Category: New Research Page 18 of 66

SICB 2017: Green Anoles, Brown Bodies: Are Brown Lizards “Losers”?

brittneyAnimals frequently compete over resources, and the outcomes of these aggressive interactions depend on a number of factors – one of which is the animals’ previous social experiences. If an animal wins a fight, it may be more likely to win subsequent fights (a “winner effect”), and if it loses, it may be more likely to lose subsequent fights (a “loser effect”).  Garcia et al. (2014, Animal Behavior) previously showed that green anoles exhibit loser effects, but not winner effects. Brittney Ivanov, research technician in Michele Johnson’s lab at Trinity University, wondered whether, since body color in green anoles is associated with social dominance, were color changes in green anoles associated with these loser effects? Could she cause a green anole to be brown if it was forced to lose social contests?

Brittney conducted an experiment using 16 male green anoles. First, in three consecutive days, these focal males interacted with a larger “trainer” male in the trainer male’s home cage for one hour. On the fourth day, the focal males interacted with a size-matched novel male in a cage that was new to both lizards. If the focal males were effectively trained to lose in the first three trials, she predicted that they would lose this fourth trial.

In the series of size-matched trials, 7 of the 16 contests resulted in a clear winner and loser, and 6 of those 7 focal males lost that trial. Further, focal males were less aggressive in the size-matched trial than they were in their previous training trials. These data support the presence of a loser effect in green anoles. Consistent with her previous work, Brittney also found that lizards that were more often green prior to the trials were more likely to win their trials, showing that body color is important in social contests.

brittneycolorgraph2This experiment revealed new findings about loser effects and body color. Focal males who lost their size-matched trial were more likely to be brown in the days after the trials – and not only that, they were more likely to become brown after the trials (so, these weren’t just loser males who had been brown all along).

All together, Brittney’s results show that body color can provide important information about a green anole’s fighting ability or motivation, or its recent social experience, and that dynamic body color influences multiple stages of social interaction in this species.

SICB 2017: Urban Anoles Like It Hot

Postdoctoral scientist, Dr. Shane Campbell-Staton, presents his work on CTmax shifts in Anolis cristatellus at SICB 2017.

Postdoctoral scientist, Dr. Shane Campbell-Staton, presents his work on CTmax shifts in Anolis cristatellus at SICB 2017.

Greetings from New Orleans, where SICB 2017 is well underway! Kicking off the conference was Dr. Shane Campbell-Staton, currently a postdoctoral researcher at the University of Illinois, Urbana-Champaign. Shane presented some work he has been doing with Kristin Winchell, a graduate student in Liam Revell’s lab at the University of Massachusetts, Boston. Kristin’s work focuses on how the crested anole, Anolis cristatellus, adjusts its biology to life in urban areas. In previous work, Kristin documented adaptive shifts in limb and toepad morphology in these anoles in urban areas, a shift she correlated with the broader perches urban anoles use.

In this neat follow-up study, Shane and Kristin have documented how perch temperatures in urban Puerto Rican habitats are higher than in natural environments on the island. In response, urban Anolis cristatellus have a higher heat tolerance. Results from a common garden experiment indicate that the urban shifts in heat tolerance are primarily due to plasticity. At the moment, Shane is performing genomic analyses to search for signatures of selection on heat tolerance.

 

Tails of the City: Caudal Autotomy of Anolis cristatellus in Urban and Natural Environments

Lead author, Kirsten Tyler, reports on her recent Journal of Herpetology paper with K. Winchell and L. Revell:

Urbanization creates drastic changes to habitats leading to differences in microclimate, perch characteristics and distribution, and ecological communities (competitors, prey, and predators) when compared to natural (forest) habitats. Studies have found increased rates of mortality of many urban species due to generalist urban-tolerant predators such as raccoons, feral cats, and domestic animals (Ditchkoff 2006). Anolis lizards are able to voluntarily drop their tails (“autotomize”) when challenged by a predator, enabling their escape in many instances. The maimed lizards are able to regenerate their lost tails, though the replacement tail is a rod of cartilage and not the original bony vertebrae. The regenerated tail portions are often a different color and texture, and the lack of vertebrae / cartilage rod are clearly visible in X-rays.

We hypothesized that autotomy rates would be more similar between urban areas in different municipalities than to natural areas in the same municipality due to similar predator regimes in urban sites across the island. We compared the frequency and pattern (number of caudal vertebrae remaining) of caudal autotomy of A. cristatellus between urban and natural areas in Puerto Rico.

X-rays of our samples with an intact tail (A) and an autotomized tail (B).

X-rays of our samples with an intact tail (A) and an autotomized tail (B).

We sampled A. cristatellus from paired natural and urban sites in four Puerto Rican municipalities: San Juan, Mayagüez, Ponce, and Arecibo. The natural sites were high quality natural forests and the urban sites were high-density residential areas. Urban sites were dominated by asphalt and other impervious surfaces, had sparse tree cover, and a large fraction of potential perches were manmade surfaces such as walls and fences. We scored 967 X-rays from these eight sites for caudal autotomy and counted the number of remaining tail vertebrae. We tested for an effect of urbanization on caudal autotomy by fitting a logistic regression model with municipality (San Juan, Mayagüez, Ponce, Arecibo) and site type (urban, natural), and their interactions, as model factors, and body size as a covariate.

Our data shows that lizards found in urban sites have a larger probability of having autotomized tails.

Our data shows that lizards found in urban sites have a larger probability of having autotomized tails.

Interestingly, we found higher rates of autotomy in all urban populations compared to nearby natural areas. Differences in autotomy might be explained by differences in predator density and efficiency (Bateman 2011). For example, inefficient predators (those that more often than not fail to capture their prey) tend to leave behind more lizards with broken and regenerated tails (Schoener 1979). In addition, a greater abundance of predators could result in more predation attempts. Unfortunately, we did not collect data on predator abundances or community composition, so we cannot distinguish between these (non-mutually exclusive) explanations. Higher rates of autotomy in urban areas could thus reflect any of a variety of factors, including (but not restricted to) inefficient predators in urban areas, a shortage of refuges offering protection from predators, or an increase in predator density.

For lizards with autotomized tails, we found no significant difference in caudal vertebrae number between urban and natural sites.

For lizards with autotomized tails, we found no significant difference in caudal vertebrae number between urban and natural sites.

Lastly, we did not find that lizards with autotomized tails in urban areas had lost more (or less) of their original tail to caudal autotomy. Since regenerated tails cannot be autotomized past the original break point (i.e. cartilage cannot autotomize), this suggests that lizards in urban areas are no more likely to be subject to multiple unsuccessful predation attempts (resulting in caudal autotomy) than lizards in natural forest. Future investigation quantifying predation attempts or predator community composition in urban and forest habitats could help us better understand the source of this intriguing pattern.

 

Read the paper:

R. Kirsten TylerKristin M. Winchell, and Liam J. Revell (2016) Tails of the City: Caudal Autotomy in the Tropical Lizard, Anolis cristatellus, in Urban and Natural Areas of Puerto Rico. Journal of Herpetology: September 2016, Vol. 50, No. 3, pp. 435-441.

 

References:

BATEMAN, P. W., AND P. A. FLEMING. 2011. Frequency of tail loss reflects variation in predation levels, predator efficiency, and the behaviour of three populations of brown anoles. Biological Journal of the Linnean Society 103:648–656.

DITCHKOFF, S. T. 2006. Animal behavior in urban ecosystems: modifica- tions due to human-induced stress. Urban Ecosystems 9:5–12.

SCHOENER, T. W. 1979. Inferring the properties of predation and other injury-producing agents from injury frequencies. Ecology 60:1110–1115.

The Genetic Consequences of Adaptive Dewlap Divergence

Figure 1 from Ng et al. 2016 showing the transect sampling spanning Anolis distichus populations differing in dewlap color (T1-4) as well as control transects (C1-4). Pie charts show dewlap color variation (top row), mitochondrial clade membership (middle row) and nuclear genetic cluster assignments (bottom row).

Figure 1 from Ng et al. 2016 showing the transect sampling spanning Anolis distichus populations differing in dewlap color (T1-4) as well as control transects (C1-4). Pie charts show dewlap color variation (top row), mitochondrial clade membership (middle row) and nuclear genetic cluster assignments (bottom row).

We sure love dewlaps here on Anole Annals! These flashy signals are incredibly diverse in size, color and pattern, and always make for a gorgeous image (e.g. 1, 2). Yet, we still have much to learn about why there is such a diversity of dewlaps and, furthermore, what are the consequences of such diversity? Previous work by Leal and Fleishman (2002, 2004) suggests that some of this dewlap diversity is due to adaptation for more efficient communication in different habitats. In a recent paper, we sought to identify whether the consequence of such adaptive trait divergence was speciation, or whether locally adapted dewlaps are maintained despite gene flow.

Anolis distichus shows remarkable geographic variation in dewlap color that predictably varies with habitat in a manner consistent with adaptation (Ng et al. 2013). This variation in color across Hispaniola gave us a great opportunity to conduct replicated analyses to identify whether adaptive differences in dewlap color consistently leads to the same genetic outcome.

We sampled populations in the Dominican Republic along five transects that transitioned from populations with orange dewlaps to those with cream or yellow dewlaps. For a comparison, we also sampled four ‘control’ transects where all populations shared a similar dewlap color. If dewlap differences are associated with speciation, we expected to see genetic differentiation between populations at either ends of the transect as this would suggest some level of reproductive isolation. Otherwise, transects showing no evidence of genetic structure would suggest that individuals are freely mating regardless of dewlap color.

Looking at the genetic structure of both nuclear and mitochondrial DNA along each transect, we found that geographic variation in dewlap color is associated with both speciation and gene flow. Three transects showed distinct genetic structure consistent with speciation, with one in particular only showing evidence of hybrids at one site which was a mere 0.89-1.55km away from other sampled sites. On the other hand, the other two transects did not look much different to the control transects, suggesting ongoing gene flow regardless of phenotypic differences.

Considering all transects together, I think there are two main take-aways from our results. First, finding evidence of gene flow across a sharp geographic shift in dewlap color must mean that strong selection is maintaining geographic variation in dewlap color; perhaps due to adaptation to different habitat types. Second, it appears that dewlap divergence does not necessarily lead to speciation. More work, however, is needed along these lines to understand whether the dewlaps we are characterizing as different are actually different from an anole’s perspective or in particular light environments (e.g. 1).

Hundreds of Genes Help to Resolve Green Anole Evolutionary History in North America

Anolis carolinensis from North Carolina. Photo from Carolina Nature.

One of the most well-known species of anole lizard is Anolis carolinensis, AKA the green anole, which is the only anole native to the continental United States. As a classic model for ecology and behavior, this lizard was the first species of reptile to have a complete genome sequence. Interestingly, only after it became a genomic model, numerous studies (Tollis et al. 2012, Campbell-Staton et al. 2012, Tollis & Boissinot 2014) sought to understand how genetic variation is structured across the geographic range of A. carolinensis,  and to infer historical migration patterns and demographic events to explain the current distribution of green anoles. However, these studies still left many questions unanswered, mostly due to the fact that they were limited in terms of numbers of genetic markers. Now, we have published a new paper in Ecology and Evolution that used a targeted enrichment method to capture more than 500 sequence markers and provide a clearer picture of A. carolinensis historical biogeography.

What we knew about Anolis carolinensis phylogeography

Collecting green anoles for phylogeographic study has been a real hoot, taking us all over the country. Anolis carolinensis ranges across subtropical North America, and consists of five geographically structured genetic clusters supported by both mitochondrial (mtDNA; see Tollis et al. 2012 and Campbell-Staton et al. 2012) and nuclear (nDNA) markers (see Tollis et al. 2012, Tollis & Boissinot 2014). Three of the clusters are found in Florida : one whose distribution primarily hugs the Northwestern coast of the peninsula, another along the Eastern coast of the peninsula, and a third relegated to South Florida. The continental mainland, while making up most of the area of green anole range, harbors only two clusters: one occupying North Carolina and South Carolina, and another from Georgia, west of the Appalachian Mountains and across the Gulf Coastal Plain into Texas.

One confusing result from earlier studies of A. carolinensis molecular phylogeography was the placement of the most basal lineage in NW Florida (Tollis et al. 2012, Campbell-Staton et al. 2012). This didn’t make sense biogeographically, since it is believed that the species dispersed to the continental mainland from western Cuba (Buth et al. 1980, Glor et al. 2005). However, a subsequent nDNA study (Tollis & Boissinot 2014) produced a multi-locus species tree to show that southern Florida harbors the most ancient lineage of A. carolinensis. This discovery of mito-nuclear discordance provided a more satisfying biogeographical explanation that only needs to invoke overwater dispersal to South Florida from Cuba.

(A) Phylogenetic relationships of the major green anole lineages inferred from the ND2 mtDNA locus. (B) Phylogenetic relationships of the major green anole lineages using multi-locus species tree approach (1 mtDNA and 3 nDNA markers).

Different genetic datasets tell different stories about Anolis carolinensis evolutionary history. (A) Phylogenetic relationships of the major green anole lineages inferred from the ND2 mtDNA locus. (B) Phylogenetic relationships of the major green anole lineages using multi-locus species tree approach (1 mtDNA and 3 nDNA markers). Adapted from Manthey et al. 2016.

From there, things remained unresolved even with nDNA. For instance, while the split between South Florida and the rest of the species received full statistical support in Tollis & Boissinot (2014), the relationships between the other clades were less supported, making it difficult to determine if the A. carolinensis mainland clades arose from separate Floridian sources.

The data used in Manthey et al. 2016

To our knowledge, this is the first Anolis phylogeography study to use targeted enrichment, so I thought I would elaborate on the nature of this kind of dataset. Anchored hybrid enrichment (AHE) relies on probes designed from conserved genomic regions ascertained from a panel of vertebrate genomes – including A. carolinensis – which are flanked by non-conserved regions (the level of conservation in determined by PhastCons scores from the UCSC Genome Browser). DNA samples are pooled, and a set containing thousands of probes is used to enrich libraries that get sequenced on an Illumina platform and assembled into contigs, producing hundreds of homologous loci.

Here’s the breakdown of what we ended up with in the new study: our sample contained 42 individual anoles from 26 localities across eight states, and we were able to obtain 487-512 loci per individual, with an average contig length of 629bp, and an average of 17 SNPs per locus including an average of six parsimony-informative SNPS per locus. Roughly speaking, that’s one parsimony-informative SNP every 100bp for 500 loci, so about 3,000 parsimony-informative SNPS  = not bad! For what it’s worth, the 10 nDNA A. carolinensis markers obtained by more traditional PCR/Sanger sequencing contained about one SNP every 100bp as well (see Tollis et al. 2012 and Tollis & Boissinot 2014). Therefore, AHE produced hundreds more informative loci at a fraction of the cost.

New insights into Anolis carolinensis phylogeography using targeted loci

Using different statistical clustering methods (DAPC and Structure), Manthey et al. supports the same five  genetic clusters as previously described. However, there is now a fully resolved species tree – arrived at using multiple methods. First, the South Florida clade is the most ancient lineage of green anoles, likely splitting off from the rest of the species during the Miocene or Pliocene. However, there is now 100% support for a sister-group relationship between the mainland clades, massively simplifying the story of A. carolinensis. Green anoles likely remained in Florida until the Pleistocene, dispersing northward and onto the mainland where two lineages evolved independently- one along the Atlantic coast in the Carolinas, and another dispersing across the Gulf Coastal Plain.

(A) Map showing geographic localities of 42 green anoles selected for targeted enrichment. (B) Results of species tree analyses. Colored symbols correspond to the five geographic and genetic clusters. Adapted from Manthey et al. (2016).

(A) Map showing geographic localities of 42 green anoles selected for targeted enrichment. (B) Results of species tree analyses. Colored symbols correspond to the five geographic and genetic clusters. Adapted from Manthey et al. (2016).

We also found that despite the best resolution to date for the A. carolinensis species tree, incomplete lineage sorting is rampant across these loci, highlighting the need for these kinds of datasets for phylogeographic studies at this evolutionary distance. For instance, the only clade with any gene trees supporting exclusive ancestry was South Florida: meaning on a given gene tree, pre-defined “clades” are often paraphyletic. The reason the species trees agreed in their topologies is due to fact that they probabilistically invoke the coalescent process, which incorporates incomplete lineage sorting. Previous studies, using ≤10 loci, simply lacked enough statistical power to do this confidently.

More work to be done

As with most scientific endeavors, the new study resolves some outstanding questions but also begs new questions. For instance, although we were able to infer gene flow between the Gulf-Atlantic and NW Florida clades, the degree of allele sharing between populations is still not clear. There seems to be some admixture between the Gulf-Atlantic and Carolinas clades south of the Appalachian Mountains in Georgia, suggesting elevational gradients provide a more effective barrier to gene flow in this species than riverine barriers. Also, the divergence times of the green anole clades are still based only on molecular clock models and could benefit greatly from informative fossils calibrations.

Super-Honest Dewlaps and Trait Scaling Relationships in Semi-Aquatic Anoles

OLYMPUS DIGITAL CAMERA

Anolis aquaticus, the semi-aqautic anole. Photo by Lindsey Swierk

From backyard anole enthusiasts to researchers with decades of experience, dewlaps are a favorite topic of discussion here on Anole Annals. We love documenting the diversity of dewlap colors and patterns (1, 2, 3, 4), judging “best/biggest dewlap” contests (1, 2, 3), and noting dewlap oddities across the genus (1, 2, 3, 4). We’re slowly piecing together an answer to the question of what role dewlaps actually serve in signaling and, in particular, what kind of information they might convey. As you might expect, it’s a pretty complex problem, made even more interesting by the fact that dewlap information content probably reflects the unique pressures placed on individual species.

I’ve recently been working on untangling the mystery of dewlaps in a quirky species of anole, Anolis aquaticus. This water-loving anole is found along streams in pockets of southern Costa Rica and northern Panama, and it has the delightful habit of diving into water when startled. Even among the aquatic-specialized anoles, A. aquaticus is different: it tends to live in ultra-close proximity to water, preferring boulders and crevices directly in the “splash zone” instead of streamside vegetation such as other aquatic species like A. oxylophus. There’s also good reason to think that A. aquaticus has a pretty rich social life – male-female, male-male, and female-female pairs can be found within a few centimeters of each other, and often in dense groups on small rocky islands.

In light of their unusual habitat and living arrangements, we decided to explore how dewlaps correlated with multiple morphological parameters in A. aquaticus. In particular, we decided to use this species to explore a long-standing, but recently debated, paradigm that most sexually selected traits (like dewlaps) scale to body size with positive allometry – or, in other words, that they’re disproportionately large in larger individuals. Last year, we captured male and female (who lack the characteristic reddish-orange dewlap) A. aquaticus and measured multiple sexual and non-sexual traits to test this idea. Our results, available in an accepted article in Integrative Zoology, allowed us to contribute our perspective to the greater understanding of the relationship of sexual selection and allometric scaling patterns. Spoiling part of the punchline, our findings do not support the traditional idea that positively allometry is a hallmark of sexual selection.

The dewlap of Anolis aquaticus. Bar represents 1 cm.

The dewlap of Anolis aquaticus. Bar represents 1 cm.

But, equally as notable, our results also suggest some interesting features of this species, including the information content of its dewlap and how allometric patterns interact to produce sexual dimorphism. We found that dewlaps are “super-honest” signals in A. aquaticus; they could serve to amplify size differences between males signaling at a distance because of their positive allometric scaling with body size. Consequently, our study and a recent study by Driessens et al. 2015 (on A. sagrei), oppose previous ideas that dewlaps approach an asymptote of optimum size to balance the pressures of signaling with predation. Our findings are also novel in that they suggest that dewlap color (redness) may serve to convey information about male weaponry: anoles with redder dewlaps were found to have head shapes that correspond to producing greater bite force.

By comparing allometric relationships between males and females, we can also begin to identify how sexual differences in proportionality link to sexual dimorphism and ecology. For instance, male hind limb length in A. aquaticus is on average larger than that of females, but becomes disproportionately smaller as male body size increases. This opens the door to the idea that, because males are larger than females, limb length sexual dimorphism might be the result of an optimal limb-body size relationship regardless of sex; A. aquaticus of either sex with overlong limbs would probably be at a disadvantage if they needed to flee over narrower surfaces such as branches or vines.

Scaling relationships of snout-vent length and a) mass, b) limb length, and c) head length for male (closed dots, solid line) and female (open dots, dashed line) Anolis aquaticus. Axes are log scaled.

Scaling relationships of snout-vent length and a) mass, b) limb length, and c) head length for male (closed dots, solid line) and female (open dots, dashed line) Anolis aquaticus. Axes are log scaled.

Finally, our results hint at the existence of two life-stage male morphs in A. aquaticus, as already identified in other anole species. Body scaling relationships show that small males have disproportionately small dewlaps, small heads, and large limbs, whereas large males have bigger dewlaps, bigger heads, and smaller limbs than should be expected for their body size. Taken together, these results provide a foundation for future research into “heavyweight” and “lightweight” male morphs and associated behaviors. With their small home ranges and apparently high tolerance for same-sex home range overlap, this could be an especially exciting avenue of exploration in A. aquaticus. In any case, it’s certain that there will be much to learn from this watery, elusive, semi-aquatic anole.

You can read more about this project in our accepted manuscript published online in Integrative Zoology. My co-author on this study, Maria Petelo, is an undergraduate at the University of Hawaiʻi who was supported by OTS/NSF’s Native American and Pacific Islander Research Experience, a program designed to increase the representation of underrepresented groups in the natural sciences.

Dragons in Asian Plantations

Hello Anole enthusiasts. This will be a slightly different post to the usual in that Anolis won’t feature at all! I am one of those weird lizard researchers that is yet to feel the scientific attraction many of my colleagues feel towards Anolis and, as such, have always preferred their much spikier old world agamid counterparts- the garden lizards of the genus Calotes.

I’m currently doing a Ph.D. (supervised by long term AA member Adam C. Algar at The University of Nottingham, UK) that looks at how agamids use the various habitat types within South East Asia (in particular, Peninsular Malaysia), focussing on oil palm plantations, rubber plantations and secondary forests. While some agamids occupy one or more of these habitat types, only a single genus has colonised all of them- Calotes. Calotes versicolor will doubtless be familiar to any herper who has been on vacation to Southern Asia, a large, red-headed take on Anolis sagrei with the same aggressive character and, unfortunately for everything else, the same ability to dominate in invasive situations. While C. versicolor is present in human modified habitats in Malaysia, the forested North of the country is home to a larger, even spikier species: C. emma, which will occupy the forest edges that C. versicolor seems to avoid. We chose C. emma for this reason as the ability to draw comparisons between the semi-natural forest edges and man-made plantations was key to this research.

dscf2553

An adult female Calotes emma in a heavily disturbed oil palm plantation (photo: James J. Hicks)

We measured such Anolis study staples as body temperature (Tb), perch type and behaviour when encountered and all the usual morphology traits. We also characterised habitats structurally using random transects and thermally using ibuttons in copper models and measured thermal performance using a racetrack (not an easy piece of equipment to build in the tropics!) and HD camera.

Differences in behaviour were noticed immediately with C. emma being highly arboreal in rubber plantations (and difficult to noose!), using the smooth trunks to ascend from predators/herpers and rotating around in typical agamid fashion. In oil palm plantations the equivalent ‘trunk’ structures consisted of blocky remnants of fronds that form a hard, smooth surface that lizards tended to avoid (shame these ones don’t have toepads!). Consequently most C. emma were seen on the ground and in and amongst discarded piles of the fronds at most ca. 40 cm from the ground, using these piles as both perches and refugia. In forest edge habitat Calotes used a wide range of perches that incorporated similar heights and structures to those used in the plantation habitat but was much less abundant than in either plantation type. These behavioural differences coincided with differing femur lengths with the arboreal Calotes in rubber having the longest femurs, for their snout-vent length alongside utilising statistically wider perches.

Thermally, plantation habitats were hotter and more variable than forest edges at relevant scales to lizards as has been documented previously at larger scales. Despite this, operative temperatures fell well within C. emma’s thermal tolerances and, currently at least, seem more favourable for this species. Whether climate change will push these habitats closer to C. emma’s critical thermal maximum remain to be seen but applying the ‘standard’ 3°C rise still won’t, theoretically at least, impact their thermal performance.

In short, plantations seem to be great habitat for Calotes emma! They are extremely abundant in these man made habitats, more so than any other vertebrate species despite each plantation type seemingly forcing them to rely on different, single axes of their fundamental niche. The downside is, as always, plantations trade low abundance and high diversity for high abundance and low diversity. While rubber plantations support at least 3 other agamid species in our study area, C. emma was the sole representative in oil palm. Whether this is due to thermal aspects (is it too hot for larger-bodied forest dragons?) or structural aspects (Draco probably can’t glide onto oil palm trunks) or something completely different remain to be tested in a future session of fieldwork. This will focus on assessing the more poorly understood forest specialist agamid species’ structural and thermal niche axes and whether these requirements are met by plantation habitats.

Gonocephalus bellii

My future fieldwork will focus on forest specialist agamids such as Gonocephalus bellii (pictured) and try to explain why they are absent from human dominated habitat types (photo: James J. Hicks)

Age Structure of Invasive Green Anole Populations near Japan

Yasumiba et al 2016 Fig 1

Figure from a new paper by Yasumiba et al. illustrating how LAGs in the cross sections of bones can be used to infer lizard age.

Anolis carolinensis is a disruptive invasive species in the Osagawara Islands near Japan, a UNESCO World Natural Heritage site.  It was first recorded on the island of Chichi-jima in the 1960’s and has since spread to surrounding islands. A recent post on Anole Annals describes efforts to improve the effectiveness of adhesive lizard traps on the islands by using cricket bait.

A new paper by Yasumiba et al. improves our understanding of these invasive A. carolinensis by quantifying their longevity and growth rates using skeletochronology.

Ecology of the San Salvador Bark Anole (Anolis distichus ocior)

 An adult male San Salvador Bark Anole (Anolis distichus ocior) displaying. Photograph by Guillermo G. Zuniga.

An adult male San Salvador Bark Anole (Anolis distichus ocior) displaying.
Photograph by Guillermo G. Zuniga.

Dayton Antley and colleagues from Avila University, the home of AA stalwart Bob Powell, recently published a detailed study of the ecology of the San Salvador bark anole (Anolis distichus ocior) in IRCF Reptiles & Amphibians (an open-access herpetological journal, with this article available here). Anolis d. ocior is one of 17 recognized subspecies of the diverse distichus group, and is found on only San Salvador and Rum Cay (Henderson and Powell 2009).

Antley et al. assessed microhabitat use, activity patterns, and approach distances of A. d. ocior in an approximately 0.3ha study area on the grounds of the Gerace Research Centre, dominated by Tropical Almonds (Terminalia catappa), Papaya (Carica papaya), and Ficus trees.

A Google Map view of the Gerace Research Centre. The study site (24°07'05.2"N 74°27'50.9"W) is outlined in white.

A Google Map view of the Gerace Research Centre. The study site
(24°07’05.2″N 74°27’50.9″W) is outlined in white.

In assessing patterns of microhabitat use throughout the day, Antley et al. conducted surveys every two hours for two days from 0700h (about 40 min after sunrise) to 1900h (about 40 min before sunset). Size class, perch height and diameter, body orientation relative to the ground, and thermal microsite (sun/shade/mixed) were recorded for every observed lizard. In the following two days, approach distances were assessed. This was achieved by a surveyor, wearing neutrally-coloured clothing, approaching an undisturbed anole at a steady pace and recording the distance at which the lizard reacted. Over two additional days, 10-minute focal animal observations were conducted of individual adult lizards (including both males and females) at a distance of 5m. The number of movements (changes in location or orientation), head turns, and head bobs were recorded for all lizards, with dewlap displays and pushups being additional recorded for males.

Lizards were active throughout the day, with activity peaking in the early morning and before midday. This was compared to ambient air temperatures recorded 1m from the ground in a shaded and sheltered location. This result surprised the authors, as a second activity peak in late afternoon/early evening was expected, as has observed in other similar studies of bark anoles (e.g. Hillbrand et al. 2011).

Mean number of lizards active (bars) and mean ambient temperatures (dots) per time period. Temperature data were collected on two consecutive days.

Mean number of lizards active (bars) and mean ambient temperatures
(dots) per time period. Temperature data were collected on
two consecutive days.

Adult males experienced highest levels of arboreality during the middle of the day, while subadult males and adult females (grouped together as they can be hard to distinguish from distance) were highly variable (see figure below). Most lizards of all classes were found in the shade, which the authors attributed as evidence for thermal conformity, and facing downward towards the ground, a common trait in many anoles that is most commonly perceived to increase an individual’s ability to monitor potential predators, competitors, or mates. 43% of lizards, however, were observed facing upwards. The author’s note that this behavior is often interpreted as an individual prepared for escape; however as all lizards were observed from distance and undisturbed, they (admirably) explain that this result is difficult to interpret.

A: Mean perch heights (cm) of adult males (L) and subadult males and females (S); B: mean perch heights of adult males at different times of day; C: mean perch heights of subadult males and females at different times of day.

A: Mean perch heights (cm) of adult males (L) and subadult males and females (S); B: mean perch heights of adult males at different times of day;
C: mean perch heights of subadult males and females at different times of day.

Adult male lizards were bolder than smaller subadult males and females, and retreated at a much closer distance when approached by a surveyor (0.99m +/- 0.07m vs. 1.54m +/- 0.18m). Focal observations revealed no significant differences between adult males vs. subadult males/females in shared behaviors, although there was a high variation in the amount of displaying behavior between adult males. The average time spent conducting dewlap displays was 3%, although one male was recorded investing 47% of his time in a combination of dewlap extensions and pushup displays.

Using all survey data combined, Antley et al. estimate that A. d. ocior in this study plot had a population density of 593 individuals/ha, with lizards observed on all but four of the smallest trees surveyed. Antley et al. note that their density estimate is extremely conservative, and much lower than previously published estimates (e.g. 1.070-5,460 individuals/ha, Schoener and Schoener 1978). The authors suggest that the small size of the study plot may have contributed to the relatively low density.

In all, this is a charming (although admittedly short) study of the natural history of the San Salvador bark anole (A. d. ocior) – a great example of an undergraduate research project that follows through to publication!

References
– Antley, D.L. et al. 2016. Microhabitat, Activity, and Approach Distances of the San Salvador Bark Anole (Anolis distichus ocior). IRCF Reptiles & Amphibians 23(2): 75-81
– Henderson, R.W. and R. Powell. 2009. Natural History of West Indian Reptiles and Amphibians. University of Florida Press, Gainesville, Florida.
– Hillbrand, P.A., A.T. Sloan, and W.K. Hayes. 2011. The terrestrial reptiles of San Salvador Island, Bahamas. Reptiles & Amphibians 18: 154–166.
– Schoener, T.W. and A. Schoener. 1978. Estimating and interpreting body-size growth in some Anolis lizards. Copeia 1978: 390–405.

ESA 2016: Niche Partitioning and Rapid Adaptation of Urban Anoles

Maintaining an already-impressive 2016 conference tour de force which included presentations at both JMIH and Evolution, Kristin Winchell presented a broad summary of her urban anole research in an invite-only Urban Ecology session at ESA 2016.

introslide

This presentation provided a synthesis of two large research projects both independently reviewed on Anole Annals (1,2), and so I will provide only a brief summary here. Kristin began by presenting an over-arching question in modern ecology: how is urbanisation going to affect biodiversity? While many may intuitively think of the process negatively, there is a large (and growing) body of research suggesting that many species are able to behaviourally respond to these novel environments and persist. So what about anoles? Kristin focuses her research on two Puerto Rican species: the crested anole (Anolis cristatellus) and the barred anole (A. stratulus).

stratulusvcristatellus

To do this, Kristin and her team employed multiple methods to explore if a) these two species have differences in their ecology in urban vs. natural areas, b) if differences in ecology are observed, does this lead to differences in morphology, and c) if differences in morphology are observed, is this related to performance? Firstly, niche partitioning between these two species in natural vs. urban areas was investigated (more details here).

novel habitat

This niche partitioning research is new and will be the main body of a manuscript currently in prep so I will keep discussions brief. One species, A. cristatellus, was observed to significantly shift its microhabitat use, which resulted in adaptive shifts in morphology. This research was documented in Winchell et al.’s recent Evolution paper and reviewed previously on AA (1,2,3). Specifically, urban lizards have longer limbs and stickier toepads (higher number of subdigital lamellae) in response to perching on broader, slippier substrates.

phenotypic shifts

This research has now developed on to the next stage of performance-related investigations. Kristin is asking the question of whether these observed morphological shifts lead to better performance (and therefore, presumably, higher fitness). Kristin presented some preliminary results, but keep your eye out for more developments!

performance

Page 18 of 66

Powered by WordPress & Theme by Anders Norén