021; https://doi.org/10.1371/journal.pone.0247009.

  98. Koch A, Arida E, Riyanto A, Hme W, Böhme W. Islands between the realms: a revised checklist of the herpetofauna of the Talaud archipelago, Indonesia, with a discussion about its biogeographic affinities. Bonner Zool Beiträge Herausgeber Zool Forschungsinstitut und Museum Alexander Koenig, Bonn. 2005;56:107–29.

    Google Scholar 

  99. Barley AJ, Diesmos AC, Siler CD, Martinez CM, Brown RM. Taxonomic revision of Philippine sun skinks (Reptilia: Squamata: Scincidae: Eutropis), and descriptions of eight new species. Herpetol Monogr. 2020;34:39–70.

    Article  Google Scholar 

  100. de Rooij N. The reptiles of the indo-Australian archipelago. Volume I. Lacertilia, Chelonia: Emydosauria. E. J. Brill, Leiden; 1915.

    Google Scholar 

  101. Nield R. China’s southernmost treaty port. Source J R Asiat Soc Hong Kong Branch. 2012;52:63–76.

    Google Scholar 

  102. Wang IJ, Bradburd GS. Isolation by environment. Mol Ecol. 2014;23:5649–62.

    Article  PubMed  Google Scholar 

  103. Fraser CI, Banks SC, Waters JM. Priority effects can lead to underestimation of dispersal and invasion potential. Biol Invasions. 2015;17:1–8.

    Article  Google Scholar 

  104. Austin CC, Rittmeyer EN, Oliver LA, Andermann JO, Zug GR, Rodda GH, et al. The bioinvasion of Guam: inferring geographic origin, pace, pattern and process of an invasive lizard (Carlia) in the Pacific using multi-locus genomic data. Biol Invasions. 2011;13:1951–67.

    Article  Google Scholar 

Download references

Acknowledgements

We thank Alan Resetar (FMNH), Carol Spencer (MVZ), Jens Vindum (CAS), Lauren Scheinberg (CAS), Lee Grismer (LSUHC), and Wolfgang Böhme (ZFMK) for their assistance in tissue loans and depositing specimens. We thank all those who helped with field work and who collected tissue samples, including Adi Shabrani, Pui Yong Min, Umilaela Arifin, Alexander Stubbs, and Ian Brennan. We thank members of the Bauer and Jackman labs for their emotional and social support through much of the lab work for this project. We also thank the McGuire lab and others in the Museum of Vertebrate Zoology for thoughtful discussions that helped substantially in writing this manuscript.

Funding

Financial support was provided by a National Science Foundation grant EF 1241885 (subaward 13–0632) to AMB and the Gerald M. Lemole, M.D. Endowed Chair Fund at Villanova University. Additional financial support was provided by Villanova University through a Biology Department Fellowship and a Graduate Student Fellowship to BRK. ID was supported by a Niche Research Grant by the Ministry of Higher Education, Government of Malaysia (IA010200–0708-0007). JAM acknowledges funding provided by the National Science Foundation (NSF; DEB 1258185 and 1457845) as well the National Geographic Society. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

B.R.K, M.L.S., T.R.J., J.A.M., and A.M.B. conceived the ideas; B.R.K., M.L.S., T.E.L., S.B.R., A.J.B., D.T.I., E.A., and J.A.M. collected specimens and data; B.R.K. and A.J.B. analyzed the data; and B.R.K. led the writing. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Benjamin R. Karin.

Ethics declarations

Ethics approval and consent to participate

Specimens from Sarawak, Malaysia were collected under Sarawak Forestry Department permit NCCD.907.4.4(11)-67 and from Indonesia under several permits including RISTEKDIKTI permit 0196/SIP/FRP/VI/2011.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1:

 Supplementary Text.

Additional file 2:

 Figures S1-S8.

Additional file 3:

 Table S1. Specimens used in phylogenetic analyses and corresponding GenBank accession numbers. Specimens at the top of the list are outgroups and followed by samples already deposited on GenBank. Published genetic sequences accessed from alternative repositories have the link available.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Karin, B.R., Lough-Stevens, M., Lin, TE. et al. The natural and human-mediated expansion of a human-commensal lizard into the fringes of Southeast Asia. BMC Ecol Evo 24, 25 (2024). https://doi.org/10.1186/s12862-024-02212-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12862-024-02212-7

Keywords