Review Article
Print
Review Article
Colonization record of the Galápagos’ vertebrate clades: Biogeographical issues plus a conservation insight
expand article infoJason R. Ali, Uwe Fritz
‡ Museum of Zoology, Senckenberg Dresden, Dresden, Germany
Open Access

Abstract

Our focus is the colonization history of the Galápagos’ vertebrate clades: 11 land-bound groups (eight reptiles, three rodents) and 13 taxa of flyers and swimmers (ten winged birds, two pinnipeds, one penguin). Using ‘colonization intervals’ and ‘colonization profiles’, it is clear that the two sets of taxa assembled very differently. The former includes older clades with between one, and potentially eight, predating the emergence of the oldest island (4 Mya). For the origin of some lineages, now-sunken landmasses associated with the Galápagos mantle-plume hotspot must have been involved, but for others it could reflect taxonomic uncertainties. In contrast, the taxa of flyers and swimmers are markedly younger, indicating either higher rates of colonization and extirpation for these sorts of animal, or continued genetic influx from mainland populations, or some combination of both factors. Concerning the first, possible drivers are the environmental stressors associated with the El Niño–La Niña climate system; the recent clades may be vulnerable to extreme events within the oscillation sequence, perhaps on ≥104-year timescales. Therefore, loose temporal thresholds might exist for the archipelago’s vertebrate groups beyond which selection fortifies them from the most challenging of seasonal states. Moreover, in a world of climate uncertainty, the findings appear relevant to conservation initiatives suggesting a focusing on the younger elements within the Galápagos’ biota.

Keywords

Caribbean flamingo, Darwin’s finches, Galápagos cormorant, Galápagos fur seal, Galápagos giant tortoise, Galápagos iguanas, Galápagos mockingbirds, Galápagos penguin, Galápagos seal, lava lizards

Introduction

For the best part of two centuries, the now-celebrated terrestrial and marine faunas associated with the Galápagos Islands have attracted much attention, both from the scientific research community and the wider public (Nicholls 2014). Knowledge of when the ancestors of each of the animal clades colonized the archipelago is important because it often underpins studies evaluating evolutionary process rates within each of the descendent lineages (e.g., Lamichhaney et al. 2015; Hedrick 2019). Such information has emerged through molecular-based estimations of the time of the taxa’s arrivals (e.g., Wright 1983; Wyles and Sarich 1983; Caccone et al. 1999; Zaher et al. 2018) combined with the dating of the volcanic-rock substrates (e.g., Cox and Dalrymple 1966; Bailey 1976; Hall 1983; Geist et al. 1985, 2014; Hickman and Lipps 1985). Interestingly, a number of the land-vertebrate groups, for instance the iguanas, geckos, and lava lizards, have long been thought to predate the islands (e.g., Wright 1983; Wyles and Sarich 1983), the oldest of which, San Cristóbal, arose just 4 Mya (Geist et al. 2014). Here, we examine the issue further by assessing (i) the region’s geographical development, and (ii) the colonization record for the native vertebrate-suite components, i.e., land-bound, flying, and swimming species (two pinnipeds and one penguin). Marine fishes and sea turtles were not considered due to their weak connection to land surfaces. Notably, this review draws upon elements of ‘colonization profile’ analysis (Ali and Hedges 2023), a technique that was recently developed to assess the arrival history of the ancestors of the land-vertebrate lineages on Madagascar Island in the SW Indian Ocean.

Galápagos mantle-plume hotspot: volcanic-island factory

Although the Galápagos Islands are renowned for the role their biota has played in the development and refinement of Natural Selection (e.g., Darwin 1859; Grant and Grant 2008; Hedrick 2019), they are important to the geoscience community with regards to mantle-plume hotspots, which are a fundamental feature of the solid-Earth system (e.g., Morgan 1971; Harpp et al. 2005; Villagómez et al. 2007; Whattam and Stern 2015; Ali and Meiri 2023; Soderman et al. 2023). A basic requirement of practically all geological investigations is a robust chronological framework. Early data for the Galápagos were provided by Cox and Dalrymple (1966), whose two-pronged radiometric- and magnetostratigraphic-dating programme indicated that the various islands were all, apparently, less than 2 Mya (n.b., material from what is now known to be the oldest island, San Cristóbal, was not examined). Later studies (Bailey 1976; Hall 1983; Geist et al. 1985; Hickman and Lipps 1985) pushed back the maximum age to ca. 4 Mya (Fig. 1). The geological work was then extended by Christie et al. (1992) and Sinton et al. (1996) to the submerged volcanic massifs on the eastern Galápagos platform and to the sea-bed to the east on the Carnegie Ridge’s western end (Figs 2 and 3). The dredge-haul samples they recovered included once-subaerial rocks (rounded clasts, weathered materials, etc.) with some yielding ages ranging from 2.8 to 9.1 Mya. Subsequently, Werner et al. (1999, 2003), Mix et al. (2003) and Sinton et al. (2017, 2018) expanded the coverage and showed that the Galápagos plume had been spawning islands in the vicinity of the Cocos-Nazca plate-spreading ridge since at least the early Middle Miocene (Fig. 2).

Figure 1. 

Map showing the ages of the various islands of the Galápagos Archipelago (figures in parentheses; Geist et al. 2014), plus the estimated time intervals when some nearby seamounts were subaerial (number ranges in brackets; Sinton et al. 2017, 2018). Also depicted are dredge-haul locations were once-subaerial rocks have yielded radiometric age dates (Christie et al. 1992; Sinton et al. 1996; yellow circles with red rims). The base image was created using GeoMapApp (Ryan et al. 2009). The ocean bed down to 1000 m water-depth is shaded white.

Figure 2. 

Map showing the key physiographical features of the eastern equatorial Pacific and the adjacent land areas. The base chart was generated using GeoMapApp (Ryan et al. 2009). The yellow-black dashed lines show the edges of the various tectonic plates based on Bird (2003), although the boundary types (spreading ridge, transform or subduction) are not depicted. Note that the origin of the Malpelo Ridge is contested; Werner et al. (2003) were uncertain whereas Marcaillou et al. (2006) thought it to be a former part of the Cocos Ridge that has now been displaced to the south. Plotted isobaths are for 1000 m and 2000 m. Yellow circles indicate locations where dredge-hauls recovered volcanic rocks that sometime after their formation experienced subaerial weathering/erosion such that today some form rounded pebbles or cobbles (Christie et al. 1992; Sinton et al. 1996). Many such symbols have associated numbers, which denotes the radiometric ages of the recovered clasts (Mya). Red triangle denotes Ocean Drilling Program (ODP) Site 1239, where a 6.8–million-year hiatus was identified between Middle Miocene (c. 14.6 Mya) and the Upper Miocene (c. 7.8 Mya) sedimentary packages (Mix et al. 2003). Notably, the lower unit includes sand grains (0.2–2.0 mm diameter), which is suggestive of nearby land, their mafic compositions indicating a basaltic massif.

Figure 3. 

Bathymetric-topographic profiles across key parts of the eastern equatorial Pacific. The data were downloaded from GeoMapApp (Ryan et al. 2009). (A) and (B) are along the main axes of the Galápagos platform-Carnegie Ridge and the western Galápagos platform-Cocos Ridge respectively, with their traces shown in Figure 2. A third section (C) provides insight into the general morphology of the ocean floor of the eastern Pacific; note its slightly lower base-level. This reflects the increased age of the crust, which formed at the Galápagos Spreading Centre (Fig. 2; also see O’Connor et al. 2007) and thermal-cooling subsidence (Parsons and Sclater 1977; Hillier and Watts 2005; Kearey et al. 2009). Depicted on (A) and (B) are the various age-dated dredge-haul samples that experienced subaerial erosion (see Fig. 3) are in their approximate transferred positions (yellow circles). Concerning the ‘plate up-flexing’, see Turcotte et al. (1978) and Kearey et al. (2009).

Appreciation of the morphology of the Galápagos platform-Carnegie Ridge and the Cocos Ridge can be gleaned from a series of bathymetric cross-sections across the various physiographic highs that are shown in Figures S1, S2, and S3. Significantly, there are sizable tracts of sea-floor that are < 2000 m deep, with a number of patches < 1000 m (Fig. 2), that when the effects of lithospheric-cooling subsidence together with geological time are considered (e.g., Parsons and Sclater 1997; Hillier and Watts 2005; Ali and Aitchison 2014), it is certain that some of these could once have been emergent (see Orellana-Rovirosa and Richards 2018; Ali and Meiri 2023).

Methods

Colonization Profile analysis

One of the problems with biogeographical studies of insular systems is the lack of an analytical framework that can be used to quantitively evaluate the colonization history of a biotic assemblage. This shortfall has, however, been recently addressed through the introduction of ‘Colonization Profile’ analysis (Ali and Hedges 2023). The procedure comprises three stages: (i) establishing colonization intervals for each of the clades, (ii) constructing the colonization profile(s), and (iii) generating equation values for use as both a quality-control metric, as well as for simulating and testing hypothetical arrival histories (e.g., temporally stochastic overwater dispersal, focused influx along a temporary land-bridge).

A clade’s ‘colonization interval’ represents the earliest and latest possible instants the lineages’ ancestors could have arrived on the island/archipelago. The initial estimation uses molecular-clock data, the former value being provided by the stem node’s older uncertainty age, while the latter one is given by the crown node’s younger uncertainty age. If the original publications do not include confidence intervals for the relevant nodes, then the preceding and succeeding divergences (i.e., of the parent and child taxa) are used. Subsequently, it may be possible to refine each clade’s interval using palaeontological data, island-emergence information (for volcanic landmasses constructed upon deep ocean floor) and regional or global extinction events. The first might pull back the latest time of arrival when direct proof of presence predates the crown-young age. The second and third may push forward the earliest time of arrival; a clade could not have been present before the landmass’ birth, nor could its descendants have survived the environmental devastation. Where on-island or on-archipelago divergence has not taken place, or relevant data are unavailable, the younger end of the colonization interval is set at 0 Mya.

A ‘colonization profile’ is then constructed in the form of a histogram by ‘stacking’ all of the colonization intervals. Its shape reflects the overall history of arrivals with spikes and troughs marking influx concentrations and reduced arrivals. The plot is referred to as the ‘actual’ data profile.

Stage three involves calculating equation values for use in quality control and model testing. Here all of the colonization intervals are plotted against their colonization interval mid-points (CIMP). This is followed by the stem-old and the crown-young ages each having best-fit lines applied with equation values deduced. Thereafter, a ‘back-modelled’ profile can be generated that uses the equation values linked to the ‘actual’ CIMP age values. Ideally, the ‘actual’ and ‘back-modelled’ profiles show close correspondence. This is assessed through two quantitative metrics. The first is a simple measure of ‘fit’ based on (Ctot−Cnm)/Ctot, where Ctot is the total number of cells associated with the ‘actual’ data and Cnm is the number of non-matching cells. The second makes use of the chi-squared test; for each 1-Myr age-bin there is an observed value related to the simulation, as well as an expected value based on the ‘actual’ and/or the ‘back-modelled’ data. Histograms are used to display the data, and in most cases the enveloping lines for the ‘actual’ and ‘back-modelled’ profiles are overlain for comparison.

Information sources for the colonization intervals for each of the Galápagos clades

Information sources of the node-age determinations for the various clades are presented below (see also Table 1), with ‘stem’ and ‘crown’ abbreviated to ‘st.’ and ‘cr.’ For processing purposes, the various clades are allocated to one of two types: land-bound vertebrates (eight reptiles, three rodents; labelled 1–11); and non-land-bound vertebrates (ten winged birds, including the now-flightless Galápagos cormorant, plus the Galápagos fur seal, Galápagos sea lion, and the Galápagos penguin; labelled A–M). Two other vertebrate groups were omitted due to their limited connections with the Galápagos’ terrestrial environments: marine fishes (the archipelago hosts no native primary freshwater fishes) and the green sea turtle, Chelonia mydas (see Jensen et al. 2019).

Table 1.

Molecular-clock data used to determine the colonization intervals for the Galápagos’ vertebrate clades (see also Figs 4 and 7). The numbering (land-locked groups) and lettering (non-land-locked lineages) schemes for the clades is based on the stem-old age. The age dates are listed to two decimal places so as make the reading of the table easier, although in some cases the actual values may have lower or higher precision. Key: CA stands for common ancestor; CIMP is an abbreviation for ‘colonization interval mid-point’; † indicates an extinct rodent species; * applies to two clades that comprise non-endemic species. Note that the MacLeod et al. (2015) reference is not mentioned in the main text. Other publications that provide context to the compiled dataset include do Prado and Percequillo (2018) and Dowler and Revelez (2021) for the three rodent groups, Lovette et al. (2012) for the mockingbird clade, Wolf et al. (2007) for the sea lion, Bollmer et al. (2006) for the hawk; Sari and Bollmer (2017) provide an excellent review of the bird groups.

Ages (Mya)
Clade code Endemic land-bound vertebrate clades Off-archipelago sister Molecular clock information sources Stem Stem-old Stem-young Crown Crown-old Crown-young CIMP
1 Chelonoidis niger giant tortoise Chelonoidis chilensis Kehlmaier et al. (2023); Poulakakis et al. (2020) reported similar ages 17.01 23.50 9.50 2.19 3.10 1.40 12.45
2 Phyllodactylus geckos’ main clade CA Phyllodactylus johnwrighti and P. reissii Torres-Carvajal et al. (2016) and Koch et al. (2016) 13.80 20.21 7.92 5.17 6.17 4.29 12.25
3 Amblyrhynchus-Conolophus iguanas Cachryx genus Malone et al. (2017); MacLeod et al. (2015) reported similar ages 8.60 12.50 4.50 5.50 8.30 2.50 7.50
4 Pseudalsophis racer snakes Pseudalsophis elegans Zaher et al. (2018) 6.90 10.61 3.20 4.40 5.45 3.35 6.98
5 Megaoryzomys curioi rodent Mindomys hammondi Estimated by Ali and Fritz (2021) using Leite et al. (2014) and Ronez et al. (2021) 4.50 6.00 3.00 0.00 0.00 0.00 3.00
6 Phyllodactylus darwini gecko Phyllodactylus leoni Torres-Carvajal et al. (2016) and Koch et al. (2016) 3.03 5.87 0.22 0.00 0.00 0.00 2.94
7 Nesoryzomys rodents Aegialomys genus Castañeda-Rico et al. (2019) 3.84 4.88 2.91 2.22 3.12 1.32 3.10
8 Microlophus lava lizards’ main clade CA Microlophus occipitalis and Clade #9 Benavides et al. (2009) 3.70 4.65 2.85 1.40 1.94 0.96 2.81
9 Microlophus lava lizards’ eastern clade Microlophus occipitalis Benavides et al. (2009) 2.80 3.78 1.95 0.42 0.20 0.68 2.23
10 Aegialomys galapagoensis rodent Aegialomys xanthaeolus Castañeda-Rico et al. (2019) 1.10 2.11 0.37 0.00 0.00 0.00 1.06
11 Phyllodactylus gilberti gecko Phyllodactylus reissii Torres-Carvajal et al. (2016) and Koch et al. (2016) 0.69 1.56 0.04 0.00 0.00 0.00 0.78
Non-land-bound vertebrate clades
A Phalacrocorax harrisi cormorant CA Phalacrocorax brasilianus and P. auratus Burga et al. (2017) 2.37 6.67 1.30 0.00 0.00 0.00 3.34
B Mimus mockingbirds Mimus gundlachii Arbogast et al. (2006) for stem, Nietlisbach et al. (2013) for crown 3.55 5.50 1.60 0.49 1.39 0.15 2.82
C Zenaida galapagoensis dove CA Zenaida auriculata and Z. graysoni Valente et al. (2015) 3.51 4.65 2.57 0.00 0.00 0.00 2.33
D Darwin’s finches CA Tiaris fulignosus and T. obscurus Funk and Burns (2018) 2.50 3.85 1.70 1.70 2.50 1.55 2.70
E Laterallus spilonota rail Laterallus jamaicensis Chaves et al. (2020) 1.20 1.90 0.50 0.00 0.00 0.00 0.95
F Spheniscus mendiculus penguin Spheniscus humboldti Gavryushkina et al. (2017) 1.06 1.88 0.63 0.00 0.00 0.00 0.94
G Arctocephalus galapagoensis fur seal Arctocephalus australis ‘B’ Yonezawa et al. (2009) 1.30 1.70 0.90 0.00 0.00 0.00 0.85
H Pyrocephalus vermillion flycatchers Pyrocephalus rubinus Carmi et al. (2016) 1.15 1.45 0.85 0.82 1.20 0.55 1.00
I Myiarchus magnirostris flycatcher Myiarchus tyrannulus Sari and Parker (2012) 0.86 1.13 0.58 0.00 0.00 0.00 0.57
J Zalophus wollebaeki sea lion Zalophus californianus Asadobay et al. (2023) 0.65 0.85 0.45 0.00 0.00 0.00 0.43
K Buteo galapagoensis hawk Buteo swainsoni do Amaral et al. (2009) 0.34 0.65 0.12 0.00 0.00 0.00 0.33
L Phoenicopterus ruber flamingo* Not applicable Frias-Soler et al. (2022) 0.16 0.45 0.00 0.00 0.00 0.00 0.23
M Dendroica petechia yellow warbler* Not applicable Browne et al. (2008), see also Chaves et al. (2012) 0.28 0.38 0.18 0.00 0.00 0.00 0.19

Regarding the target study groups, several clades had to be excluded due to them lacking phylogenetic data, for instance the dark-billed cuckoo (Coccyzus melacoryphus), Galápagos martin (Progne modesta), South American hoary bat (Aeorestes villosissimus), and western red bat (Lasiurus blossevillii brachyotis).

The colonization interval literature sources are: 1, Chelonoidis niger tortoises, st. and cr. – Kehlmaier et al. (2023); 2, main clade of the Phyllodactylus geckos, st. and cr. – Koch et al. (2016) and Torres-Carvajal et al. (2016); 3, Amblyrhynchus-Conolophus iguanas, st. and cr. – Malone et al. (2017); 4, Pseudalsophis racer snakes, st. and cr. – Zaher et al. (2018); 5, †Megaoryzomys curioi rodent, st. – estimated by Ali and Fritz (2021) based on Leite et al. (2014) and Ronez et al. (2021); 6, Phyllodactylus darwini gecko, st. – Koch et al. (2016) and Torres-Carvajal et al. (2016); 7, Nesoryzomys rodents, st. and cr. – Castañeda-Rico et al. (2019); 8, main clade of the Microlophus lava lizards, st. and cr. – Benavides et al. (2009); 9, eastern clade of the Microlophus lava lizards, st. and cr. – Benavides et al. (2009); 10, Aegialomys galapagoensis rodent, st. – Castañeda-Rico et al. (2019); 11, Phyllodactylus gilberti gecko, st. – Koch et al. (2016) and Torres-Carvajal et al. (2016); A, Phalacrocorax harrisi cormorant, st. – Burga et al. (2017); B, Mimus mockingbirds, st. – Arbogast et al. (2006), cr. – Nietlisbach et al. (2013); C, Zenaida galapagoensis dove, st. – Valente et al. (2015); D, Darwin’s finches, st. and cr. – Funk and Burns (2018); E, Laterallus spilonota rail, st. – Chaves et al. (2020); F, Spheniscus mendiculus penguin, st. – Gavryushkina et al. (2017); G, Arctocephalus galapagoensis fur seal, st. – Yonezawa et al. (2009); H, Pyrocephalus vermillion flycatchers, st. and cr. – Carmi et al. (2016); I, Myiarchus magnirostris flycatcher, st. – Sari and Parker (2012); J, Zalophus wollebaeki sea lion, st. – Asadobay et al. (2023); K, Buteo galapagoensis hawk, st. – do Amaral et al. (2009); L, Phoenicopterus ruber Caribbean flamingo, st. – Frias-Soler et al. (2022); M, Dendroica petechia yellow warbler, st. – Browne et al. (2008), see also Chaves et al. (2012). As Clades 5, 6, 10, 11, A, C, E, F, G, I, J, K, L, and M do not have crown nodes, the younger bound of each’s colonization interval is set at 0 Mya. Regarding the giant tortoises (Clade #1), we note the recent study of Torres et al. (2024), which used a total evidence matrix (morphology combined with mitochondrial DNA information). However, this resulted in a highly unlikely phylogenetic topology with the African crown group testudinine Stigmochelys pardalis placed basal to all other extant testudinids, i.e., the testudinine and xerobatine tortoises, thereby compromising any age estimates.

Results

Colonization intervals for the land vertebrates

The initial presentation of the colonization intervals is shown in Figure 4. Notably, those for the land-bound vertebrates have wide ranges (Fig. 4A), with four of the eleven groups having CIMP ages greater than that of the oldest island (4 Mya). This contrasts with the flying and swimming vertebrates, which collectively have narrower ranges (Fig. 4B), with all of the CIMP values <4 Mya.

Figure 4. 

Colonization interval data for Galápagos’ vertebrate clades. (A) is for the land-bound lineages, (B) is for the swimming and flying groups that are made up of endemic species. Clade key (A): 1, Chelonoidis niger tortoises; 2, main clade of the Phyllodactylus geckos; 3, Pseudalsophis racer snakes; 4, Amblyrhynchus-Conolophus iguanas; 5, †Megaoryzomys curioi rodent; 6, Phyllodactylus darwini gecko; 7, Nesoryzomys rodents; 8, main clade of Microlophus lava lizards; 9, eastern clade of the Microlophus lava lizards; 10, Aegialomys galapagoensis rodent; 11, Phyllodactylus gilberti gecko. Clade key (B): A, Phalacrocorax harrisi cormorant; B, Mimus mockingbirds; C, Zenaida galapagoensis dove; D, Darwin’s finches; E, Laterallus spilonota rail; F, Spheniscus mendiculus penguin; G, Arctocephalus galapagoensis fur seal; H, Pyrocephalus vermillion flycatchers; I, Myiarchus magnirostris flycatcher; J, Zalophus wollebaeki sea lion; K, Buteo galapagoensis hawk; L, Phoenicopterus ruber Caribbean flamingo; M, Dendroica petechia yellow warbler. In (B), the asterisks highlight the three swimming clades (F, G, and J).

Generation of equation values for the land-bound vertebrates

Figure 5A shows the various colonization intervals in Figure 4A with the colonization interval mid-point (CIMP) plotted against the interval range. The accompanying graph (Fig. 5B) presents all of the stem-old and the crown-young age dates, with best-fit lines calculated, respectively (1.721 * CIMP) + 0.114, where r2 = 0.983 and (0.278 * CIMP) – 0.114, where r2 = 0.607 (n = 11).

Figure 5. 

(A) Individual colonization intervals plotted against their mid-point values. (B) Best-fit lines through the stem-old ages and the crown-young ages are used to generate simulated colonization profiles (Fig. 4). CIMP, colonization interval mid-point; cr., crown; st. stem.

Colonization profiles ‘actual’ and simulated

The colonization profile for the ‘actual’ data is shown in Figure 6A. In comparison, the ‘back-modelled’ data profile, Figure 6B, has a ‘fit’ of 0.889 with P = 1. The plot reiterates the strong ‘older than the islands’ issue. Indeed, if all of the ancestral colonizers had arrived at 4 Mya, the colonization profile would show an obvious mismatch (Fig. 6C), with ‘fit’ values around 0.38 and p ≤ 0.001.

Figure 6. 

Galápagos vertebrate colonization profiles. (A) is the ‘actual’ data (grey shaded region and red line) for the land vertebrates, while (B) is the ‘back-modelled’ data (grey shaded region and blue line, with the ‘actual’ data line [red] overlain). (C) is a simulation based upon all of the colonizations taking since the oldest extant island (San Cristóbal) formed, with an arrival every c. 381 kya starting at 4 Mya and ending at c. 190 kya, a scenario that is termed ‘constant rate’. (D) is the profile for the non-land vertebrate clades, with the ‘actual’ and ‘back-modelled’ land-vertebrate lines added. Where appropriate, ‘fit’ and p values are shown on the respective plots.

Colonization profile for the flying and swimming vertebrates

The colonization profile for the non-land-bound vertebrates, both endemic and non-endemic, is shown in Figure 6D (respectively 10 flying and 3 swimming clades). Its shape is conspicuously different to that of land-bound groups and indicates that it comprises more recent arrivals.

Discussion

Acknowledgement of the limitations of the dataset

The assembled dataset makes use of divergence dates from more than twenty different studies that were published over a 15-year interval (Browne et al 2008; Kehlmaier et al. 2023). Hence, they differ in terms of the quality of the data as well as the level of the analyses. However, a general bias is unlikely to have impacted them all, and we are confident that the features we outline below, as well as the conclusions that follow, are likely to be valid. At the very least, perceived weakness may spur discussion and/or further work.

Explanations for the ‘old’ colonization ages within the land-bound vertebrates

Early studies reporting ‘older than the islands colonizations’ included Wright (1983) for the geckos and lava lizards, Lopez et al. (1992) for the lava lizards, and Rassmann (1997) for the marine and land iguanas (Amblyrhynchus, Conolophus). The data compiled for this study indicate that the condition applies to a least one lineage and up to seven others (Fig. 7A,B). There are a number of reasons why this might be the case. Firstly, the ancestral colonizers arrived on a now-drowned landmass; as explained above, the Galápagos hotspot has been generating islands in the eastern equatorial Pacific since at least the Middle Miocene, and likely without a break. This has to be the case with the oldest Phyllodactylus gecko clade (#2), plus it probably also applies to the Amblyrhynchus-Conolophus iguanas (#3) and the Pseudalsophis racer snakes (#4), and possibly the Chelonoidis niger tortoises (#1), the †Megaoryzomys curioi rodent (#5), the Phyllodactylus darwini gecko (#6), Nesoryzomys rodents (#7), and the main clade of the Microlophus lava lizards (#8). However, for some of these groups, the supposed sister species on the mainland could have been incorrectly identified and closer relatives actually exist, or off-archipelago extinction has removed intermediate lineages, thus increasing the ages of the stem nodes as an artefact. Notably, the ‘lost taxa’ issue was pondered in early molecular studies of the Galápagos tortoises (e.g., Caccone et al. 1999, 2002). Recent dating of the split between Chelonoidis niger and its genetically closest extant relative C. chilensis, which is found in parts of Argentina, Bolivia, and Paraguay, yields a stem-node uncertainty range of 23.5–9.5 Mya (Kehlmaier et al. 2023; these numbers are consistent with the independent studies of Sánchez et al. 2017 and Poulakakis et al. 2020). Significantly, though, Chelonoidis-genus subfossils have been recovered from numerous Pliocene and younger sedimentary formations in western South America, Central America, and the Caribbean, with uppermost Pleistocene ‘giants’ found in coastal areas of mainland Ecuador (Kehlmaier et al. 2017 and references therein; Ali and Fritz 2021).

Figure 7. 

Stem (red) and crown (blue) intervals for the various vertebrate clades and how they relate to the extant islands. The land-bound groups are shown in (A) and (B), while the non-land-bound lineages are in (C) and (D). The data are plotted using two different y-axis measures, the colonization interval mid-points (A) and (C) and the crown-old ages (B) and (D). For those clades where on-archipelago diversification has not taken place, or no relevant age data are available, the crown-young ages are set at 0 Mya. In such cases, the associated horizontal bars have been given a width of –0.10 to 0.10 Mya to make them visible.

Comparing the land-vertebrate colonization record with that of the flying and swimming vertebrates

The colonization profile for the flying and swimming vertebrate clades is unlike that for the land-bound groups (Fig. 7C,D). Notably, ten of the clades postdate the oldest island. The conspicuous lack of old components suggests that those taxa that can cross easily to the archipelago have either experienced higher rates of turnover, or that following the initial colonization a steady trickle of mainland migrants has slowed genetic differentiation, thus lowering the ages of the stems and, where relevant, the crowns (e.g., Leaché et al. 2014; Tiley et al. 2023), or that both processes have operated.

Concerning the turnover explanation, the Galápagos Islands are located in the eastern equatorial Pacific and the biota is subject to significant environmental stresses imposed by the El Niño–La Niña climate cycles (Wang and Fiedler 2006; Chen et al. 2019). Notably, the phenomenon has deleteriously impacted a wide variety of taxa (Table 1 of Dueñas et al. 2021). The hot, wet El Niño state negatively affects the marine feeders, decimating their food chains, while the cool, dry La Niña phase poses a greater challenge to the land-based foragers (Hedrick 2019). Moreover, the system, which has a somewhat erratic periodicity of 3–7 years (Fig. 8), appears to have operated since at least the Late Miocene, although with intervals of reduced intensity as compared to the modern configuration (e.g., Galeotti et al. 2010; Weiss et al. 2017; Drury et al. 2018). Thus, a potential extirpation mechanism is in place that might regularly pressure any recent arrivals, the older clades presumably having evolved such that sufficient numbers of their members can ride out the extremes of the climatic phases. A suggested example of a taxon being affected in such a manner is the Galápagos penguin Spheniscus mendiculus, which has a colonization interval of 1.88 to 0 Mya. Notably, its small, geographically-confined population shows an almost complete lack of genetic diversity, which is thought to result from it having passed through multiple bottleneck events (Arauco-Shapiro et al. 2020). This contrasts with the marine iguana Amblyrhynchus cristatus whose colonization interval is 12.5 to 2.5 Mya (Table 1); Steinfartz et al. (2007) showed it experienced little genetic loss following the severe El Niño event of 1997–1998 (Fig. 8), despite the depletions of populations on various islands reaching up to 90%.

Figure 8. 

Temperature anomalies associated with the El Niño-La Niña climate system. The data are from NOAA’s Climate Prediction Center (https://origin.cpc.ncep.noaa.gov/products/analysis_monitoring/ensostuff/ONI_v5.php, accessed April 2024) and are presented as 3-month means centred on January, April, July, and October. El Niño phases are associated with increased precipitation, whereas La Niña intervals are comparatively dry.

The Galapagos’ extant biotic components are potentially threatened by habitat degradation, food-supply depletion, invasive species and climate change. Concerning the last, awareness about the vulnerability of certain taxa has emerged following studies of the El Niño–La Niña system’s impact on low-population species (e.g., Grant et al. 2000; Vargas et al. 2007, 2008; Tindle et al. 2013; Karnauskas et al. 2015; Páez-Rosas et al. 2021). As a consequence, the superposed threat of anthropogenic climate change has led some to outline related conservation strategies (e.g., Dueñas et al. 2021). For instance, with regards to the Galápagos penguin Karnauskas et al. (2015: 6436–6347) discussed re-population initiatives and the setting up of protected marine areas for the animals and their food supply. Concerning the Caribbean flamingo (a non-endemic species that is sometimes referred to as Phoenicopterus ruber glyphorhynchus), Tindle et al. (2016) and Vargas et al. (2021: 263) suggested a variety of measures including behaviour monitoring, habitat-safeguarding, and minimizing human interferences. Considering the findings reported above, perhaps clade-age should be another factor that informs preservation-related strategies, with a specific focus on the groups that on a geological time scale have arrived more recently.

Summary and conclusions

‘Colonization intervals’ have been established for 24 of the Galápagos Archipelago’s land-breeding vertebrate lineages for which reliable molecular-clock data exist.

Using ‘colonization profile’ analysis, we have shown that the land-bound vertebrate clades collectively (eight reptile and three rodent groups) have a rather different arrival record to that of the non-land-bound vertebrates (ten winged birds, two pinnipeds, and one penguin). The former includes at least one and possibly eight lineages that predate the emergence of the oldest island (San Cristóbal, 4 Mya); the flying and swimming groups are much younger, with the arrivals of at least ten postdating this key geophysical event.

The inferred ages of clades with ‘old’ colonization intervals may reflect reality, with their landings on now-submerged islands on the Galápagos Platform–Carnegie Ridge or the Cocos Ridge, but for some it may be an effect of phylogenetic uncertainties and/or taxonomic ambiguities. The younger arrival ages of the flying and swimming clades might be a consequence of either high rates of migration combined with high rates of extirpation for these types of taxa, or that speciation (anagenetic and cladogenetic) is stymied by extended genetic contact with the source population, or even a combination of both. With the former, the El Niño–La Niña climate system likely played a key role. If so, it may be argued that those groups that have been present for ≥106 years are less susceptible to such forcings due to ‘selection hardening’.

Finally, our results indicate that conservation efforts should also focus on the archipelago’s recently-arrived vertebrate groups. If the turnover idea is correct, they may be more likely to succumb to natural-climate variations, especially those that are operating in tandem with effects associated with the human-modified atmosphere.

Acknowledgments

The present study benefited much from formal critiques by Oliver Hawlitschek and an anonymous reviewer. Ralf Britz is thanked for his editorial guidance and further comments.

References

  • Ali JR, Aitchison JC (2014) Exploring the combined role of eustasy and oceanic island subsidence in shaping biodiversity on the Galápagos. Journal of Biogeography 41: 1227–1241. https://doi.org/10.1111/jbi.12313
  • Arauco-Shapiro G, Schumacher KI, Boersma D, Bouzat JL (2020) The role of demographic history and selection in shaping genetic diversity of the Galápagos penguin (Spheniscus mendiculus). PLoS ONE 15: e0226439. https://doi.org/10.1371/journal.pone.0226439
  • Asadobay P, Urquía DO, Künzel S, Espinoza-Ulloa SA, Vences M, Páez-Rosas D (2023) Time-calibrated phylogeny and full mitogenome sequence of the Galapagos sea lion (Zalophus wollebaeki) from scat DNA. PeerJ 11: e16047. https://doi.org/10.7717/peerj.16047
  • Benavides E, Baum R, Snell HM, Snell HL, Sites JW (2009) Island biogeography of Galápagos lava lizards (Tropiduridae: Microlophus): Species diversity and colonization of the archipelago. Evolution 63: 1606–1626. https://doi.org/10.1111/j.1558-5646.2009.00617.x
  • Bollmer JL, Kimball RT, Whiteman NK, Sarasola JH, Parker PG (2006) Phylogeography of the Galápagos hawk (Buteo galapagoensis): A recent arrival to the Galápagos Islands. Molecular Phylogenetics and Evolution 39: 237–247. https://doi.org/10.1016/j.ympev.2005.11.014
  • Burga A, Wang W, Ben-David E, Wolf PC, Ramey AM, Verdugo C, Lyons K, Parker PG, Kruglyak L (2017) A genetic signature of the evolution of loss of flight in the Galapagos cormorant. Science 356: eaal3345. https://doi.org/10.1126/science.aal3345
  • Caccone A, Gibbs JP, Ketmaier V, Suatoni E, Powell JR (1999) Origin and evolutionary relationships of giant Galápagos tortoises. Proceedings of the National Academy of Sciences of the United States of America 96: 13223–13228. https://doi.org/10.1073/pnas.96.23.13223
  • Carmi O, Witt CC, Jaramillo A, Dumbacher JP (2016) Phylogeography of the vermilion flycatcher species complex: Multiple speciation events, shifts in migratory behavior, and an apparent extinction of a Galápagos-endemic bird species. Molecular Phylogenetics and Evolution 102: 152–173. https://doi.org/10.1016/j.ympev.2016.05.029
  • Castañeda-Rico S, Johnson SA, Clement SA, Dowler RC, Maldonado JE, Edwards CW (2019) Insights into the evolutionary and demographic history of the extant endemic rodents of the Galápagos Islands. Therya 10: 213–228. https://doi.org/10.12933/therya-19-873
  • Chaves JA, Martinez-Torres PJ, Depino EA, Espinoza-Ulloa S, García-Loor J, Beichman AC, Stervander M (2020) Evolutionary history of the Galápagos rail revealed by ancient mitogenomes and modern samples. Diversity 12: 425. https://doi.org/10.3390/d12110425
  • Chaves JA, Parker PG, Smith TB (2012) Origin and population history of a recent colonizer, the yellow warbler in Galápagos and Cocos Islands. Journal of Evolutionary Biology 25: 509–521. https://doi.org/10.1111/j.1420-9101.2011.02447.x
  • Chen Y, Lu Y, Zhao P, Xu F, Huang X, Huang X (2019) Seasonal and interannual variations of sea temperature influenced by Galápagos Islands in eastern tropical Pacific Ocean. Journal of Geophysical Research: Oceans 124: 3007–3020. https://doi.org/10.1029/2018JC014523
  • Christie DM, Duncan RA, McBirney AR, Richards MA, White WM, Harpp KS, Fox CG (1992) Drowned islands downstream from the Galápagos hotspot imply extended speciation times. Nature 355: 246–248. https://doi.org/10.1038/355246a0
  • Cox A, Dalrymple GB (1966) Palaeomagnetism and potassium-argon ages of some volcanic rocks from the Galapagos Islands. Nature 209: 776–777. https://doi.org/10.1038/209776a0
  • Darwin CR (1859) On the Origin of Species by Means of Natural Selection, or, the Preservation of Favoured Races in the Struggle for Life. John Murray, London, ix, [1], 502 pp.
  • do Amaral FR, Sheldon FH, Gamauf A, Haring E, Riesing M, Silveira LF, Wajntal A (2009) Patterns and processes of diversification in a widespread and ecologically diverse avian group, the buteonine hawks (Aves, Accipitridae). Molecular Phylogenetics and Evolution 53: 703–715. https://doi.org/10.1016/j.ympev.2009.07.020
  • do Prado JR, Percequillo AR (2018) Systematic studies of the genus Aegialomys Weksler et al., 2006 (Rodentia: Cricetidae: Sigmodontinae): Geographic variation, species delimitation, and biogeography. Journal of Mammalian Evolution 25: 71–118. https://doi.org/10.1007/s10914-016-9360-y
  • Dowler RC, Revelez MA (2021) Chromosomal relationships among the native rodents (Cricetidae: Oryzomyini) of the Galápagos Islands, Ecuador. Therya 12: 317–329. https://doi.org/10.12933/therya-21-1126
  • Drury AJ, Lee GP, Gray WR, Lyl M, Westerhold T, Shevenell AE, John CM (2018) Deciphering the state of the Late Miocene to Early Pliocene equatorial Pacific. Paleoceanography and Paleoclimatology 33: 246–263. https://doi.org/10.1002/2017PA003245
  • Frias-Soler RC, Bauer A, Grohme MA, Espinosa-López G, Gutiérrez-Costa M, Llanes-Quevedo A, Van Slobbe F, Frohme M, Wink M (2022) Phylogeny of the order Phoenicopteriformes and population genetics of the Caribbean flamingo (Phoenicopterus ruber: Aves). Zoological Journal of the Linnean Society 196: 1485–1504. https://doi.org/10.1093/zoolinnean/zlac040
  • Galeotti S, von der Heydt AS, Huber M, Bice D, Dijkstra H, Jilbert T, Lanci L, Reichart GJ (2010) Evidence for active El Niño Southern Oscillation variability in the Late Miocene greenhouse climate. Geology 38: 419–422. https://doi.org/10.1130/G30629.1
  • Gavryushkina A, Heath TA, Ksepka DT, Stadler T, Welch D, Drummond AJ (2017) Bayesian total-evidence dating reveals the recent crown radiation of penguins. Systematic Biology 66: 57–73. https://doi.org/10.1093/sysbio/syw060
  • Geist D, Snell HL, Snell HM, Goddard C, Kurz M (2014) Paleogeography of the Galápagos Islands and biogeographical implications. In: Harpp K, Mittelstaedt E, d’Ozouville N, Graham D (Eds) The Galápagos: A Natural Laboratory for the Earth Sciences. American Geophysical Union (American Geophysical Union Monograph 204), Washington, DC, 145–166.
  • Grant PR, Grant BR (2008) How and Why Species Multiply: The Radiation of Darwin’s Finches. Princeton University Press, Princeton, NJ, 272 pp.
  • Grant PR, Grant BR, Keller LF, Petren K (2000) Effects of El Niño events on Darwin’s finch productivity. Ecology 81: 2442–2457. https://doi.org/10.2307/177466
  • Harpp KS, Wanless VD, Otto H, Hoernle K, Werner R (2005) The Cocos and Carnegie aseismic ridges: A trace element record of long-term plume-spreading center interaction. Journal of Petrology 46: 109–133. https://doi.org/10.1093/petrology/egh064
  • Jensen MP, FitzSimmons NN, Bourjea J, Hamabata T, Reece J, Dutton PH (2019) The evolutionary history and global phylogeography of the green turtle (Chelonia mydas). Journal of Biogeography 46: 860–870. https://doi.org/10.1111/jbi.13483
  • Karnauskas KB, Jenouvrier S, Brown CW, Murtugudde R (2015) Strong sea surface cooling in the eastern equatorial Pacific and implications for Galápagos penguin conservation. Geophysics Research Letters 42: 6432–6437. https://doi.org/10.1002/2015GL064456
  • Kearey P, Klepeis KA, Vine FJ (2009) Global Tectonics (3rd ed.). Wiley-Blackwell, Chichester, xii, 496 pp.
  • Kehlmaier C, Barlow A, Hastings AK, Vamberger M, Paijmans JLA, Steadman DW, Albury NA, Franz R, Hofreiter M, Fritz U (2017) Tropical ancient DNA reveals relationships of the extinct Bahamian giant tortoise Chelonoidis alburyorum. Proceedings of the Royal Society 284B: 20162235. https://doi.org/10.1098/rspb.2016.2235
  • Kehlmaier C, Graciá E, Ali JR, Campbell PD, Chapman SD, Deepak V, Ihlow F, Jalil NE, Pierre-Huyet L, Samonds KE, Vences M, Fritz U (2023) Ancient DNA elucidates the lost world of western Indian Ocean giant tortoises and reveals a new extinct species from Madagascar. Science Advances 9: eabq2574. https://doi.org/10.1126/sciadv.abq257
  • Koch C, Flecks M, Venegas PJ, Bialke P, Valverde S, Rödder D (2016) Applying n-dimensional hypervolumes for species delimitation: Unexpected molecular, morphological, and ecological diversity in the leaf-toed gecko Phyllodactylus reissii Peters, 1862 (Squamata: Phyllodactylidae) from northern Peru. Zootaxa 4161: 41–80. http://doi.org/10.11646/zootaxa.4161.1.2
  • Lamichhaney S, Berglund J, Almén MS, Maqbool K, Grabherr M, Alvaro MB, Promerova M, Rubin CJ, Wang C, Zamani N, Grant BR, Grant PR, Webster MT, Andersson L (2015) Evolution of Darwin’s finches and their beaks revealed by genome sequencing. Nature 518: 371–375. https://doi.org/10.1038/nature14181
  • Leaché AD, Harris RB, Rannala B, Yang Z (2014) The influence of gene flow on species tree estimation: A simulation study. Systematic Biology 63: 17–30. https://doi.org/10.1093/sysbio/syt049
  • Leite RN, Kolokotronis SO, Almeida FC, Werneck FP, Rogers DS, Weksler M (2014) In the wake of invasion: Tracing the historical biogeography of the South American cricetid radiation (Rodentia, Sigmodontinae). PLoS ONE 9: e100687. https://doi.org/10.1371/journal.pone.0100687
  • Lopez TJ, Hauselman ED, Maxson LR, Wright JW (1992) Preliminary analysis of phylogenetic relationships among Galápagos Island lizards of the genus Tropidurus. Amphibia-Reptilia 13: 327–339. https://doi.org/10.1163/156853892X00030
  • Lovette IJ, Arbogast BS, Curry RL, Zink RM, Botero CA, Sullivan JP, Talaba AL, Harris RB, Rubenstein DR, Ricklefs RE, Bermingham E (2012) Phylogenetic relationships of the mockingbirds and thrashers (Aves: Mimidae). Molecular Phylogenetics and Evolution 63: 219–229. https://doi.org/10.1016/j.ympev.2011.07.009
  • MacLeod A, Rodríguez A, Vences M, Orozco-terWengel P, García C, Trillmich F, Gentile G, Caccone A, Quezada G, Steinfartz S (2015) Hybridization masks speciation in the evolutionary history of the Galápagos marine iguana. Proceedings of the Royal Society B 282: 20150425. https://doi.org/10.1098/rspb.2015.0425
  • Malone CL, Reynoso VH, Buckley L (2017) Never judge an iguana by its spines: Systematics of the Yucatan spiny tailed iguana, Ctenosaura defensor (Cope, 1866). Molecular Phylogenetics and Evolution 115: 27–39. https://doi.org/10.1016/j.ympev.2017.07.010
  • Marcaillou B, Charvis P, Collot JY (2006) Structure of the Malpelo Ridge (Colombia) from seismic and gravity modelling. Marine Geophysical Researches 27: 289–300. https://doi.org/10.1007/s11001-006-9009-y
  • Mix AC, Tiedemann R, Blum P, Abrantes FF, Benway H, Cacho-Lascorz I, Chen MT, Delaney ML, Flores JA, Giosan L, Holbourn AE, Irino T, Iwai M, Joseph LH, Kleiven HF, Lamy F, Lund SP, Martinez P, McManus JF, Ninnemann US, Pisias NG, Robinson RS, Stoner JS, Sturm A, Wara MW, Wei W (2003) . Proceedings of the ODP, Initial Reports, Volume 202. College Station, TX (Ocean Drilling Program), 93 pp. http://www-odp.tamu.edu/publications/202_IR/202TOC.htm
  • Nicholls H (2014) The Galápagos: A Natural History. Basic Books, New York, NY, xviii, 195 pp., 8 plates.
  • Nietlisbach P, Wandeler P, Parker PG, Grant PR, Grant BR, Keller LF, Hoeck PEA (2013) Hybrid ancestry of an island subspecies of Galápagos mockingbird explains discordant gene trees. Molecular Phylogenetics and Evolution 69: 581–592. https://doi.org/10.1016/j.ympev.2013.07.020
  • O’Connor JM, Stoffers P, Wijbrans JR, Worthington TJ (2007) Migration of widespread long-lived volcanism across the Galápagos volcanic province: Evidence for a broad hotspot melting anomaly? Earth and Planetary Science Letters 263: 339–354. https://doi.org/10.1016/j.epsl.2007.09.007
  • Orellana-Rovirosa F, Richards M (2018) Emergence/subsidence histories along the Carnegie and Cocos Ridges and their bearing upon biological speciation in the Galápagos. Geochemistry Geophysics Geosystems 19: 4099–4129. https://doi.org/10.1029/2018GC007608
  • Páez-Rosas D, Torres J, Espinoza E, Marchetti A, Seim H, RiofríoLazo M (2021) Declines and recovery in endangered Galapagos pinnipeds during the El Niño event. Scientific Reports 11: 8785. https://doi.org/10.1038/s41598-021-88350-0
  • Parent CE, Caccone A, Petren K (2008) Colonization and diversification of Galápagos terrestrial fauna: A phylogenetic and biogeographical synthesis. Philosophical Transactions of the Royal Society 363B: 3347–3361. https://doi.org/10.1098/rstb.2008.0118
  • Parsons B, Sclater JG (1977) An analysis of the variation of ocean floor bathymetry and heat flow with age. Journal of Geophysical Research 82B: 803–827. https://doi.org/10.1029/JB082i005p00803
  • Poulakakis N, Miller JM, Jensen EL, Beheregaray LB, Russello MA, Glaberman S, Boore J, Caccone A (2020) Colonization history of Galapagos giant tortoises: Insights from mitogenomes support the progression rule. Journal of Zoological Systematics and Evolutionary Research 58: 1262–1275. https://doi.org/10.1111/jzs.12387
  • Rassmann K (1997) Evolutionary age of the Galápagos iguanas predates the age of the present Galápagos islands. Molecular Phylogenetics and Evolution 7: 158–172. https://doi.org/10.1006/mpev.1996.0386
  • Ronez C, Brito J, Hutterer R, Martin RA, Pardiñas UFJ (2021) Tribal allocation and biogeographical significance of one of the largest sigmodontine rodents, the extinct Galápagos Megaoryzomys (Cricetidae). Historical Biology 33: 1920–1932. https://doi.org/10.1080/08912963.2020.1752202
  • Ryan WBF, Carbotte SM, Coplan JO, O’Hara S, Melkonian A, Arko R, Weissel RA, Ferrini AV, Goodwillie A, Nitsche F, Bonczkowski J, Zemsky R (2009) Global multi-resolution topography synthesis. Geochemistry Geophysics Geosystems 10: Q03014. https://doi.org/10.1029/2008GC002332
  • Sánchez J, Holley JA, Poljak S, Bolzán AD, Bravi CM (2017) Phylogenetic and divergence time analysis of the Chelonoidis chilensis complex (Testudines: Testudinidae). Zootaxa 4320: 487–504. https://doi.org/10.11646/zootaxa.4320.3.5
  • Sari EHR, Bollmer JL (2017) Colonization of Galápagos birds: Identifying the closest relative and estimating colonization. In: Parker GP (Ed.) Disease Ecology: Galapagos Birds and their Parasites. Springer, Cham, 15–43. https://doi.org/10.1007/978-3-319-65909-1_2
  • Sari EHR, Parker PG (2012) Understanding the colonization history of the Galápagos flycatcher (Myiarchus magnirostris). Molecular Phylogenetics and Evolution 63: 244–254. https://doi.org/10.1016/j.ympev.2011.10.023
  • Sinton CW, Christie DM, Duncan RA (1996) Geochronology of Galapagos seamounts. Journal of Geophysical Research 101: 13689–13700. https://doi.org/10.1029/96JB00642
  • Sinton CW, Harpp KS, Geist DJ, Mittelstaedt E, Fornari DJ, Soule SA (2017) Banco Tuzo: An ancient Galapagos island and potential stepping stone for species dispersal. Galapagos Research 69: 9–17.
  • Sinton CW, Hauff F, Hoernle K, Werner R (2018) Age progressive volcanism opposite Nazca plate motion: Insights from seamounts on the northeastern margin of the Galapagos Platform. Lithos 310–311: 342–354. https://doi.org/10.1016/j.lithos.2018.04.014
  • Soderman CR, Shorttle O, Gazel E, Geist DJ, Matthews S, Williams HM (2023) The evolution of the Galápagos mantle plume. Science Advances 9: eadd5030. https://doi.org/10.1126/sciadv.add50
  • Steinfartz S, Glaberman S, Lanterbecq D, Marquez C, Rassmann K, Caccone A (2007) Genetic impact of a severe El Niño Event on Galápagos marine iguanas (Amblyrhynchus cristatus). PLoS ONE 2: e1285. https://doi.org/10.1371/journal.pone.0001285
  • Tiley GP, Flouri T, Jiao X, Poelstra JW, Xu B, Zhu T, Rannala B, Yoder AD, Yang Z (2023) Estimation of species divergence times in presence of cross-species gene flow. Systematic Biology 72: 820–836. https://doi.org/10.1093/sysbio/syad015
  • Tindle RW, Tindle LE, Vagenas D, Harris MP (2013) Population dynamics of the Galápagos flightless cormorant Phalacrocorax harrisi in relation to sea temperature. Marine Ornithology 41: 121–133.
  • Tindle RW, Tupiza A, Blomberg SP, Tindle LE (2016) The biology of an isolated population of the American flamingo Phoenicopterus ruber in the Galapagos Islands. Galapagos Research 68: 15–27.
  • Torres F, Huang EJ, Román-Carrion JL, Bever GS (2024) New insights into the origin of the Galápagos tortoises with a tip-dated analysis of Testudinidae. Journal of Vertebrate Paleontology 43: in press. https://doi.org/10.1080/02724634.2024.2313615
  • Torres-Carvajal O, Rodríguez-Guerra A, Chaves JA (2016) Present diversity of Galápagos leaf-toed geckos (Phyllodactylidae: Phyllodactylus) stems from three independent colonization events. Molecular Phylogenetics and Evolution 103: 1–5. https://doi.org/10.1016/j.ympev.2016.07.006
  • Valente LM, Phillimore AB, Etienne RS (2015) Equilibrium and non-equilibrium dynamics simultaneously operate in the Galápagos islands. Ecology Letters 18: 844–852. https://doi.org/10.1111/ele.12461
  • Vargas FH, Barlow S, Hart T, Jiménez-Uzcátegui G, Chavez J, Naranjo S, Macdonald DW (2008) Effects of climate variation on the abundance and distribution of flamingos in the Galapagos Islands. Journal of Zoology 276: 252–265. https://doi.org/10.1111/j.1469-7998.2008.00485.x
  • Vargas FH, Lacy RC, Johnson PJ, Steinfurth A, Crawford RJM, Boersma PD, Macdonald DW (2007) Modelling the effect of El Niño on the persistence of small populations: The Galápagos penguin as a case study. Biological Conservation 137: 138–148. https://doi.org/10.1016/j.biocon.2007.02.005
  • Villagómez DR, Toomey DR, Hooft EEE, Solomon SC (2007) Upper mantle structure beneath the Galápagos Archipelago from surface wave tomography. Journal Geophysical Research 112: B07303. https://doi.org/10.1029/2006JB004672
  • Weiss TL, Denniston RF, Wanamaker AD, Villarini G, von der Heydt AS (2017) El Niño-Southern Oscillation-like variability in a Late Miocene Caribbean coral. Geology 45: 643–646. https://doi.org/10.1130/G38796.1
  • Werner R, Hoernle K, Barckhausen U, Hauff F (2003) Geodynamic evolution of the Galápagos hot spot system (Central East Pacific) over the past 20 m.y.: Constraints from morphology, geochemistry, and magnetic anomalies. Geochemistry Geophysics Geosystems 4: 1108. https://doi.org/10.1029/2003GC000576
  • Whattam SA, Stern RJ (2015) Late Cretaceous plume-induced subduction initiation along the southern margin of the Caribbean and NW South America: The first documented example with implications for the onset of plate tectonics. Gondwana Research 27: 38–63. https://doi.org/10.1016/j.gr.2014.07.011
  • Wolf JB, Tautz D, Trillmich F (2007) Galápagos and Californian sea lions are separate species: Genetic analysis of the genus Zalophus and its implications for conservation management. Frontiers of Zoology 4: 20. https://doi.org/10.1186/1742-9994-4-20
  • Wright JW (1983) The evolution and biogeography of the lizards of the Galápagos Archipelago: Evolutionary genetics of Phyllodactylus and Tropidurus populations. In: Bowman RI, Berson M, Leviton AE (Eds) Patterns of Evolution in Galápagos Organisms. American Association for the Advancement of Science, Pacific Division, San Francisco, CA, 123–155.
  • Wyles JS, Sarich VM (1983) Are the Galápagos iguanas older than the Galápagos? In: Bowman RI, Berson M, Leviton AE (Eds) Patterns of Evolution in Galápagos Organisms. American Association for the Advancement of Science, Pacific Division, San Francisco, CA, 177–186.
  • Zaher H, Yánez-Muñoz MH, Rodrigues MT, Graboski R, Machado FA, Altamirano-Benavides M, Bonatto SL, Grazziotin FG (2018) Origin and hidden diversity within the poorly known Galápagos snake radiation (Serpentes: Dipsadidae). Systematics and Biodiversity 16: 614–642. https://doi.org/10.1080/14772000.2018.1478910

Supplementary material

Supplementary material 1 

Figures S1–S3

Ali JR, Fritz U (2024)

Data type: .pdf

Explanation notes: Figure S1. Map showing the endpoints of the various bathymetric-topographic profiles presented in Figures S2 and S3. — Figure S2. Series of S–N bathymetric-topographic profiles approximately perpendicular to the long axis of the Galápagos platform-Carnegie Ridge terrain. — Figure S3. Series of bathymetric profiles approximately perpendicular to the long axis of the Cocos Ridge.

This dataset is made available under the Open Database License (http://opendatacommons.org/licenses/odbl/1.0). The Open Database License (ODbL) is a license agreement intended to allow users to freely share, modify, and use this dataset while maintaining this same freedom for others, provided that the original source and author(s) are credited.
Download file (2.90 MB)
login to comment