Summary
Using a multidisciplinary approach, we investigated whether human-controlled fire has historically influenced temporal niche partitioning between dark-diurnal and pale-nocturnal lineages of the Moorish gecko (Tarentola mauritanica). The pale-nocturnal variant exhibited lower skin melanin levels, smaller and fewer melanosomes, and lower plasma α-Melanocyte Stimulating Hormone levels than its dark-diurnal counterpart. Mitochondrial genome analyses indicated that the common ancestor of these gecko lineages diverged approximately 6,600 years ago, coinciding with the transition of modern humans from nomadic hunter-gatherers to settled agricultural societies. Species distribution models suggested coexistence between humans and geckos during the emergence of these lineages. Additionally, we demonstrated that fire attracts phototactic arthropods, concentrating prey resources. These findings imply that human-controlled fire may have created a novel foraging niche for pale-nocturnal geckos, likely driving the divergence of the two variants.
Subject areas: Zoology, Phylogenetics, Evolutionary biology
Graphical abstract
Highlights
-
•
Tarentola mauritanica shows two variants: a dark-diurnal and a pale-nocturnal form
-
•
According to mtDNA, these forms began diverging ca. 6,600 years ago
-
•
Human-controlled fire may have created a niche droving the split between these gecko forms
-
•
This phenomenon may represent an incipient speciation driven by human civilization
Zoology; Phylogenetics; Evolutionary biology
Introduction
Humans are the only animal species capable of changing the environment globally.1 Human action has largely affected the fate of wildlife throughout humankind’s history, promoting profound changes in community composition and geographic ranges.2 Humans have also caused numerous extinction events,1,3 filtering out many species from human-altered landscapes where more opportunistic species persist or thrive,4,5 and not infrequently exerting selective pressures that have led to new phyletic lineages and speciation.6
Artificial lighting at night (hereafter, ALAN) is one of the most pervasive forms of environmental alteration caused by humans. Numerous organisms adjust their behavior and life cycles according to the availability of light across various time frames, ranging from daily to seasonal cycles. By introducing changes to the natural cycle of light and darkness, ALAN significantly impacts several crucial aspects of animal existence,7,8 especially affecting nocturnal species, whose sensitivity is evident as they either actively avoid light, as seen in most bat species,9,10 or are attracted to it, as in numerous nocturnal insects showing positive phototaxis (e.g.,11).
While there is much knowledge about the global, overwhelming impact of ALAN (e.g.,12) the effects of the first form of human-generated lighting, fire, are practically unknown. The intentional use of fire traces back to the Middle Pleistocene of the Levant13,14 but spread as a common practice soon afterward.15 The earliest use of fire may have been for cooking and repelling predators, but it also facilitated social interactions. The use of fire likely became widespread during the transition of modern humans to settled agricultural communities, associated with increased human density and a new lifestyle. This may have led to a “daylight extension,” in which humans have developed an activity peak during late evening hours, an unusual condition compared to the mammalian standard, including other primates.16,17
Besides changing the microclimate, fire generates light at night, when numerous phototactic insects may get lured and fall victim to opportunistic predators exploiting the new environmental conditions and their associated food bonanza. However, no study has explored the possibility that fire has historically created a previously untapped foraging niche suitable for insectivorous vertebrates. Although there is no way of retrospectively investigating the soundness of this hypothesis, several clues may be gathered at least to test the potential influence of ancient human-governed fires on animal ecology and evolution.
We selected the Moorish gecko (Tarentola mauritanica) as a case study to reach this goal. This small saurian is widespread in North Africa and Europe and has the widest range of all gecko species in the Mediterranean Basin.18 Because of the close relationship with humans, the species has been accidentally introduced to the Balearic Islands (Spain), Madeira (Portugal), some Balkan islands, Crete, and South America.19,20 European populations of Moorish geckos diverged from the central Morocco clade during the early Pliocene, around 4.14 Mya. The colonization of the Iberian Peninsula by Moorish geckos from North Africa could have occurred during the Messinian Salinity Crisis or after the opening of the Strait of Gibraltar during the early Pliocene.21
The Moorish gecko is active during diurnal and nocturnal hours, occupying several ecological niches, such as tree trunks, houses, and stone walls.22,23,24 This species represents an interesting example of phenotypically plastic variation, highlighted by its ability to adapt skin color according to the substrate.25 Two distinct and sympatric populations can be easily distinguished22 (Figure 1): the “dark diurnal gecko,” which mainly lives on trees, and the “pale nocturnal gecko,” found in human settlements on walls, especially near artificial lights, to increase the chances of prey capture.22,26,27 These two phenotypes are likely adapted to improve camouflage and are observed in many areas of the Mediterranean Basin (Figure S1). Experimental work showed that pale-nocturnal geckos are less exposed to predation on walls (their natural substrate) at night while dark-diurnal geckos are less attacked in trees (their natural substrate) during the day.22 The dark color of diurnal geckos may have primarily evolved to improve thermoregulation.28,29,30 We propose that as Neolithic humans spread across Europe, especially the Mediterranean, human-controlled fire played a significant ecological role, introducing a new foraging opportunity for Moorish geckos. This opportunity arose from positive phototactic arthropods being attracted to the light provided by fire. It is currently unknown whether nocturnal Moorish geckos are the ancestral form from which diurnal geckos have evolved, or vice versa. We argue that both evolutionary paths are conceivable and may be associated with the spread of human-generated fire. The following scenarios illustrate this association:
Figure 1.
Dark-diurnal (A) and pale-nocturnal (B) Moorish geckos (Tarentola mauritanica) morphotypes
The two lineages occupy distinct ecological niches: the diurnal form inhabits trees, while the nocturnal variant prefers building walls at night. Insets (C) and (D) show the plasticity of a dark-diurnal gecko changing skin color on different substrates (see25 for more details). The bar scale is representative of 1mm (see also Figure S1).
“Out of the dark” scenario: the ancestral form was nocturnal and largely exploited human-generated fire. However, Moorish geckos are highly territorial and aggressive,31,32 so only some dominant individuals controlled the food bonanza offered by the fire. Less competitive individuals were excluded from these profitable foraging sites and confined to peripheral sites whose food availability was impoverished by the “vacuum effect” exerted by the fire. The more phenotypically plastic geckos exploited their capacity to adapt their color to daytime and mitigate predation in daylight, so they conquered a novel diurnal niche, leaving the less plastic, dominant individuals associated with humans and their fires.
“Into the dark” scenario: the phenotypic variation in skin color shown by this species has set the scene for the evolution of a night-adapted lineage and led to the split between dark and pale geckos. The direct ancestors of the nocturnal form may have included particularly pale individuals whose coloration protected them from nocturnal predators22 while allowing them to exploit habitats around human-set fires in villages, near walls of the houses or rocky surfaces, as successful foraging sites. By attracting positively phototactic insect prey, illuminated surfaces may have offered a new foraging niche that the geckos have successfully exploited. This may have led to the evolution of a paler, nocturnal, and synanthropic lineage that has ever since accompanied our species and the creation of urban settlements up to modern times.
To explore these evolutionary scenarios, we formulated the following hypotheses and predictions:
-
(1)
Although the two forms, diurnal and nocturnal respectively, retain the ability to vary the level of darkness/lightness to some extent,22 nocturnal geckos are characterized by a more marked depigmentation resulting from their high adaptation to the nocturnal niche. We predict that the two forms will differ in skin reflectance in response to the blood levels of α-MSH (α-Melanocyte Stimulating Hormone) and the size and numerosity of melanophores, overall revealing distinct structural features between the lineages adapted to the two temporal niches.
-
(2)
During the Holocene, a wave of human populations from the Middle East spread in Europe via Anatolia.33,34,35 Studies based on radiocarbon suggest that this migration wave spread farming practices into the region, initiating the Neolithic revolution in Europe.36,37,38,39,40 The advent of farming practices brought about the widespread use of fire, potentially creating a unique foraging niche for a specialized gecko lineage. It is plausible that Moorish geckos and humans encountered each other in the Mediterranean region during this period, assuming they inhabited the same areas. Consequently, we predict that the common ancestor in the two gecko lineages occurred during sympatry with humans.
-
(3)
Fire may significantly increase prey availability when lit near wall surfaces by attracting arthropods eaten by the nocturnal gecko form. Preliminary work has established a clear diet difference between diurnal and nocturnal geckos, with the former mostly preying on Formicidae, and the latter feeding on moths, small dipterans, and orthopterans (pers. obs.). We, therefore, predict that such potential prey will be more abundant near fire than under dark conditions, acting as the control.
Results
Skin reflectance, dermal melanin content, and α-MSH levels differ between pale-nocturnal and dark-diurnal geckos
Pale-nocturnal geckos showed a significantly lower skin reflectance than dark-diurnal geckos (df = 1, 14; F = 62.37 p = 0.00001) (Figure 2A; Table S1). Accordingly, both dermal melanin production (df = 1, 16; F = 32.3; p = 0.00003) (Figure 2B; Table S2) and α-MSH’s blood levels (df = 1, 10; F = 22.63 p = 0.0007) (Figure 2C; Table S3) differed between forms.
Figure 2.
Comparisons between the two geckos’ morphotypes
Comparisons between dark-diurnal (“day”) and pale-nocturnal (“night”) Moorish geckos (Tarentola mauritanica) for (A) Integral under skin reflectance curve, (B) spectrophotometric skin melanin assay (ng/μg protein) and C) α-MSH hormone blood concentration (mg/mL). Data are presented as boxplots showing the mean (dotted line), median (bold line), interquartile range (box), and whiskers representing ±1 standard deviation. Individual data points are shown as dots. One-way ANOVA tests, ∗∗p < 0.001 (see also Tables S1–S3).
Pale-nocturnal geckos show smaller and less abundant melanophores than dark-diurnal individuals
The comparison of histological skin samples taken from the dorsal surface of pale-nocturnal and dark-diurnal Moorish geckos revealed clear differences, in agreement with our hypothesis. Dark phenotype geckos exhibited numerous, predominantly large melanophores. Additionally, a thick and continuous layer of pigment granules was observed in the connective tissue of the underlying dermis. In contrast, the skin of pale-nocturnal specimens displayed remarkably reduced melanophores in size and abundance, and the layer of pigment surrounding them was sparse (Figure 3).
Figure 3.
Skin sections stained with toluidine blue
(A) Skin from the dorsal area of dark-diurnal Moorish geckos (Tarentola mauritanica), with single melanophores (inset). A high concentration of melanophores (thick arrow) is localized in the dermal region, along with a thick layer of melanin granules (arrow).
(B) Digitally stained version of image A to emphasize melanophores and melanin pigment.
(C) Skin from the dorsal area of a pale-nocturnal gecko, with a single melanophore (inset).
(D) Digitally stained version of image C to emphasize melanophores and a layer of melanin pigment. The bar scale is representative of 50μm.
Significant Spatial Overlap Between Humans and Geckos at the Time of Gecko Divergence
The SDMs developed for humans and geckos were projected into the past to estimate their potential distribution in the Mediterranean area during the 0.06 Myr−0.01 Myr interval covering the estimated divergence time between the dark and pale gecko forms. These models exhibited high predictive performance. Specifically, the SDM developed for humans showed an AUC of 0.818 (S.D. = 0.014) and a TSS of 0.502 (S.D. = 0.028). Similarly, the SDMs developed for T. mauritanica achieved high performance, with an AUC of 0.96 (S.D. = 0.002) and a TSS of 0.803 (S.D. = 0.008). We assessed the spatial correlation between humans and the Moorish gecko in the past using the Boyce Index, obtaining a Spearman correlation value of 0.422. Furthermore, the linear mixed-effects model revealed a positive and significant (angular coefficient of the distribution, a = 0.6; p < 0.001) relationship between H. sapiens and T. mauritanica potential distributions (Figure S2). The niche overlap between T. mauritanica and H. sapiens was highest in Southern Iberia and north-western Africa (Figure 4). Additionally, the model exhibited a relevant age-driven random effect structure, with increasing overlaps toward more recent time intervals (p = 0.016, Figure 4).
Figure 4.
Past sympatry between Tarentola mauritanica and Homo sapiens
(A, B, and C) Spatial distribution models of T. mauritanica and H. sapiens and their overlap during the time interval 0.06 Myr−0.02 Myr. The models indicate significant overlap in potential distribution within the Mediterranean region, supported by a linear mixed-effect model demonstrating a positive relationship between potential distributions of the two species.
(D) Spatial distribution model of T. mauritanica approximately 10,000 years ago (Boyce Index, Spearman correlation test and the linear mixed-effects model p < 0.001).
(E) Distribution of the 317 ancient genomes of H. sapiens mainly from the Mesolithic and Neolithic periods (filled squares, modified from41) and further archaeological records (see also Figure S2; Table S4). The bar scale is representative of 500 km of distance.
Mitogenome analysis traces the origin of pale-nocturnal and dark-diurnal lineages approximately 6,600 years ago
The phylogenetic analysis based on the complete mitochondrial genome revealed that the Moorish gecko’s pale and dark populations are monophyletic, sharing a common ancestor estimated to have emerged approximately 6,600 years ago (Figure 5A). Samples from Calabria, Apulia, and northern Campania were identified as sister taxa to all pale-nocturnal and dark-diurnal geckos included in the analysis.
Figure 5.
Phylogenetic characterization
(A) Phylogeny and divergence time estimation derived from molecular-clock analysis of 10 pale-nocturnal (N1-N10) and 6 dark-diurnal (D1-D6) Tarentola mauritanica, in addition to other Italian haplotypes.42 Divergence times for the two gecko populations were calculated using the complete mitochondrial genome. The red bars on the nodes represent the 95% credibility intervals of the estimated posterior distributions of the divergence times. The bar scale is representative of 1 Kya, thousand years ago.
(B) DensiTree. All trees created in the analysis (except the burn-in phase) are displayed. Trees with the most common topology are highlighted in dark green, and trees with the second most common topology in light green.
The tree set with a dominant topology in the Bayesian analyses provided the distribution and origination time of the Most Recent Common Ancestor (tMRCA) (Figure 5B). All post-burn-in trees are shown with their estimated branch lengths and topologies, where the fuzziness of the horizontal plane of the branches visualizes variation in branch lengths across the trees. This reconstruction also demonstrates strong support for separating pale-nocturnal and dark-diurnal geckos, as indicated by the topology (dark green branches). In contrast, increased uncertainty in the tMRCA is represented by light green shading.
Fire-illuminated sites provide higher prey availability
The arthropods sampled at the walls illuminated by fire differed significantly from those recorded at the unlit walls (Figure 6). Specifically, we recorded higher numbers of arthropods at fire-illuminated sites considering the total number of individuals at the order level (two-tailed t-test = 2.50, d.f. = 9, p = 0.033) (Table 1). Samples were dominated by moths, mosquitos, grasshoppers, wasps, ants, and true bugs (see also Table S5).
Figure 6.
Characterization of trophic availability
Numbers of arthropods collected from walls exposed to fire (F) vs. walls not exposed to fire (NF).
(A) Sankey diagram connecting numbers of individuals collected by order with treatment.
(B) The total number of individuals captured under the two conditions is shown. Data are represented as boxplots with the mean (dotted line), median (bold line), interquartile range (box), and whiskers representing ±1 standard deviation. (two-tailed t-test; t = 2.50; d.f. = 9; p < 0.05) (see also Table S5).
Table 1.
Paired sample t-test results for comparison between arthropods collected at walls exposed to fire vs. unlit walls
Taxon | d.f. | t | p |
---|---|---|---|
Diptera | 63 | 4.75 | 0.00001a |
Orthoptera | 63 | 2.42 | 0.01840a |
Hymenoptera | 63 | 3.31 | 0.00152a |
Hemiptera | 63 | 4.39 | 0.00004a |
Lepidoptera | 63 | 2.69 | 0.00894a |
Coleoptera | 63 | 1.74 | 0.08638 |
Neuroptera | 63 | 1.27 | 0.20840 |
Arachnida | 63 | 0.38 | 0.70040 |
Plecoptera | 63 | 0.44 | 0.65827 |
Pseudoscorpionida | 63 | 1.00 | 0.32113 |
statistically significant difference (p < 0.05). d.f. = degrees of freedom.
Discussion
The existence of Moorish geckos that are active during the night or daytime, and their extreme color variation, have long been recognized in the scientific literature (e.g.,43,44,45,46,47,48). However, no study has explored the phylogenetic history and the adaptations of the two forms. Although focusing on a defined geographical area, our study can be considered representative of all other populations that feature these two morphotypes (Figure S1).
Did human-controlled fire influence the temporal niche partitioning in Moorish geckos?
We used a multidisciplinary approach to demonstrate that the two gecko variants have undergone phenotypic divergence over their evolutionary paths.
The pale-nocturnal form has smaller and fewer melanosomes and is associated with lower levels of α-MSH than the dark-diurnal form. Additionally, mitogenome analysis revealed that the pale-nocturnal lineage split from the dark-diurnal lineage around 6,600 years ago, making it unclear which of the two forms appeared first. However, molecular dating indicates that both forms diverged when Neolithic humans spread across Europe, coinciding with the use of fire in association with human settlement.
The area where we carried out our study is characterized by an abundance of caves long inhabited by humans (49,50). Charcoal analysis from the Cilento coast reveals a significant, human-driven shift in vegetation between 5600 and 7490 years ago (51 and G. Di Pasquale, pers. comm.) overlapping with the divergence time frame we estimated for the two gecko variants. The charcoal samples were obtained from stratified deposits inside caves, providing evidence of fire use within these sites. The presence of fire, charcoal remains and various archaeological materials, serves as clear evidence of human activity in these caves. Consequently, strong evidence suggests that humans extensively used caves and controlled fires in these underground sites, likely supporting our light-related evolutionary hypothesis.
This probably facilitated the evolution of the nocturnal lineage by enhancing the availability of suitable arthropod prey for pale geckos, thereby providing a new and previously unexploited foraging niche. Also, our spatial modeling exercise confirmed sympatry between humans and geckos when the two lineages arose. Finally, we showed that fire lit near walls mimicking a natural rock surface considerably increased the amount of arthropod food, attracting selectively insects such as moths, mosquitos, and wasps that are potential prey for the pale variant.
Overall, this multifaceted body of evidence suggests that ancient fires used by humans may have facilitated the separation between the two gecko variants leading to the situation we observe today. However, we could not fully clarify whether the two lineages originated from diurnal or nocturnal ancestors. While human-controlled fire may have represented an environmental novelty providing an important foraging niche at night, the evidence we gathered is still compatible with either scenario. The presence in Italy of geckos since the Upper Pleistocene is supported by fossil remains43 and excludes its historical, anthropogenic introduction in the country, in agreement with previous phylogenetic work.44
The timing of human control over fire remains a contentious issue.45,46 Evidence from Europe suggests that Neanderthals and early modern humans already used controlled and consistent fire starting from the Middle Pleistocene onwards in Europe, and from the Late Pleistocene in Europe and South Africa, for purposes beyond warmth and cooking.47,48 However, the hypothesis that controlled fires became prevalent upon the arrival of modern humans in Europe, particularly in the Mediterranean, is highly plausible.
We verified that the phenotypic divergence between the two gecko variants may have occurred broadly across the Mediterranean. However, we interpret our results cautiously, as the genetic samples used in this analysis are restricted to our study area. Expanding the sampling to include other Mediterranean regions (e.g., Spain, Portugal, France) would help clarify whether this phenomenon reflects a broader pattern or is a result of recent local adaptation. Such an expanded analysis should adopt the same multidisciplinary approach we used to investigate the two populations in our study area. Indeed, phylogenetic analyses incorporating sequences from outside our study area would either complicate or reinforce the hypothesis of a local phenomenon. Moreover, in a pan-Mediterranean analysis, it is important to consider that this species has been widely translocated by humans, often unintentionally.
The “out of the dark” scenario
The hypothesis of a nocturnal ancestor that has led to a first pale-nocturnal lineage and later to a diurnal descendant is certainly parsimonious and agrees with the fact that based on a phylogenetic analysis of temporal activity patterns, nocturnality would be an ancient trait appearing at the base of the gecko tree.52
Geckos and skinks independently evolved nocturnality, diverging from the ancestral diurnality typical in lizards,53 with noticeable metabolic54,55 and sensory56 adaptations to the nocturnal niche. While the spread of human-controlled fire may have represented a strong advantage for more night-specialized phenotypes, these may have monopolized such preferred foraging sites excluding other less competitive individuals. In behavioral studies of T. mauritanica foraging near artificial lights, large individuals exclude smaller and younger geckos from key foraging sites,57 and a similar situation is conceivable near the fires set by ancient humans. On the other hand, living on the periphery of such sites for subordinate geckos constrained in their marginal territories may have been highly unprofitable because of the “vacuum effect” caused by the fire, attracting arthropod food and depleting the surrounding areas. This situation is well-known today in response to artificial lighting. It is regarded as one of the adverse effects of light pollution on natural environments, favoring a few opportunistic, light-tolerant predators at the expense of light-averse species whose dark habitat is continuously impoverished.10,58 The process has been proposed as the way light-tolerant common pipistrelles (Pipistrellus pipistrellus) have outcompeted light-averse lesser horseshoe bats (Rhinolophus hipposideros) in Switzerland, contributing to the decline of the latter species.59 Under such a scenario, the most phenotypically plastic individual geckos, capable of acquiring a sufficiently dark color to shift their temporal niche to the daytime, might have led to the evolution of the diurnal lineage observed today, escaping the strong competition by the nocturnal dominants.
The “into the dark” scenario
The alternative “into the dark” scenario we considered would originate from a dark-diurnal ancestor56,60 from where the pale-nocturnal lineage, and later, the dark-diurnal lineage would evolve. During the evolutionary path leading to the current nocturnal specialists, an ancestral dark form would have lost the ability to conspicuously change its coloration, fixing a pale phenotype that would effectively mitigate predatory pressure at night.22 The attainment of the pale phenotype involved a reduction in α-MSH hormone levels and alterations in the number and size of melanosomes in the dermis. In other words, predators would have exerted strong selective pressure, killing all individuals whose coloration was not pale enough to fade their contour on rock walls at night, leaving only the palest ones on the scene. The latter individuals were probably those exhibiting a greater ability to lighten their color,25,61 and subsequently underwent depigmentation adapting to nocturnal life. The paleo-synanthropic relationship between geckos and humans has persisted into modern times. Today the pale-nocturnal gecko inhabits urban areas, often waiting for prey near artificial light sources. Although nocturnal lizards such as geckos may be active at body temperatures that are considerably lower than those characterizing activity in diurnal lizards,62,63 nocturnal species may still experience suboptimal locomotion in the cold of the night,62,63,64 which might, in theory, limit successful foraging. However, locomotion in nocturnal lizards is surprisingly efficient, as they outrun 3-fold diurnal lizards and show low-temperature performances like those of diurnal lizards at higher temperatures. This efficiency is attributed to a low minimum cost of locomotion62,65 and higher metabolic rates at low temperatures.54,55 In the case of pale geckos, hunting near fires might have represented a further way of warming up and achieving even higher locomotion performances.
The role of phenotypic plasticity in promoting the diversification of new lineages has been reconsidered in recent times,66 and today is seen as potentially important albeit its mechanisms are still debated (e.g.,.67,68,69,70,71,72 Phenotypic plasticity might serve as a protective mechanism against environmental fluctuations, fostering the persistence of populations.66,73,74 On the other hand, phenotypic plasticity could lay the groundwork for a process called “genetic assimilation,” where a phenotype becomes integrated into the genotype, potentially resulting in a loss of plasticity known as “canalization.”75 Under the “into the dark” scenario, the plasticity still observed in the dark-diurnal lineage and hypothetically shared with its diurnal ancestor facilitated the initial colonization and survival of the population in a new environment, allowing time for subsequent genetic adaptation to fine-tune responses to this environment. The plastic capabilities of the phenotype can thus respond to selection, anticipating the adaptation process of the genotype and enabling the colonization of the new nocturnal niche.25,61
Current artificial lighting at night offers numerous tests of how animals, including geckos, respond to novel light sources.12 For instance, six gecko species from the genus Phelsuma, although mostly diurnal, shifted their activity from diurnal to nocturnal hours to exploit prey concentrating near artificial lights,76 a behavioral change closely resembling that we propose as a potential origin of the Moorish gecko’s nocturnal lineage.
The influence of human-controlled fire on temporal niche partitioning in Moorish geckos
In conclusion, our investigation into the potential influence of human-controlled fire on the temporal niche partitioning in Moorish geckos presents intriguing insights but also leaves some questions unanswered. The multifaceted evidence we gathered suggests that ancient fires used by humans may have played a role in facilitating the separation between the two gecko variants observed today. However, it does not definitively clarify whether the two lineages originated from diurnal or nocturnal ancestors. The “out of the dark” scenario posits a parsimonious hypothesis where a nocturnal ancestor led to a first nocturnal lineage, followed later by a diurnal descendant, whereas the “into the dark” scenario suggests an alternative origin, where a dark-diurnal ancestor gave rise to a less plastic, more specialized pale-nocturnal lineage. Phenotypic plasticity emerges as a crucial factor in both scenarios, potentially facilitating the colonization of new environments and the subsequent fine-tuning of genetic adaptations.
The ongoing impact of artificial lighting on nocturnal behavior underscores the relevance of our findings, drawing parallels to the adaptation of the Moorish gecko’s nocturnal lineage. Future research could explore further the mechanisms underlying phenotypic plasticity and its interplay with genetic adaptation in this species, shedding light on the evolutionary dynamics of temporal niche partitioning in response to past and current human influences.
Limitations of the study
Although our work provides evidence supporting the hypothesis of anthropogenic light conditioning in the separation of dark-diurnal and pale-nocturnal gecko populations, it is necessary to clarify which of the two hypotheses (into-the-dark or out-of-the-dark) is the most plausible. Further investigation is needed into the phylogenetics, physiology, and ecology of nocturnal and diurnal forms. Furthermore, while the pattern we described is fascinating and could have potentially occurred in any location with both human communities and geckos in the Mediterranean Basin, it should be verified in additional areas where well-differentiated dark-diurnal and pale-nocturnal populations have been observed. Based on our data, however, we cannot speculate on the existence of such a phenomenon outside the study area.
Resource availability
Lead contact
Further information and requests for resources and reagents should be directed to and will be fulfilled by the lead contact, Danilo Russo (danrusso@unina.it).
Materials availability
This study did not generate new reagents.
Data and code availability
-
•
The sequencing data are available at National Center for Biotechnology Information (NCBI) as GenBank: MK275668 - MK275678, MK275681, MK275682, MK275684 - MK275686, JQ425045 - JQ425050, JQ425055, JQ425056, JQ425060 and are publicly available as of the date of publication.
-
•
This article does not report original code.
-
•
Any additional information required to reanalyze the data reported in this article is available from the lead contact upon request and are also directly available in the supplementary materials of this article as the distribution of polymorphic gecko populations; phenotypic data on the examined geckos; Homo sapiens past and present habitat suitability map; arthropods captured with and without light.
Acknowledgments
We thank Greger Larson for his suggestions on a preliminary version of our text. Gaetano Di Pasquale shared important information regarding charcoal analysis and human presence in the study area. We are indebted to two anonymous reviewers for their valuable comments on the first article version.
Author contributions
Conceptualization: D.F. and M.B.; sample collection: D.F., E.R., V.M., and M.B; laboratory experiments: E.R., V.M., B.A., and M.B; Analyses of data: D.F., D.R., V.M., E.R., B.A., A.M., G.G., and M.B; supervision: D.F.; writing original draft: D.F., E.R., and M.B.; review and editing: D.F., D.R., E.R., V.M., B.A., A.M., G.G., and M.B. All the authors read and approved the article.
Declaration of interests
The authors declare no conflicts of interest.
STAR★Methods
Key resources table
REAGENT or RESOURCE | SOURCE | IDENTIFIER |
---|---|---|
Chemicals | ||
MS-222 | Sigma-Aldrich | Cat#E10521-10g |
Epon 812 resin | Fluka | |
Toluidine blue | Sigma-Aldrich | Cat#T3260-5g |
Sodium tetraborate buffer solution | Sigma-Aldrich | Cat#B9876-500g |
Tris-HCl pH 7.4 | Thermo Scientific | Cat#J60202.K2 |
NaCl | Sigma-Aldrich | Cat#S9888-25G |
NP-40 | Thermo Scientific | Cat#13434269 |
Protease inhibitors | Roche | Cat#4693116001 |
Deposited data | ||
Complete mitochondrial sequence | Fulgione et al., 201922 and this study | National Center for Biotechnology Information (NCBI) accession numbers MK275668 to MK275678, MK275681 - MK275682, MK275684 to MK275686 |
Partial mitochondrial sequence | Rato et al., 202323 | National Center for Biotechnology Information (NCBI) accession numbers JQ425045 to JQ425050, JQ425055, JQ425056, JQ425060 |
Software and algorithms | ||
Bayesian Evolutionary Analysis Sampling Trees (BEAST) | Bouckaert et al., 201977 | |
Bayesian Evolutionary Analysis Utility (BEAUti) | Drummond et al., 201278 | |
Molecular Evolutionary Genetics Analysis version 11 (MEGA11) | Tamura et al., 202179 | |
Global Biodiversity Information Facility (GBIF) database | www.GBIF.org | |
R software environment version 4.10 | R Core Team 2020 | https://www.R-project.org/ |
iNaturalist | https://www.inaturalist.org/ | |
Tracer v.1.6 | Rambaut et al., 201880 | |
Treeannotator v2.7.7 | Drummond et al., 201278 | |
FigTree v.1.4.4 | Rambaut 201081 | http://tree.bio.ed.ac.uk/software/figtree/ |
DensiTree v.2.7.7 | Bouckaert et al., 201082 | |
LogCombiner 2.7.7 | Drummond et al., 201278 | |
Other | ||
Heparinised syringe | N/A | |
Yellow sticky traps | N/A | |
Super Nova Leica Ultratome | Leica Microsystem | |
Leica EZ4 W stereomicroscope | Leica Microsystem | |
Zeiss Axioskop 5 microscope | Zeiss | Cat#490980-0001-000 |
Zeiss Axiocam camera | Zeiss | Cat#426560-9061-000 |
AvaSpec-2048-USB2-UA-50 | Avantes | |
Halogen & Deuterium Halogen light sources | Avantes | AvaLight-DH-S |
Reflection probe | Avantes | FCR-7UV200-2-ME |
White reference tile WS2 | Avantes |
Method details
Study area and samples collection
The study was conducted in Southern Italy (40°15′N, 14°54′E, Salerno province) in a region characterised by Mediterranean vegetation, predominantly featuring olive groves, rural structures, and stone walls.
To perform laboratory procedures, we sampled 16 individuals: 6 dark-diurnal geckos collected from tree trunks, and 10 pale-nocturnal geckos captured on walls, using nylon loops. The subjects were then released at the capture site, except for six individuals (three dark-diurnal geckos and three pale-nocturnal geckos) sacrificed to obtain tissues suitable for histological analyses. We sampled Moorish geckos with the approval of the relevant nature conservation authority (Cilento, Vallo di Diano e Alburni National Park, protocol 2013/0010678). The experimental protocols were approved by the Ethical Committee for Animal Experiments of the first author’s University (protocol 2013/0032826).
Skin histology
We anesthetized three dark-diurnal and three pale-nocturnal geckos with 250 mg/kg 1% MS-222 (Sigma Chemical Co. St. Louis, MO) injected into the intracoelomic cavity and euthanised them by decapitation. Following this, we processed skin samples taken from the backs of three dark and three light geckos for light microscopy, following the methods outlined in.83 In brief, we fixed the samples in 2% paraformaldehyde and 2.5% glutaraldehyde (for 4 h at 4°C) and subsequently post-fixed them in a 2% osmium tetroxide solution (for 1 h at 4°C). After dehydration in an ascending series of ethyl alcohol, the samples were embedded in Epon 812 resin (Fluka). We then cut semi-thin sections of 1.5 μm thickness using a Super Nova Leica Ultratome and stained them with 1% toluidine blue in 1% sodium tetraborate buffer. Finally, we examined the sections using a Zeiss Axiocam camera attached to a Zeiss Axioskop microscope (Zeiss, Jena, Germany).
Skin reflectance
To objectively determine the skin coloration of the sampled Moorish geckos, we measured skin reflectance using spectrophotometry (250–1000 nm, AvaSpec-2048-USB2-UA-50; Avantes, Apeldoorn, Netherlands) of the 16 geckos, focusing on the range between 300 and 700 nm.25 We employed a white reference tile (WS2; Avantes) as a calibration reference.
The spectrophotometer probe, featuring a 0.2 mm hole end, was positioned perpendicular to the animals' body surface, and the reflectance (R%) was recorded at three locations on their backs. Subsequently, the average of the integrals subtending the reflectance curves was considered representative of the entire back of each individual.22,25,84 We assessed the difference between the dark-diurnal and pale-nocturnal groups using a one-way ANOVA test, with statistical significance defined as p < 0.05.
Melanocyte Stimulating Hormone and melanin assay
Plasma and skin samples were obtained from three pale-nocturnal and three dark-diurnal geckos.
Blood was drawn using a heparinised syringe from the interdiscal vertebra windows of the tail. Plasma was then isolated by centrifugation at 2000g for 10 min at room temperature, and the samples were stored at −80°C until processing in the laboratory.
We determined the levels of α-MSH using the ELISA assay described by.71 Statistical significance differences (defined with p < 0.05) between the two groups were assessed using an ANOVA. We followed the protocol proposed by,85 with some modifications, to quantify melanin. In brief, tissues were collected in 50 mM Tris-HCl pH 7.4, 300 mM NaCl, 0.5% NP-40, and protease inhibitors (Roche). Cells were lysed using a combination of freeze-thawing (3 cycles of dry ice at −37°C), Dounce homogenisation (200 strokes), and sonication (2 min, 10 s on and 10 s off).
Phylogenetic characterisation and time tree
To infer the divergence time of pale-nocturnal and dark-diurnal Moorish gecko populations, we first aligned the mtDNA sequences of T. mauritanica from the 6 dark-diurnal (accession number MK275676 to MK275678, MK275684 to MK275686) and the 10 pale-nocturnal specimens (accession number MK275668 to MK275675, MK275681, MK275682) with T. mauritanica from Central and Southern Morocco (accession number JQ425056-Marrakech and JQ425055-Anezal),42 Iberian Peninsula (accession number JQ425045 - Rambla del Cañar_ Murcia),42 and Southern Italy (accession number JQ425046 to JQ425050 and JQ425060),21,42 using “complete deletion” strategies in Molecular Evolutionary Genetics Analysis v.11 (MEGA11).79 Then, we constructed a Bayesian phylogeny in Bayesian Evolutionary Analysis Sampling Trees (BEAST) v.1.7.577 with T. mauritanica JQ425056-Marrakech as outgroup. We considered General Time Reversible (GTR) substitution model as suggested by “find best DNA model” option in MEGA11.79 We used a strict clock model with the default parameters and default operators set in Bayesian Evolutionary Analysis Utility (BEAUti)78 to estimate the divergence times.
Because no suitable fossil record exists for calibrating the mutation rate in T. mauritanica, we adopted a substitution rate of the mtDNA of lizards (5.29 × 10−9).86 Each Markov chain Monte Carlo (MCMC) sample was based on a run of 100,000,000 generations and sampled every 1,000 generations. The first 10% of samples were treated as burn-in and removed before individual runs were combined in LogCombiner v.2.7.7.78 Stationarity was assessed using the program Tracer v.1.680 with ESS values >200 taken as evidence for convergence. A final MCMC maximum clade credibility tree was generated from the cumulative post-burn-in sample of the combined analyses in the program Treeannotator v2.7.7.78 The resulting consensus tree was visualized using FigTree v.1.4.481 and the distribution of the trees was visualized with DensiTree v.2.7.7,82 using the same parameters described above, with burn-in of 10%.
Potential human-gecko co-occurrence
To establish whether Moorish geckos and humans coexisted when the dark-diurnal and pale-nocturnal geckos split, we generated Species Distribution Models (SDM) for T. mauritanica and projected them to the past. We collected modern occurrence data for the gecko from the Global Biodiversity Information Facility (GBIF) database by selecting “human observation” and “machine observation” as the basis of the record (www.GBIF.org, 26 January 2024; GBIF Occurrence Download https://doi.org/10.15468/dl.xprybx). The data were further filtered by excluding occurrences without geographical coordinates. For the human dataset, we used the H. sapiens records published by87 and,88 focusing on fossil presence observations dated within a time window coherent with the divergence time estimates between the gecko lineages we found (see below). Radiocarbon data were calibrated using the “Bchron R” package89 through the “intcal20” curve90 in R environment.91 As environmental predictors, we employed the monthly bioclimatic variables generated through the 2Ma CESM1.2 simulation,88 downscaled at a 0.5 ° × 0.5 ° grid resolution. The native set of predictors was converted into bioclimatic variables according to WorldClim using the “dismo” R package.92 Lastly, variables were projected onto the Mollweide coordinate reference system. To prevent model overfitting, duplicate occurrences in the raster grid were removed. After this step, we gathered 520 modern T. mauritanica and 740 past H. sapiens occurrences.
We defined the Mediterranean region as the study area for both H. sapiens and T. mauritanica, according to the current and historical geographical distribution of the Moorish gecko.21 Then, we randomly generated 10,000 background points within the study area. To account for potential sampling biases, pseudoabsences were geographically placed according to the density of the occurrence data, making them more abundant where presences are denser.93,94,95 For H. sapiens only, the record was divided into 1000-year consecutive time bins according to the time resolution of bioclimatic predictors. Subsequently, we partitioned the pseudoabsences proportionally to the number of presences per time bin. After extracting climatic values at each occurrence and pseudoabsence data point, we accounted for multicollinearity among predictors by considering the Variance Inflation Factor (VIF). The latter was assessed using the function “vifcor” embedded in the “usdm” R package, selecting a threshold of 0.75.96,97 After applying VIF, the selected predictors were BIO4 (temperature seasonality), BIO8 (mean temperature of the wettest quarter), BIO9 (mean temperature of the driest quarter), BIO13 (Precipitation of Wettest Month), BIO14 (precipitation of the driest month), BIO15 (Precipitation Seasonality), and BIO19 (precipitation of the coldest quarter).
To estimate the geographic distribution of both species in the past, we adopted the SDM ensemble approach. Specifically, we trained SDMs through an ensemble forecasting approach, using the R package “biomod2”.98 We considered four different algorithms: Maximum Entropy Models (MaxEnt), Generalized Boosted Models (GBM), Random Forests (RF), and Generalized Linear Models (GLM). We adopted the default settings described in the “biomod2” R package for model tuning. To evaluate the predictive accuracy of SDMs, we randomly split the dataset into a 70% sample used for model calibration and the remaining 30% used to assess model performance. Then, we calculated the area under the receiver operating characteristic curve (AUC;99) and the true skill statistic (TSS;100). This procedure was iterated 10 times, changing the randomly selected training/testing data points at each iteration. Model averaging was performed by weighting the individual model projections by their AUC values and averaging the results,101 excluding the model with AUC <0.7. Lastly, SDM predictions were projected into the past to obtain the potential distribution of T. mauritanica in the Mediterranean area during the time interval predicted by the phylogenetic reconstruction as the divergence time between the two Moorish gecko forms.
To test whether there was a significant spatial correlation between H. sapiens and T. mauritanica, we adopted the Boyce Index.98 This index requires only presence data points and measures how much model predictions differ from a random distribution of observed presences across prediction gradients.99 The Boyce Index ranges from −1 to +1, with positive values indicating a model whose predictions are consistent with the distribution of presences in the evaluation dataset. Values close to zero mean that the model does not differ from a random model, and negative values indicate counter predictions, i.e., predicting poor-quality areas where presences are more frequent.98
We used the Boyce Index to test whether the predicted suitability for T. mauritanica is higher in the presence of humans than otherwise throughout the study area. In addition to the Boyce Index, we trained a Linear Mixed-Effects Model to test whether the suitability of the two species is associated, setting the variable “time” as a random effect. This was done using the “lme4” R package.102 To provide a visual rendering of the potential spatial overlap between the two species across space and time, we first binarised the individual species prediction maps by adopting the threshold that maximises the sum of sensitivity and specificity (“MaxSens + Spec”). We then stacked the binary maps of the two species at each 1 kyr according to the time resolution of the bioclimatic variables. Map binarisation was performed using the “PresenceAbsence” R package.103
Testing differences in prey availability due to fire
We conducted a paired-design experiment to assess the power of fire in attracting arthropods at night, affecting food availability for Moorish geckos. We selected four light-coloured walls, resembling rock surfaces, and compared the number of arthropods lying on each wall over 5 hours when fire was set nearby vs a “dark” (no fire) control. The distance of the fire from the wall was carefully calibrated to prevent excessive heating while ensuring a broad surface of light. The flame never exceeded 40 cm in height, was 30 cm wide and was fuelled by dry wood collected from the surrounding area. The night of treatment was randomly assigned.
To collect arthropods comprehensively, we hung to each wall fifteen 15x20cm yellow sticky traps regularly spaced over the wall surface. Besides, we also collected arthropods lying in the spaces between sticky traps using hand nets and tweezers. We identified all collected arthropods using a Leica EZ4 W stereomicroscope (Leica Microsystems) following104,105 and using local reference collections. Identification was done at the order level and all individuals were counted. We used a paired-sample t-test to compare the number of individuals captured by order between lit vs unlit walls. The results was considered statistically significant with p<0.05.
Quantification and statistical analysis
To test differences in skin reflectance, dermal melanin content and α-MSH plasmatic levels between pale-nocturnal and dark-diurnal geckos, we performed an one-way ANOVA test in R (version 4.3.0) considering p<0.05 as statistically significant. The results of significance are reported in the main text (Result section) and legend of Figure 2.
To evaluate the spatial correlation between humans and the Moorish gecko in the past, we used the Boyce Index, the Spearman correlation test and the linear mixed-effects model, with p<0.001 considered statistically significant. The results are reported in Figure 4.
To test the differences in arthropods sampled at unlit versus fire-lit walls, we conducted a paired two-tailed t-test, with p<0.05 considered statistically significant. The results are reported in the legend of Figure 6 and in Table 1.
Published: December 30, 2024
Footnotes
Supplemental information can be found online at https://doi.org/10.1016/j.isci.2024.111715.
Supplemental information
AMS, Accelerator mass spectrometry; 14C, Radioncarbon dating; OSL, Optical stimulated luminescence.
References
- 1.Cowie R.H., Bouchet P., Fontaine B. The Sixth Mass Extinction: fact, fiction or speculation? Biol. Rev. 2022;97:640–663. doi: 10.1111/brv.12816. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Lyons S.K., Amatangelo K.L., Behrensmeyer A.K., Bercovici A., Blois J.L., Davis M., DiMichele W.A., Du A., Eronen J.T., Faith J.T., et al. Holocene shifts in the assembly of plant and animal communities implicate human impacts. Nature. 2016;529:80–83. doi: 10.1038/nature16447. [DOI] [PubMed] [Google Scholar]
- 3.Braje T.J., Erlandson J.M. Human acceleration of animal and plant extinctions: A Late Pleistocene, Holocene, and Anthropocene continuum. Anthropocene. 2013;4:14–23. doi: 10.1016/j.ancene.2013.08.003. [DOI] [Google Scholar]
- 4.Hulme-Beaman A., Dobney K., Cucchi T., Searle J.B. An Ecological and Evolutionary Framework for Commensalism in Anthropogenic Environments. Trends Ecol. Evol. 2016;31:633–645. doi: 10.1016/j.tree.2016.05.001. [DOI] [PubMed] [Google Scholar]
- 5.McKinney M.L. Effects of urbanization on species richness: A review of plants and animals. Urban Ecosyst. 2008;11:161–176. doi: 10.1007/s11252-007-0045-4. [DOI] [Google Scholar]
- 6.Ålund M., Cenzer M., Bierne N., Boughman J.W., Cerca J., Comerford M.S., Culicchi A., Langerhans B., McFarlane S.E., Möst M.H., et al. Anthropogenic Change and the Process of Speciation. Cold Spring Harbor Perspect. Biol. 2023;15:a041455. doi: 10.1101/cshperspect.a041455. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Bennie J., Davies T.W., Cruse D., Gaston K.J. Ecological effects of artificial light at night on wild plants. J. Ecol. 2016;104:611–620. doi: 10.1111/1365-2745.12551. [DOI] [Google Scholar]
- 8.Gaston K.J., Bennie J., Davies T.W., Hopkins J. The ecological impacts of nighttime light pollution: a mechanistic appraisal. Biol. Rev. 2013;88:912–927. doi: 10.1111/brv.12036. [DOI] [PubMed] [Google Scholar]
- 9.Rowse E.G., Harris S., Jones G. The Switch from Low-Pressure Sodium to Light Emitting Diodes Does Not Affect Bat Activity at Street Lights. PLoS One. 2016;11 doi: 10.1371/journal.pone.0150884. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Stone E.L., Harris S., Jones G. Impacts of artificial lighting on bats: a review of challenges and solutions. Mamm. Biol. 2015;80:213–219. doi: 10.1016/j.mambio.2015.02.004. [DOI] [Google Scholar]
- 11.Van Langevelde F., Ettema J.A., Donners M., WallisDeVries M.F., Groenendijk D. Effect of spectral composition of artificial light on the attraction of moths. Biol. Conserv. 2011;144:2274–2281. doi: 10.1016/j.biocon.2011.06.004. [DOI] [Google Scholar]
- 12.Rich C., Longcore T., editors. Ecological consequences of artificial night lighting. Island Press; 2006. [Google Scholar]
- 13.Stepka Z., Azuri I., Horwitz L.K., Chazan M., Natalio F. Hidden signatures of early fire at Evron Quarry (1.0 to 0.8 Mya) Proc. Natl. Acad. Sci. USA. 2022;119 doi: 10.1073/pnas.2123439119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Zohar I., Alperson-Afil N., Goren-Inbar N., Prévost M., Tütken T., Sisma-Ventura G., Hershkovitz I., Najorka J. Evidence for the cooking of fish 780,000 years ago at Gesher Benot Ya’aqov, Israel. Nat. Ecol. Evol. 2022;6:2016–2028. doi: 10.1038/s41559-022-01910-z. [DOI] [PubMed] [Google Scholar]
- 15.Shimelmitz R., Kuhn S.L., Jelinek A.J., Ronen A., Clark A.E., Weinstein-Evron M. ‘Fire at will’: The emergence of habitual fire use 350,000 years ago. J. Hum. Evol. 2014;77:196–203. doi: 10.1016/j.jhevol.2014.07.005. [DOI] [PubMed] [Google Scholar]
- 16.Bowman D.M.J.S., Balch J., Artaxo P., Bond W.J., Cochrane M.A., D’Antonio C.M., DeFries R., Johnston F.H., Keeley J.E., Krawchuk M.A., et al. The human dimension of fire regimes on Earth: The human dimension of fire regimes on Earth. J. Biogeogr. 2011;38:2223–2236. doi: 10.1111/j.1365-2699.2011.02595.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Gowlett J.A.J. The discovery of fire by humans: a long and convoluted process. Phil. Trans. R. Soc. B. 2016;371 doi: 10.1098/rstb.2015.0164. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Rieppel O. Handbuch der Reptilien und Amphibien Europas (Akademische Verlagsgesellschaft) 1981. Tarentola mauritanica(Linnaeus, 1758) - Wiesbade; pp. 119–133. [Google Scholar]
- 19.Carranza S., Arnold E.N., Mateo J.A., López-Jurado L.F. Long-distance colonization and radiation in gekkonid lizards, Tarentola (Reptilia: Gekkonidae), revealed by mitochondrial DNA sequences. Proc. Biol. Sci. 2000;267:637–649. doi: 10.1098/rspb.2000.1050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.IUCN Tarentola mauritanica: Vogrin, M., Corti, C., Pérez Mellado, V., Baha El Din, S. & Martínez-Solano, I.: The IUCN Red List of Threatened Species 2017: e.T61578A63716927. 2017. [DOI]
- 21.Rato C., Carranza S., Harris D.J. Evolutionary history of the genus Tarentola (Gekkota: Phyllodactylidae) from the Mediterranean Basin, estimated using multilocus sequence data. BMC Evol. Biol. 2012;12:14. doi: 10.1186/1471-2148-12-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Fulgione D., Buglione M., Rippa D., Trapanese M., Petrelli S., Monti D.M., Aria M., Del Giudice R., Maselli V. Selection for background matching drives sympatric speciation in Wall Gecko. Sci. Rep. 2019;9:1288. doi: 10.1038/s41598-018-37587-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Rato C., Silva-Rocha I., Sillero N. What does the future hold for a thermophilic and widely introduced gecko, Tarentola mauritanica (Squamata: Phyllodactylidae)? Biol. Invasions. 2023;26:1061–1074. doi: 10.1007/s10530-023-03229-7. [DOI] [Google Scholar]
- 24.Valverde J.A. Junta de Andalucía; 2004. Estructura de una comunidad mediterránea de vertebrados terrestres (Consejería de Medio Ambiente. [Google Scholar]
- 25.Vroonen J., Vervust B., Fulgione D., Maselli V., Van Damme R. Physiological colour change in the Moorish gecko, Tarentola mauritanica(Squamata: Gekkonidae): effects of background, light, and temperature. Biol. J. Linn. Soc. Lond. 2012;107:182–191. doi: 10.1111/j.1095-8312.2012.01915.x. [DOI] [Google Scholar]
- 26.Arnold N. New edition. Harper Collins; 2002. Collins Field Guide to the Reptiles and Amphibians of Britain and Europe Collins. [Google Scholar]
- 27.Rato C., Carretero M.A. Ecophysiology Tracks Phylogeny and Meets Ecological Models in an Iberian Gecko. Physiol. Biochem. Zool. 2015;88:564–575. doi: 10.1086/682170. [DOI] [PubMed] [Google Scholar]
- 28.Clusella Trullas S., Van Wyk J.H., Spotila J.R. Thermal melanism in ectotherms. J. Therm. Biol. 2007;32:235–245. doi: 10.1016/j.jtherbio.2007.01.013. [DOI] [Google Scholar]
- 29.Clusella-Trullas S., van Wyk J.H., Spotila J.R. Thermal benefits of melanism in cordylid lizards: a theoretical and field test. Ecology. 2009;90:2297–2312. doi: 10.1890/08-1502.1. [DOI] [PubMed] [Google Scholar]
- 30.McNamara M.E., Rossi V., Slater T.S., Rogers C.S., Ducrest A.-L., Dubey S., Roulin A. Decoding the Evolution of Melanin in Vertebrates. Trends Ecol. Evol. 2021;36:430–443. doi: 10.1016/j.tree.2020.12.012. [DOI] [PubMed] [Google Scholar]
- 31.Lisičić D., Drakulić S., Herrel A., Đikić D., Benković V., Tadić Z. Effect of competition on habitat utilization in two temperate climate gecko species. Ecol. Res. 2012;27:551–560. doi: 10.1007/s11284-011-0921-5. [DOI] [Google Scholar]
- 32.Salvador A. Salamanquesa común. Tarentola mauritanica. Enciclopedia Virtual de los Vertebrados Españoles. 2016. http://www.vertebradosibericos.org/
- 33.Sikora M., Carpenter M.L., Moreno-Estrada A., Henn B.M., Underhill P.A., Sánchez-Quinto F., Zara I., Pitzalis M., Sidore C., Busonero F., et al. Population Genomic Analysis of Ancient and Modern Genomes Yields New Insights into the Genetic Ancestry of the Tyrolean Iceman and the Genetic Structure of Europe. PLoS Genet. 2014;10 doi: 10.1371/journal.pgen.1004353. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Lazaridis I., Patterson N., Mittnik A., Renaud G., Mallick S., Kirsanow K., Sudmant P.H., Schraiber J.G., Castellano S., Lipson M., et al. Ancient human genomes suggest three ancestral populations for present-day Europeans. Nature. 2014;513:409–413. doi: 10.1038/nature13673. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Lazaridis I., Nadel D., Rollefson G., Merrett D.C., Rohland N., Mallick S., Fernandes D., Novak M., Gamarra B., Sirak K., et al. Genomic insights into the origin of farming in the ancient Near East. Nature. 2016;536:419–424. doi: 10.1038/nature19310. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Ammerman A.J., Cavalli-Sforza L.L. Measuring the Rate of Spread of Early Farming in Europe. Man. 1971;6:674. doi: 10.2307/2799190. [DOI] [Google Scholar]
- 37.Silva F., Steele J. New methods for reconstructing geographical effects on dispersal rates and routes from large-scale radiocarbon databases. J. Archaeol. Sci. 2014;52:609–620. doi: 10.1016/j.jas.2014.04.021. [DOI] [Google Scholar]
- 38.Fort J. Demic and cultural diffusion propagated the Neolithic transition across different regions of Europe. J. R. Soc. Interface. 2015;12 doi: 10.1098/rsif.2015.0166. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Pinhasi R., Fort J., Ammerman A.J. Tracing the Origin and Spread of Agriculture in Europe. PLoS Biol. 2005;3 doi: 10.1371/journal.pbio.0030410. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Vander Linden M., Silva F. Comparing and modeling the spread of early farming across Europe. PAGES Mag. 2018;26:28–29. doi: 10.22498/pages.26.1.28. [DOI] [Google Scholar]
- 41.Allentoft M.E., Sikora M., Refoyo-Martínez A., Irving-Pease E.K., Fischer A., Barrie W., Ingason A., Stenderup J., Sjögren K.-G., Pearson A., et al. Population genomics of post-glacial western Eurasia. Nature. 2024;625:301–311. doi: 10.1038/s41586-023-06865-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Rato C., Carranza S., Perera A., Harris D.J. Evolutionary patterns of the mitochondrial genome in the Moorish gecko, Tarentola mauritanica. Gene. 2013;512:166–173. doi: 10.1016/j.gene.2012.09.032. [DOI] [PubMed] [Google Scholar]
- 43.Villa A., Delfino M. Fossil lizards and worm lizards (Reptilia, Squamata) from the Neogene and Quaternary of Europe: an overview. Swiss J. Palaeontol. 2019;138:177–211. doi: 10.1007/s13358-018-0172-y. [DOI] [Google Scholar]
- 44.Belluardo F., Pellitteri-Rosa D., Cocca W., Liuzzi C., Rato C., Crottini A., Bellati A. Multilocus phylogeography of Italian Moorish geckos adds insights into the evolutionary history of European populations. Zool. Scripta. 2024;53:129–141. doi: 10.1111/zsc.12642. [DOI] [Google Scholar]
- 45.Clark J.D., Harris J.W.K. Fire and its roles in early hominid lifeways. Afr. Archaeol. Rev. 1985;3:3–27. doi: 10.1007/BF01117453. [DOI] [Google Scholar]
- 46.Gowlett J.A.J., Harris J.W.K., Walton D., Wood B.A. Early archaeological sites, hominid remains and traces of fire from Chesowanja, Kenya. Nature. 1981;294:125–129. doi: 10.1038/294125a0. [DOI] [PubMed] [Google Scholar]
- 47.Henry A.G., Brooks A.S., Piperno D.R. Microfossils in calculus demonstrate consumption of plants and cooked foods in Neanderthal diets (Shanidar III, Iraq; Spy I and II, Belgium) Proc. Natl. Acad. Sci. USA. 2011;108:486–491. doi: 10.1073/pnas.1016868108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Roebroeks W., Villa P. On the earliest evidence for habitual use of fire in Europe. Proc. Natl. Acad. Sci. USA. 2011;108:5209–5214. doi: 10.1073/pnas.1018116108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Martini, F., and Abbate, M. (1993). Grotta della Serratura a Marina di Camerota. Culture e ambienti dei complessi oloceneici (Garlatti e Razzai).
- 50.Martini I., Ronchitelli A., Arrighi S., Capecchi G., Ricci S., Scaramucci S., Spagnolo V., Gambassini P., Moroni A. Cave clastic sediments as a tool for refining the study of human occupation of prehistoric sites: insights from the cave site of La Cala (Cilento, southern Italy) J. Quat. Sci. 2018;33:586–596. doi: 10.1002/jqs.3038. [DOI] [Google Scholar]
- 51.Di Pasquale G., Saracino A., Bosso L., Russo D., Moroni A., Bonanomi G., Allevato E. Coastal Pine-Oak Glacial Refugia in the Mediterranean Basin: A Biogeographic Approach Based on Charcoal Analysis and Spatial Modelling. Forests. 2020;11:673. doi: 10.3390/f11060673. [DOI] [Google Scholar]
- 52.Gamble T., Greenbaum E., Jackman T.R., Bauer A.M. Into the light: diurnality has evolved multiple times in geckos: Diurnality Evolved Multiple Times in Geckos. Biol. J. Linn. Soc. Lond. 2015;115:896–910. doi: 10.1111/bij.12536. [DOI] [Google Scholar]
- 53.Vitt L.J., Pianka E.R., Cooper W.E., Jr., Schwenk K. History and the Global Ecology of Squamate Reptiles. Am. Nat. 2003;162:44–60. doi: 10.1086/375172. [DOI] [PubMed] [Google Scholar]
- 54.Benson A.K., Kelly S.A., Legge R., Ma F., Low S.J., Kim J., Zhang M., Oh P.L., Nehrenberg D., Hua K., et al. Individuality in gut microbiota composition is a complex polygenic trait shaped by multiple environmental and host genetic factors. Proc. Natl. Acad. Sci. USA. 2010;107:18933–18938. doi: 10.1073/pnas.1007028107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Hare K.M., Pledger S., Thompson M.B., Miller J.H., Daugherty C.H. Nocturnal lizards from a cool-temperate environment have high metabolic rates at low temperatures. J. Comp. Physiol. B. 2010;180:1173–1181. doi: 10.1007/s00360-010-0489-3. [DOI] [PubMed] [Google Scholar]
- 56.Roth L.S.V., Kelber A. Nocturnal colour vision in geckos. Proc. Biol. Sci. 2004;271:S485–S487. doi: 10.1098/rsbl.2004.0227. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Martín B., Pérez H., Ferrer M. Effects of natural and artificial light on the nocturnal behaviour of the wall gecko. Anim. Biodivers. Conserv. 2018;41:209–215. [Google Scholar]
- 58.Russo D., Cosentino F., Festa F., De Benedetta F., Pejic B., Cerretti P., Ancillotto L. Artificial illumination near rivers may alter bat-insect trophic interactions. Environ. Pollut. 2019;252:1671–1677. doi: 10.1016/j.envpol.2019.06.105. [DOI] [PubMed] [Google Scholar]
- 59.Arlettaz R., Godat S., Meyer H. Competition for food by expanding pipistrelle bat populations ( Pipistrellus pipistrellus) might contribute to the decline of lesser horseshoe bats ( Rhinolophus hipposideros) Biol. Conserv. 2000;93:55–60. doi: 10.1016/S0006-3207(99)00112-3. [DOI] [Google Scholar]
- 60.Roth L.S.V., Lundstrom L., Kelber A., Kroger R.H.H., Unsbo P. The pupils and optical systems of gecko eyes. J. Vis. 2009;9:27. doi: 10.1167/9.3.27. [DOI] [PubMed] [Google Scholar]
- 61.Fulgione D., Trapanese M., Maselli V., Rippa D., Itri F., Avallone B., Van Damme R., Monti D.M., Raia P. Seeing through the skin: dermal light sensitivity provides cryptism in moorish gecko. J. Zool. 2014;294:122–128. doi: 10.1111/jzo.12159. [DOI] [Google Scholar]
- 62.Autumn K., Weinstein R.B., Full R.J. Low Cost of Locomotion Increases Performance at Low Temperature in a Nocturnal Lizard. Physiol. Zool. 1994;67:238–262. doi: 10.1086/physzool.67.1.30163845. [DOI] [Google Scholar]
- 63.Huey R.B., Niewiarowski P.H., Kaufmann J., Herron J.C. Thermal Biology of Nocturnal Ectotherms: Is Sprint Performance of Geckos Maximal at Low Body Temperatures? Physiol. Zool. 1989;62:488–504. doi: 10.1086/physzool.62.2.30156181. [DOI] [Google Scholar]
- 64.Autumn K., Farley C.T., Emshwiller M., Full R.J. Low Cost of Locomotion in the Banded Gecko: A Test of the Nocturnality Hypothesis. Physiol. Zool. 1997;70:660–669. doi: 10.1086/515880. [DOI] [PubMed] [Google Scholar]
- 65.Autumn K., Jindrich D., DeNardo D., Mueller R. Locomotor performance at low temperature and the evolution of nocturnality in geckos. Evolution. 1999;53:580–599. doi: 10.1111/j.1558-5646.1999.tb03793.x. [DOI] [PubMed] [Google Scholar]
- 66.Pfennig D.W., Wund M.A., Snell-Rood E.C., Cruickshank T., Schlichting C.D., Moczek A.P. Phenotypic plasticity’s impacts on diversification and speciation. Trends Ecol. Evol. 2010;25:459–467. doi: 10.1016/j.tree.2010.05.006. [DOI] [PubMed] [Google Scholar]
- 67.Fox R.J., Donelson J.M., Schunter C., Ravasi T., Gaitán-Espitia J.D. Beyond buying time: the role of plasticity in phenotypic adaptation to rapid environmental change. Phil. Trans. R. Soc. B. 2019;374 doi: 10.1098/rstb.2018.0174. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Fulgione D., Maselli V., Rivieccio E., Aceto S., Salvemini M., Buglione M. Evolutionary Plasticity in Insular Lizard, Adapting over Reproduction, Metabolism, and Color Variation. Biology. 2023;12:1478. doi: 10.3390/biology12121478. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Ghalambor C.K., Hoke K.L., Ruell E.W., Fischer E.K., Reznick D.N., Hughes K.A. Non-adaptive plasticity potentiates rapid adaptive evolution of gene expression in nature. Nature. 2015;525:372–375. doi: 10.1038/nature15256. [DOI] [PubMed] [Google Scholar]
- 70.Laland K.N., Uller T., Feldman M.W., Sterelny K., Müller G.B., Moczek A., Jablonka E., Odling-Smee J. The extended evolutionary synthesis: its structure, assumptions and predictions. Proc. Biol. Sci. 2015;282 doi: 10.1098/rspb.2015.1019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Monti D.M., Raia P., Vroonen J., Maselli V., Van Damme R., Fulgione D. Physiological change in an insular lizard population confirms the reversed island syndrome: Physiology in Insular Lizards. Biol. J. Linn. Soc. Lond. 2013;108:144–150. doi: 10.1111/j.1095-8312.2012.02019.x. [DOI] [Google Scholar]
- 72.Pfennig D.W. 1st ed. CRC Press; 2021. Phenotypic Plasticity & Evolution: Causes, Consequences, Controversies. [DOI] [Google Scholar]
- 73.Ghalambor C.K., McKAY J.K., Carroll S.P., Reznick D.N. Adaptive versus non-adaptive phenotypic plasticity and the potential for contemporary adaptation in new environments. Funct. Ecol. 2007;21:394–407. doi: 10.1111/j.1365-2435.2007.01283.x. [DOI] [Google Scholar]
- 74.Wong B.B.M., Candolin U. Behavioral responses to changing environments. Behav. Ecol. 2015;26:665–673. doi: 10.1093/beheco/aru183. [DOI] [Google Scholar]
- 75.Ehrenreich I.M., Pfennig D.W. Genetic assimilation: a review of its potential proximate causes and evolutionary consequences. Ann. Bot. 2016;117:769–779. doi: 10.1093/aob/mcv130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Baxter-Gilbert J., Baider C., Florens F.B.V., Hawlitschek O., Mohan A.V., Mohanty N.P., Wagener C., Webster K.C., Riley J.L. Nocturnal foraging and activity by diurnal lizards: Six species of day geckos (Phelsuma spp.) using the night-light niche. Austral Ecol. 2021;46:501–506. doi: 10.1111/aec.13012. [DOI] [Google Scholar]
- 77.Bouckaert R., Vaughan T.G., Barido-Sottani J., Duchêne S., Fourment M., Gavryushkina A., Heled J., Jones G., Kühnert D., De Maio N., et al. BEAST 2.5: An advanced software platform for Bayesian evolutionary analysis. PLoS Comput. Biol. 2019;15 doi: 10.1371/journal.pcbi.1006650. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Drummond A.J., Suchard M.A., Xie D., Rambaut A. Bayesian Phylogenetics with BEAUti and the BEAST 1.7. Mol. Biol. Evol. 2012;29:1969–1973. doi: 10.1093/molbev/mss075. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Tamura K., Stecher G., Kumar S. MEGA11: Molecular Evolutionary Genetics Analysis Version 11. Mol. Biol. Evol. 2021;38:3022–3027. doi: 10.1093/molbev/msab120. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Rambaut A., Drummond A.J., Xie D., Baele G., Suchard M.A. Posterior Summarization in Bayesian Phylogenetics Using Tracer 1.7. Syst. Biol. 2018;67:901–904. doi: 10.1093/sysbio/syy032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Rambaut A. Institute of Evolutionary Biology, University of Edinburgh; 2010. FigTree v1.3.1. [Google Scholar]
- 82.Bouckaert R.R. DensiTree: making sense of sets of phylogenetic trees. Bioinformatics. 2010;26:1372–1373. doi: 10.1093/bioinformatics/btq110. [DOI] [PubMed] [Google Scholar]
- 83.Avallone B., Tizzano M., Cerciello R., Buglione M., Fulgione D. Gross anatomy and ultrastructure of Moorish Gecko, Tarentola mauritanica skin. Tissue Cell. 2018;51:62–67. doi: 10.1016/j.tice.2018.03.002. [DOI] [PubMed] [Google Scholar]
- 84.Fulgione D., Lega C., Trapanese M., Buglione M. Genetic factors implied in melanin-based coloration of the Italian wall lizard. J. Zool. 2015;296:278–285. doi: 10.1111/jzo.12242. [DOI] [Google Scholar]
- 85.Kalie E., Razi M., Tooze S.A. ULK1 Regulates Melanin Levels in MNT-1 Cells Independently of mTORC1. PLoS One. 2013;8 doi: 10.1371/journal.pone.0075313. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Eo S.H., DeWoody J.A. Evolutionary rates of mitochondrial genomes correspond to diversification rates and to contemporary species richness in birds and reptiles. Proc. Biol. Sci. 2010;277:3587–3592. doi: 10.1098/rspb.2010.0965. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Raia P., Mondanaro A., Melchionna M., Di Febbraro M., Diniz-Filho J.A., Rangel T.F., Holden P.B., Carotenuto F., Edwards N.R., Lima-Ribeiro M.S., et al. Past Extinctions of Homo Species Coincided with Increased Vulnerability to Climatic Change. One Earth. 2020;3:480–490. doi: 10.1016/j.oneear.2020.09.007. [DOI] [Google Scholar]
- 88.Timmermann A., Yun K.-S., Raia P., Ruan J., Mondanaro A., Zeller E., Zollikofer C., Ponce De León M., Lemmon D., Willeit M., Ganopolski A. Climate effects on archaic human habitats and species successions. Nature. 2022;604:495–501. doi: 10.1038/s41586-022-04600-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Haslett J., Parnell A. A Simple Monotone Process with Application to Radiocarbon-Dated Depth Chronologies. J. Roy. Stat. Soc. C Appl. Stat. 2008;57:399–418. doi: 10.1111/j.1467-9876.2008.00623.x. [DOI] [Google Scholar]
- 90.Reimer P.J., Austin W.E.N., Bard E., Bayliss A., Blackwell P.G., Bronk Ramsey C., Butzin M., Cheng H., Edwards R.L., Friedrich M., et al. The IntCal20 Northern Hemisphere Radiocarbon Age Calibration Curve (0–55 cal kBP) Radiocarbon. 2020;62:725–757. doi: 10.1017/RDC.2020.41. [DOI] [Google Scholar]
- 91.R CORE Team A Language and Environment for Statistical Computing R Foundation for Statistical Computing. In (Foundation for Statistical Computing).
- 92.Hijmans R., Phillips S., Leathwick J., Elith J. dismo: methods for species distribution modeling, that is, predicting the environmental similarity of any site to that of the locations of known occurrences of a species. Versiones. 2023;1:3–14. [Google Scholar]
- 93.Phillips S.J., Dudík M. Modeling of species distributions with Maxent: new extensions and a comprehensive evaluation. Ecography. 2008;31:161–175. doi: 10.1111/j.0906-7590.2008.5203.x. [DOI] [Google Scholar]
- 94.Roy-Dufresne E., Saltré F., Cooke B.D., Mellin C., Mutze G., Cox T., Fordham D.A. Modeling the distribution of a wide-ranging invasive species using the sampling efforts of expert and citizen scientists. Ecol. Evol. 2019;9:11053–11063. doi: 10.1002/ece3.5609. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Syfert M.M., Smith M.J., Coomes D.A. Correction: The Effects of Sampling Bias and Model Complexity on the Predictive Performance of MaxEnt Species Distribution Models. PLoS One. 2013;8 doi: 10.1371/annotation/35be5dff-7709-4029-8cfa-f1357e5001f5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Naimi B., Hamm N.A.S., Groen T.A., Skidmore A.K., Toxopeus A.G. Where is positional uncertainty a problem for species distribution modelling? Ecography. 2014;37:191–203. doi: 10.1111/j.1600-0587.2013.00205.x. [DOI] [Google Scholar]
- 97.Zuur A.F., Ieno E.N., Elphick C.S. A protocol for data exploration to avoid common statistical problems: Data exploration. Methods Ecol. Evol. 2010;1:3–14. doi: 10.1111/j.2041-210X.2009.00001.x. [DOI] [Google Scholar]
- 98.Thuiller W., Lafourcade B., Engler R., Araújo M.B. BIOMOD - a platform for ensemble forecasting of species distributions. Ecography. 2009;32:369–373. doi: 10.1111/j.1600-0587.2008.05742.x. [DOI] [Google Scholar]
- 99.Hanley J.A., McNeil B.J. The meaning and use of the area under a receiver operating characteristic (ROC) curve. Radiology. 1982;143:29–36. doi: 10.1148/radiology.143.1.7063747. [DOI] [PubMed] [Google Scholar]
- 100.Allouche O., Tsoar A., Kadmon R. Assessing the accuracy of species distribution models: prevalence, kappa and the true skill statistic (TSS): Assessing the accuracy of distribution models. J. Appl. Ecol. 2006;43:1223–1232. doi: 10.1111/j.1365-2664.2006.01214.x. [DOI] [Google Scholar]
- 101.Marmion M., Parviainen M., Luoto M., Heikkinen R.K., Thuiller W. Evaluation of consensus methods in predictive species distribution modelling. Divers. Distrib. 2009;15:59–69. doi: 10.1111/j.1472-4642.2008.00491.x. [DOI] [Google Scholar]
- 102.Bates D., Mächler M., Bolker B., Walker S. Fitting Linear Mixed-Effects Models Using lme4. J. Stat. Software. 2015;67 doi: 10.18637/jss.v067.i01. [DOI] [Google Scholar]
- 103.Freeman E.A., Moisen G. PresenceAbsence : An R Package for Presence Absence Analysis. J. Stat. Software. 2008;23:1–31. doi: 10.18637/jss.v023.i11. [DOI] [Google Scholar]
- 104.Bellmann, H. (2019). Guida agli insetti d’Europa. Ediz. illustrata (Ricca Editore).
- 105.Hurlbert S.H. Pseudoreplication and the Design of Ecological Field Experiments. Ecol. Monogr. 1984;54:187–211. doi: 10.2307/1942661. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
AMS, Accelerator mass spectrometry; 14C, Radioncarbon dating; OSL, Optical stimulated luminescence.
Data Availability Statement
-
•
The sequencing data are available at National Center for Biotechnology Information (NCBI) as GenBank: MK275668 - MK275678, MK275681, MK275682, MK275684 - MK275686, JQ425045 - JQ425050, JQ425055, JQ425056, JQ425060 and are publicly available as of the date of publication.
-
•
This article does not report original code.
-
•
Any additional information required to reanalyze the data reported in this article is available from the lead contact upon request and are also directly available in the supplementary materials of this article as the distribution of polymorphic gecko populations; phenotypic data on the examined geckos; Homo sapiens past and present habitat suitability map; arthropods captured with and without light.