Next Article in Journal
Genotyping-by-Sequencing Analysis Shows That Siberian Lindens Are Nested within Tilia cordata Mill
Previous Article in Journal
Real-Time Classification of Invasive Plant Seeds Based on Improved YOLOv5 with Attention Mechanism
Previous Article in Special Issue
How the Strength of Monsoon Winds Shape Forest Dynamics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Evolutionary and Biogeographical History of Penguins (Sphenisciformes): Review of the Dispersal Patterns and Adaptations in a Geologic and Paleoecological Context

by
Jonathan S. Pelegrín
1,* and
Carolina Acosta Hospitaleche
2
1
Grupo de Investigación en Ecología y Conservación de la Biodiversidad (EcoBio), Área de Biología y Programa de Maestría en Educación Ambiental y Desarrollo Sostenible, Universidad Santiago de Cali, Cali 760001, Colombia
2
División Paleontología de Vertebrados, Museo de la Plata, Facultad de Ciencias Naturales y Museo, Universidad Nacional de La Plata-CONICET, La Plata B1900FWA, Argentina
*
Author to whom correspondence should be addressed.
Diversity 2022, 14(4), 255; https://doi.org/10.3390/d14040255
Submission received: 26 December 2021 / Revised: 13 March 2022 / Accepted: 14 March 2022 / Published: 30 March 2022

Abstract

:
Despite its current low diversity, the penguin clade (Sphenisciformes) is one of the groups of birds with the most complete fossil record. Likewise, from the evolutionary point of view, it is an interesting group given the adaptations developed for marine life and the extreme climatic occupation capacity that some species have shown. In the present contribution, we reviewed and integrated all of the geographical and phylogenetic information available, together with an exhaustive and updated review of the fossil record, to establish and propose a biogeographic scenario that allows the spatial-temporal reconstruction of the evolutionary history of the Sphenisciformes, discussing our results and those obtained by other authors. This allowed us to understand how some abiotic processes are responsible for the patterns of diversity evidenced both in modern and past lineages. Thus, using the BioGeoBEARS methodology for biogeographic estimation, we were able to reconstruct the biogeographical patterns for the entire group based on the most complete Bayesian phylogeny of the total evidence. As a result, a New Zealand origin for the Sphenisciformes during the late Cretaceous and early Paleocene is indicated, with subsequent dispersal and expansion across Antarctica and southern South America. During the Eocene, there was a remarkable diversification of species and ecological niches in Antarctica, probably associated with the more temperate climatic conditions in the Southern Hemisphere. A wide morphological variability might have developed at the beginning of the Paleogene diversification. During the Oligocene, with the trends towards the freezing of Antarctica and the generalized cooling of the Neogene, there was a turnover that led to the survival (in New Zealand) of the ancestors of the crown Sphenisciform lineages. Later these expanded and diversified across the Southern Hemisphere, strongly linked to the climatic and oceanographic processes of the Miocene. Finally, it should be noted that the Antarctic recolonization and its hostile climatic conditions occurred in some modern lineages during the Pleistocene, possibly due to exaptations that made possible the repeated dispersion through cold waters during the Cenozoic, also allowing the necessary adaptations to live in the tundra during the glaciations.

1. Introduction

Penguins (Aves, Sphenisciformes) constitute a group of birds that are exclusively marine and flightless. All the species present extreme anatomical and physiological modifications directly related with the diving habit and the adaptations to cold-temperature waters [1,2]. From an evolutionary point of view, there is consensus to include the Sphenisciformes along with other aquatic birds in Aequornithes, and within this clade they are closely related to the Procellariiformes [3,4,5,6]. More precisely, the origin of penguins would be linked to a flying ancestor that secondarily would have lost the ability to fly as they became excellent divers capable of traveling long distances ([7] and numerous later contributions) and reaching extreme depths [1,2,8].
The Sphenisciformes would have originated at the ends of the Cretaceous [9,10,11,12,13] in Zealandia [14] or Te Riu-a-Māui (Māori) or Tasmantis, lands that emerge today as New Zealand. Their appearance and diversification would be closely related to the extinction of the large marine reptiles that played the role of top predators in the southern oceans [15]. Later, these niches became vacant and were occupied by other vertebrates such as penguins in the Southern Hemisphere ([16] and references therein). Although no Cretaceous penguins are known, the fossil record is consistent with this idea. The oldest records of penguins correspond to forms that are morphologically archaic [11,17,18,19,20] that probably acquired a great size, a non-pneumatic skeleton, a flattening of the wing bones constituting propelling blades for diving, and an incipient widening and shortening of the tarsometatarsus, during the lower Paleocene.
These and other specializations for wing-propelled diving are already present in the Paleocene species (Kupoupou stilwelli, Waimanu manneringi, Sequiwaimanu rosieae, Kumimanu biceae, Muriwaimanu tuatahi, Crossvallia waiparensis, and Crossvallia unienwillia), although in the Eocene, forms with more extreme morphophysiological specializations are evident. In this regard, features such as the development of a blood plexus in the wing are observed early in the evolution of penguins (see details in [21]). This acquisition allowed them better thermal regulation during cold-water forays [22,23], as did the presence of highly modified feathers transformed into scales that cover the wings and substantially improve hydrodynamic skills during diving [24].
An increase in body size and a greater adaptation for diving in cold water would have conferred an important adaptive advantage in this context, since a greater body size implies a greater diving capacity, both in terms of depth reached and the duration of the dive [25]. The maximum expression of body size was achieved in the Eocene, when Palaeeudyptes klekowskii reached more than two meters in height [26]. Although there is no consensus about how the size of Paleogene penguins should be calculated, several cases of giant species have been reported in Antarctica, South America, New Zealand, and Australia, covering almost all the areas where penguins are recorded. Thus, penguins reached their apogee with many shapes and an incredible diversity of sizes [27].
It has been proposed that large and robust penguins would have arrived at the Peruvian coasts through two successive colonizations from different areas. The first spread, from Antarctica, would have occurred by the middle Eocene, whereas the second colonization, from New Zealand, would have occurred by the end of the Eocene. According to this proposal, based on the Eocene record of Peruvian penguins [24,28], the presence of Antarctic forms in the middle Eocene in Chile [29] and Argentina [30] is also explained. This stage does not extend beyond the Oligocene. It is not possible to determine the causes or the exact mechanisms that caused these faunal changes, but the diversity of diving birds is inversely proportional to the diversity of marine mammals, especially odontocetes cetaceans. Giant penguins were extinguished where marine mammals became successful as the top predators in the oceans [31]. A new stage in the evolution of the group begins in the Neogene, which includes the appearance of modern forms closely related to living species [32,33]. Taxonomic and morphological diversification in living species is notably less than what was known in the past, and post-Pliocene species are almost entirely attributed to modern genera [34,35].
An example of the transition that occurred during the Neogene is the avian assemblage of Horcón, on the central coast of Chile, which reflects the existence of a mixed fauna during the Pliocene, connecting the seabird associations of the late Miocene with the modern regional avifauna [36]. However, the Cenozoic history of penguins seems to have been somewhat more complex than previously believed. The current avifauna would be the result of a series of successive colonizations and extinctions closely linked to the establishment and development of the ocean currents and the ecological dynamics of species [35,37]. A recent analysis identified New Zealand (either exclusively or with South America) as the most likely ancestral area for crown clade penguins [38].
Despite being a group with a low current diversity (18 species), considering the species known from the fossil record, the Sphenisciformes are one of the best-known avian clades, with about 65 recognized species [6,39]. Likewise, the phylogenetic relationships have led to the proposition of various phylogenetic hypotheses, which have been possible due to the good state of preservation of many fossils and the deep and widely comparative studies of the morphological features among the described lineages. In recent years, extensive morphological knowledge and the consolidation of molecular analysis techniques have allowed phylogenetic approaches to reconstruct the evolution of penguins by integrating extant and extinct forms [17,19,32,33,35,36,40,41,42,43].
Some approaches have generated hypotheses where the influence of events such as those that occurred during the Neogene on the biogeographic patterns and the evolution of the Sphenisciformes niche are reconstructed; however, many scenarios only consider the current species [35,43]. In this sense, the richness of the penguin fossil record [6] allows the possibility of considering and integrating all the available information to propose broader approximations in a deep time approach.
Thanks to the vast amount of information available on the presence of species during the Cenozoic in several locations of the Southern Hemisphere and modern biogeographic analyses methodologies, it is possible to reconstruct geographical scenarios of evolution over time and to understand the influence of environmental and geological changes on the diversification of penguins. In particular, BioGeoBEARS [44,45] analysis allows the reconstruction of ancestral areas in a context of maximum likelihood and employs Bayesian modeling from a calibrated phylogeny. Some previous contributions have dealt with this topic (e.g., [28,35,38]). We focus our review on detecting ancestral areas of origin and describing the paleobiogeographical patterns of the Sphenisciformes lineage based on a broad and complete analysis of the Sphenisciformes fossil record and the most recently published phylogenetic proposal based on the total evidence for the group [32]. This approach allows us to visualize the speciation, dispersal, and extinction events that would have occurred throughout their evolutionary history, shedding more light on how the environmental changes that occurred throughout the Cenozoic could have influenced the evolution and diversification patterns of penguins. This gives us the possibility of comparing our own results with the previous proposals.

2. Materials and Methods

2.1. Fossil Record and Penguin Phylogenies

According to the available scientific literature and the Paleobiology Database, we consolidated a new biogeographical and temporal matrix, considering all the records for penguin species (Table 1). In this way, we recorded the time intervals according to their chronostratigraphic distribution range and encoded the presence (1) or absence (0) of the species in each geographical area. It should be mentioned that although a single occurrence is the only data for some fossil species (e.g., Crossvallia unienwillia), the stratigraphical range provided in Table 1 corresponds to the age of the level where the fossil was collected. The same criterion applies for species with multiple records (e.g., Palaeeudyptes klekowskii), in which the stratigraphical range corresponds to the ages of the levels where it was reported. For species with an uncertain age, due to the lack of a strict stratigraphic control (i.e., Marplesornis novaezealandiae), the range includes a different-ages proposal. Table 1 includes the source of the data.
Data from 83 species (18 living and 65 fossil ones) were obtained. Given the need for a completely resolved and calibrated phylogeny to perform the BioGeoBEARS analysis, a review of the latest phylogenies proposed for Sphenisciformes was carried out. After considering the number of species included, the consistency of the calibrated ages, the degree of resolution, and the integration of multiple information sources, we applied the Bayesian total evidence phylogeny proposed by Gavryushkina et al. [32]. Another proposal, the Bayesian Markov Chain Monte Carlo (MCMC) framework for phylogenetic analysis, takes an extensive data source from molecular sequences derived from extant species and morphological traits from extant and fossil species. It also considers the stratigraphic intervals as the fossil occurrences. The phylogenetic proposal of Gavryushkina et al. considers the evolutionary affinities of penguins according to 202 morphological characters [42] derived from reasonably complete fossil specimens (n = 36), together with molecular and morphological information from the 18 living species [32]. With this input, this approach estimates and dates species phylogenies. The Bayesian method integrates the fossil information under a new perspective, unlike other methods that only use fossils to calibrate nodes or stablished origin intervals. For our purposes, we used the maximum sampled-ancestor clade credibility tree (the MSACC tree). This tree is a summary tree derived from a posterior sample that maximizes the product of posterior clade probabilities (see details in [80], cited in [32]). Other biogeographical proposals discussed below are based on different phylogenetic approaches [38,43] and references cited therein).

2.2. Species Considered in the Present Analysis

The description of the first fossil penguin was followed by a great proliferation of new genera and species, which after some years were re-evaluated and, in many cases, dismissed or considered as synonyms. This work took, as a starting point, a complete review of the fossil record for the Sphenisciformes lineage (Table 1, Figure 1). Even though the list of penguin fossil species is much more extensive, rigorous analyses carried out over recent decades have established long synonymic lists of species and genera that are no longer considered valid. Table 1 follows the taxonomic arrangements proposed for Argentinian [63,64,65,66,67,68,69,70,71], Chilean [67,69,72,73,78,81], Peruvian [24,28,41,57,69,73], Antarctic [27,49,50,51,52,53,54,57,58,59], New Zealand [11,17,19,20,33,38,42,48,55,60], Australian [56,67,75], and African [76,77] taxa. This compilation is essential to obtain complementary information for the discussion and the palaeoecological analysis.
A particular case worth commenting on is that of Eudyptula. In this work, the traditional and most widely analyzed proposal, in which Eudyptula minor would be the only modern species of the genus Eudyptula, was adopted as input for the present analysis. According to that proposal, the diversity of Eudyptula forms is reflected in the six subspecies inhabiting Australia and New Zealand [2,82]. Other more recent proposals consider that Eudyptula would be constituted by E. minor and E. novaehollandiae, species of recent divergence [83,84] that have been accepted as such by the ornithological community [85]. The inclusion of Eudyptula as the only living species does not modify or bias our results. Further, the incorporation of extinct species was constrained by several additional factors, including taxonomic status, given that some taxa are currently synonyms or have been considered non-valid taxa in subsequent revisions, and their previous inclusion in a phylogeny.
On the other hand, Spheniscus anglicus, a species described from materials that presumably come from the Miocene Bahía Inglesa Formation of Chile [86], was excluded from this analysis due to serious irregularities. The material was bought and removed the country illegally, violating the laws for the protection of the paleontological heritage in Chile. In this context, the species’ geographical and stratigraphic origin is not reliable. In addition, the characters used for its diagnosis are not adequate, and the proposal of a new species is unjustified. For these reasons, we decided to exclude S. anglicus from our analysis, a species that has never been listed or considered in any of the subsequent specialized scientific publications.
In short, despite not being included in the present biogeographical analysis, the information derived from all the species was not included in phylogenetic proposal of Gavryushkina et al. [32], which provided complementary information on the presence and diversity in the continental areas considered, allowing the enrichment of aspects of the discussion. The details of the fossil species considered here, and those included in our analysis, are provided in Table 1.

2.3. Paleobiogeographical Analyses

The biogeographic regions established for the analysis were chosen based on the extant and ancient distribution of Sphenisciformes species, as well as on geological and climatic criteria. Thus, we established six biogeographic regions or areas: north-central South America, including Galapagos (from 23° S), southern South America, southern Africa, Antarctica, Australia, and New Zealand, unlike the nine [28], ten [35], or the twelve [38] areas included in other contributions. For analysis, we proposed a flexible scenario for the dispersal events among the various study areas. This criterion was determined by the proximity and distances among the six areas and their geological histories linked to the fragmentation and drift of Gondwana since the Cretaceous and, later, during the Paleogene and the Neogene [87,88]. These drift processes triggered the oceanographical evolution of marine currents [89], which are key factors in the dispersal possibilities for penguins. Likewise, the possibilities for the colonization of areas were established based on the long-distance swimming characteristics observed in current penguins, which were presumably present in Paleogene forms according to fossil distribution since the Paleogene [33]. Given the outstanding dispersal and marine movement capacity reflected in modern species, as well as the Southern Hemisphere distribution of the fossil and modern species, a matrix of the probability of colonization was adjusted to 1 with respect to the studied areas.
In accordance with the BioGeoBEARS analysis [44], we carried out the evaluation of three models: Dispersal—Extinction Cladogenesis (DEC); a likelihood version of the Dispersal—Vicariance model (DIVALIKE); and a likelihood version of the BayArea model (BAYAREALIKE). The DEC model considers and emphasizes changes in the range of distribution in speciation events (cladogenesis). Under that model, during events a descendant lineage will always occupy a single region of the ancestral area, considering sympatry or vicariance. The DIVALIKE model allows a daughter lineage to retain more than a single geographical region of ancestral occupation during the vicariant event. This model does not allow a daughter lineage to inherit a small rank that is sympatric to the rank of another descendant lineage. Conversely, the BAYAREALIKE model does not emphasize geographic range variation at speciation events; instead, it estimates range changes along speciation events through range expansion−contraction dynamics., We assessed these models including the Jump-dispersal (+J) parameter [44,90]. This parameter allows evolutionary founding events, where an emerging novel lineage disperses outside of the area(s) occupied by its ancestor during the speciation process.
All the models were compared, considering the p-value for the LRT (Likelihood Ratio Test) and the value of the AIC for each evaluated scenario [91]. The estimation models using the methodologies derived from BioGeoBEARS have been applied to various bird taxa, and despite receiving criticism for the inclusion of the parameter J (the founder effect) [92], these models have been reevaluated, reinforcing, and supporting the validity of the models [93]. Thus, we incorporated founder-event speciation (+J), which results in a process that is important in island systems for birds, considering the importance of transcontinental colonization events during different bird clades diversification, and especially for penguins [35]. Specifically, models that included the +J parameter have been broadly consistent in explaining the colonization processes in biogeographic and macroevolutionary studies. Examples are the contributions on several lineages of modern birds, including the Megapodidae family within Galliformes [94], Thraupidae (Coerebinae) [90], Motacillidae [95], Coraciiformes [96], Trogoniformes [97], Rallidae [98], and those studies on fossil lineages, such as Coelurosauria clade [99], and mammals, such as horses (Equinae) [100]. With particular reference to penguins, previous works analyzed the crown group species [35] as well as fossil representatives [38]. In line with these works and the life-history traits of penguins, we considered that the +J parameter would be associated with Sphenisciformes macroevolutionary process, due to the remarkable oceanic dispersion capacity evidenced by modern and ancient forms [77,101,102]. The statistical analyses were performed using the software RASP powered by R software [103].

3. Results and Discussion

3.1. Paleogene History of Penguins

According to the results obtained here, the best model was the BAYAREALIKE + J, which provided the statistical support with the lowest AICc value and the highest AICw, compared to the other models (Figures S1–S6, Table 2 and Table S1); there are significant differences between the BAYAREALIKE + J and BAYAREALIKE model scenarios (Table S2). Similar results were obtained in previous contributions [38], although in other analyses the selected model was DIVALIKE + J [35]. As we expected, our results confirmed the relevance of the +J parameter (founder events) to explain the biogeographical history of penguins, a clade with a presumably well-stablished dispersal capacity due the early development of adaptative traits to navigate across marine environments; this is supported by the analyses of Paleogene forms, including those of Waimanu [11]. In addition, the BAYAREALIKE + J, as a better model, provides support for the importance of geographical expansion−contraction dynamics to explain the evolutionary patterns of Sphenisciformes. The Cenozoic cooling trends triggered many biomic expansions−contractions in Southern Hemisphere continents, which influenced the dispersal processes and possibly the speciation and extinction patterns.
In general terms, all the models pinpoint and concur with a center of origin for Sphenisciformes in New Zealand (Figures S1–S6). These results are consistent with previous estimates based on fossil findings, which also estimate the origin of the lineage towards the late Cretaceous [9,10,11,12,13]. This is a logical proposal given the high diversification and specialization already present in the Paleocene. The oldest records for penguins correspond to the Paleocene and are concentrated in New Zealand [17,18,19,20,104].
The results of our analysis pinpoints New Zealand as the most likely ancestral area, and secondly points to Antarctica with slightly lower probabilities (Figure 2; see the Supplementary Materials for details). It is noteworthy that New Zealand’s importance as a center of origin is strongly supported by a high concentration of records, many of them being the oldest penguin records reported to date [6,19,20,104] (Figure 1, Table 1). It should be noted that New Zealand’s geographical proximity to the Antarctic territory during the Upper Cretaceous and the early Paleogene provides some evidence of the significance of both continents during the initial diversification of the group. Findings for the Chatham Islands, and specifically associated with the Takatika Grit, show that since the Upper Cretaceous (c.83−79 Ma) Zealandia began to present a progressive rupture with respect to West Antarctica [14,87,105], continuing until the Eocene with respect to the eastern Antarctic region [106]. At the end of this stage, Zealandia would have experienced a strong marine transgression [107,108]. This process might explain the notable radiation and rapid diversification of penguins during the Eocene for Antarctica, as compared with New Zealand. The abundant fossil record of Seymour Island (Antarctica) strongly supports this idea. In this sense, the wider Antarctic territory would have offered greater opportunities, in terms of colonization of new niches and thus the generation of diverse processes of speciation, due to geographic isolation.
Our results, like previous findings [28,38], allow us to postulate New Zealand as a probable main ancestral territory. In addition, is important to consider the geographical proximity between New Zealand and Antarctica during the Paleogene; both territories during the Paleocene-Eocene climatic optimum would have presented very similar environmental conditions at the continental level, with cold temperate environments that would have presented periods of fluctuation towards warm temperate climates during the Early Eocene Climatic Optimum (EECO). This would have made it possible to configure humid temperate forest biomes, with tropical floristic components during several Paleogene intervals [109,110,111]. Likewise, marine estimates show significant warming of Pacific waters from the upper Paleocene to the middle Eocene [46,112,113]. This paleoenvironmental context could have favored the dispersal between New Zealand and Antarctica, given the importance that oceanic temperatures possibly played for the dispersal of the first penguins. Added to this is the Crossvallia record in Antarctica and New Zealand, which further strengthened the links between these two large areas during the first million years of the group’s evolution. Crossvallia unienwillia was a large penguin species, with a single record in the Paleocene of Seymour Island (Antarctica); due to the incompleteness of its skeleton, it has been repeatedly omitted from phylogenetic analyses. Its presence, however, indicates the presence of Sphenisciformes in Antarctica since the Paleocene [46,47]. Crossvallia waiparensis is the second species of Crossvallia that has been described for the Paleocene of New Zealand (Figure 3).
The recent description of numerous taxa for the Paleocene of New Zealand indicates favorable conditions for the establishment and flourishing of the group. Although only two species of the genus Waimanu have been included (Figure 2), the New Zealand Paleocene sphenicofauna also includes other species such as Kupoupou stilwelli, Crossvallia waiparensis, Sequiwaimanu rosieae, and Kumimanu biceae (Figure 1 and Figure 2). Although the Antarctic record is scarce during the Paleocene, this is probably due to a taphonomic bias rather than to regional environmental conditions, since the changes in the depositional environment of the James Ross Basin during the Eocene caused a more abundant and diverse penguin record [114].
During the middle Eocene, several lineages diversified in Antarctica, including forms of a wide spectrum of sizes, including some giant penguins such as Anthropornis grandis that reached 1.7 m high, and other tiny penguins such as Aproskoditos microtero that were only 0.35 m high. This shows great diversity, also evidenced in the number of species included in other genera, such as Delphinornis, Tonniornis, Palaeeudyptes, and others (see Table 1). This broad diversity is probably associated with niche partitioning processes powered by the development of different bill morphologies and specializations in a wide range of trophic possibilities [115,116]. Among these taxa are the forms that reached the southern and central South American coasts, allowing the establishment of the Perudyptes devriesi lineages on the Peruvian coasts. The record of taxa typically Antarctic in southern South America during the middle Eocene [29,30] supports this hypothesis (Figure 4).
During the late Eocene and probably the early Oligocene, the Antarctic species would have completely disappeared. A highlighted diversification of the New Zealand lineages is evidenced at this time. Some colonizations in South America, such as those of Icadyptes salasi and Inkayaku paracacensis, are verified during the Eocene of Peru, and would be closely linked with the penguin fauna of New Zealand. According to our results, the three lineages (together with Perudyptes devriesi) would have independently colonized the subtropical Pacific coasts of South America during the late Eocene. These colonizations could be related with migrations produced by oceanic currents established from New Zealand to South America during the Eocene after the EEOC, with the opening of the Tasman Strait and the Drake Passage. The currents suffered notable alterations at the latitudinal level, ending in the establishment of the circum-Antarctic current, the main influencing factor in progressive Antarctic freezing during the Oligocene [117,118,119] (Figure 5).
In this sense, we proposed that New Zealand could have played an important role as a refugee during the Oligocene for penguins that faced the climatic changes that transformed the Antarctic continent and the marine current regimes [117,118,119,120]. This idea is aligned with the presence of the Kaiika lineage, a taxa endemic to New Zealand [48], and the diversity of the genus Kairuku with three species recorded for the New Zealand Oligocene [60]. Likewise, Palaeeudyptes, of presumably Antarctic origin, would have been present in New Zealand, as evidenced, for example, by Palaeeudyptes marplesi [42]. Therefore, the fossil findings suggest that after the extinction of almost all of the Antarctic forms, Palaeeudyptes could have been one of the few lineages that would have colonized and persisted in New Zealand (Figure 5).

3.2. Neogene History of Penguins

According to our results, the taxa recorded in Patagonia (Argentina), would have had a New Zealand origin. Presumably, the establishment of the Antarctic circumpolar current would have allowed the dispersion of the lineages from New Zealand to southern South America, possibly given the similar environmental conditions in both places that might have been decisive for aspects related to the feeding and breeding areas. In this way, from the colonization of southern South America at the beginning of the Miocene, several lineages would have developed. First, Paraptenodyes (including P. robustus and P. antarcticus), and later, Eretiscus tonni and Palaeospheniscus (with a high diversity constituted by P. bergi, P. patagonicus, and P. biloculata), established a wide presence in southern South America, as evidenced by the fossil record of Patagonia Argentina [61,64,70,71], and reached the coasts of Chile and Peru by the middle Miocene [81,121].
The Miocene was a crucial time for the establishment of the most modern faunas [36,71,122]. Our results suggest that from New Zealand to Southern South America, three biogeographical events that were probably related with the intensification of the Antarctic circumpolar current during the middle and late Miocene [89], deserve to be highlighted. This new scenario favors the selection of physiological and biochemical adaptations to face colder environmental regimes, an idea strongly supported by genomic studies [35,122]. Thus, our results are consistent with the biogeographical proposals from the crown group [35,38]; see also [123].
The first event corresponds to the diversification of Spheniscus, widely recorded during the Mio-Pliocene of Chile and Peru [36]. Here, an origin in the south of South America and a dispersion towards the north, which was probably influenced by the beginning of the establishment process of the Humboldt current during the middle Miocene (15−12 Ma) is proposed. The ecological preferences of the Spheniscus lineage are consistent with colder waters and a diet based on fish [116]; it could be possible that these traits were inherited from their ancestors. This is also supported by the fossil record of these areas and even by the fossil record of Antarctic [54]. By the middle Miocene, there is a vast record of penguins attributed to Spheniscus, mainly in Chile and Perú, represented by species such as Spheniscus urbinai, S. megaramphus, S. muizoni, and S. chilensis. The diversification of the modern lineages corresponding to Spheniscus would have been relatively recent. Respectively, S. humboldti and S. demersus might have colonized the central-northern Pacific coast of South America and the coasts of southern Africa from southern South America during the Pleistocene [77,124]. In addition, and recently, S. humboldti colonized the Galapagos archipelago, allowing the origin of S. mendiculus [122]. These processes were probably related to the expansion of the polar caps during the glaciations, reaching almost 40° South latitude, altering the structure of the marine currents and the latitudinal thermal gradient [125]. Thus, our results are consistent with previous proposals [77] of multiple colonization events in Africa for Sphenisciformes. This is supported by the presence of Nucleornis insolitus, Dege hendeyi, and Inguza predemersus in the fossil record, which colonized Africa independently at the end of the Miocene (Figure 6 and Figure 7).
A second biogeographical event corresponds to the clade Megadyptes-Eudyptes, with a probable common ancestor in New Zealand. These results are consistent with previous analyses about the biogeographical history for this clade [38]. In addition, our findings suggest a strong generalist condition for these geographical occupations. The lineage would have developed wide dispersal capacities around Antarctica, reaching multiple continental islands close to the mainland masses, which would have generated advantages in terms of the absence of possible predators and competition for resources. However, the cooling processes that intensified during the Plio-Pleistocene led to the formation and growth of ice caps in the Antarctic Ocean. Therefore, these glacial and interglacial intervals would have generated isolation and subsequent speciation in some of these lineages [35]. On the other hand, the scenarios of allopatric speciation by isolation in islands for the Eudyptes lineage are discussed by some authors, such as Frugone et al. [126], who proposed a greater effect of the thermal zonation of the Antarctic polar front and the subtropical currents on the definition of species. Consequently, the strong dispersal capacity and a more generalist condition would not have allowed the necessary genetic isolation and subsequent speciation, as seems to be evidenced in E. schlegeli and E. chrysolophus. In this way, Eudyptes would be an example of a generalist lineage that, with its different species, migrated from New Zealand throughout the Southern Hemisphere, reaching southern South America, subantarctic islands, and Africa [35,126] (Figure 6 and Figure 7).
During the middle Miocene a third diversification process of the clade Pygoscelis-Aptenodytes is revealed in our results. The radiation center was probably from southern South America with lineages such as Madrynornis endemic to Patagonia, and a common ancestor of Pygoscelis and Aptenodytes with an outstanding skill of dispersion in the circum-Antarctic waters. These ocean currents became colder and colder since the intensification of the circum-Antarctic current 11 Ma ago, thanks to the development of biochemical and metabolic adaptations associated with thermoregulation, optimization of oxygen consumption and ATP production [35,122,126]. This would have allowed them to reach a circum-Antarctic distribution during the middle and late Miocene, as evidenced by the presence of Pygoscelis tyreei and Aptenodytes ridgeni in New Zealand, as well as the presence of P. grandis and P. calderensis in southern South America. The southern parts of Argentina and New Zealand may have been linked during the late Middle Miocene as areas of constant exchange of species with other regions, powered by the latitudinal direction of marine currents [118]. Finally, together with the cooling and the Pleistocene glaciations, processes of population isolation occurred by the formation of polar caps and the consequent changes in the currents. Some patterns triggered by glacial-interglacial intervals modified the genetic flow between populations and promoted isolation scenarios and the subsequent speciation in the Antarctic lineages. Consequently, the adaptations previously developed along the crown group lineage evolution, which allowed the occupation of more extreme thermal niches in increasingly cold waters, would have been key exaptations in the colonization and subsequent biomic specialization in extreme tundra conditions. The modern species Aptenodytes forsteri and Pygoscelis adeliae are examples of that process (Figure 6 and Figure 7).

4. Conclusions

Despite using a different phylogenetic proposal, a flexible scenario for dispersal possibilities, and an alternative areas delimitation, our results are broadly consistent with previous findings about the main paleobiogeographical patterns during penguins’ evolutionary history. Thus, our findings are broadly consistent with a New Zealand center of origin for Sphenisciformes during the late Cretaceous and early Paleogene, supporting the hypothesis generated by analyses proposed by diverse authors [28,38]. With respect to the Eocene, we found an outstanding diversification and dispersion of penguins geographically associated with Antarctica, due to the establishment of temperate conditions triggered by PETM and EECO. The Oligocene and early Miocene represented a turnover in the Sphenicofauna; the extinction of Antarctic lineages consolidated New Zealand and Southern South America as refuges associated with the latitudinal contraction of temperate biomes and warm marine currents. The outcomes suggest that crown group Sphenisciformes flourished during the Miocene and many adaptations from their ancestors would probably be established as exaptations to face increasingly cold environmental conditions during the Neogene. Thus, some lineages expanded their areas towards subtropical latitudes in South America and Africa, while other lineages (Pygoscelis and Aptenodytes) developed the colonization capacity for the hardest climatic environments, such as the tundra conditions in Antarctica during the Pleistocene glaciations. All these statements are, however, provisional, and subject to new findings and subsequent analyses. Although the penguin record is quite complete in comparison with those of other bird taxa, several deficiencies and important gaps are recognized during the time periods considered. We trust, however, that the efforts of numerous researchers currently working on these studies will at least partially reverse these findings.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/d14040255/s1, Figure S1: Biogeographical results of BioGeoBEARS under DEC model; Figure S2: Biogeographical results of BioGeoBEARS under DEC + J model; Figure S3: Biogeographical results of BioGeoBEARS under DIVALIKE model; Figure S4: Biogeographical results of BioGeoBEARS under DIVALIKE + J model; Figure S5: Biogeographical results of BioGeoBEARS under BAYAREALIKE model; Figure S6: Biogeographical results of BioGeoBEARS under BAYAREALIKE + J model; Table S1: Results of model test with statistical support of BioGeoBEARS analyses for all models; Table S2: Complementary results of model test with statistical support of BioGeoBEARS analyses comparing all models according to +J parameter.

Author Contributions

Conceptualization, J.S.P., C.A.H.; methodology, J.S.P., C.A.H.; formal analysis, J.S.P., C.A.H.; investigation, J.S.P., C.A.H.; resources, J.S.P., C.A.H.; data curation, J.S.P., C.A.H.; writing—original draft preparation, J.S.P., C.A.H. writing—review and editing, J.S.P., C.A.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the General Research Council (Dirección General de Investigaciones-DGI) at the Universidad Santiago de Cali (Colombia) under call No. 01-2021; La Plata University PI N955 (Argentina), and the National Scientific and Technical Research Council PIP 0096 (Argentina).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on the text in the Section 2.

Acknowledgments

We thank A. Givryushkina for allowing us to use her phylogenetic proposal and for being ready to collaborate with us; Leonardo Belalcazar for the computational logistical support; and Jacobo Sabogal for the opinions and technical support in graphic design and penguin illustrations. Finally, we acknowledge the anonymous reviewers who undoubtedly contributed to the improvement of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Williams, T.D.; Busby, J. Bird Families of the World. 2. The Penguins: Spheniscidae; Oxford University Press: Oxford, UK, 1995. [Google Scholar]
  2. Winkler, D.W.; Billerman, S.M.; Lovette, I.J. Bird Families of the World: An Invitation to the Spectacular Diversity of Birds; Lynx Editions: Cerdanyola del Vallès, Spain, 2015. [Google Scholar]
  3. Mayr, G. Metaves, Mirandornithes, Strisores and other novelties–a critical review of the higher-level phylogeny of neornithine birds. J. Zoolog. Syst. Evol. 2011, 49, 58–76. [Google Scholar] [CrossRef]
  4. Jarvis, E.D.; Mirarab, S.; Aberer, A.J.; Li, B.; Houde, P.; Li, C.; Zhang, G. Whole-genome analyses resolve early branches in the tree of life of modern birds. Science 2014, 346, 1320–1331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Prum, R.O.; Berv, J.S.; Dornburg, A.; Field, D.J.; Townsend, J.P.; Lemmon, E.M.; Lemmon, A.R. A comprehensive phylogeny of birds (Aves) using targeted next-generation DNA sequencing. Nature 2015, 526, 569–573. [Google Scholar] [CrossRef] [PubMed]
  6. Mayr, G. Avian Evolution: The Fossil Record of Birds and Its Paleobiological Significance; John Wiley & Sons: Hoboken, NJ, USA, 2017. [Google Scholar]
  7. Simpson, G.G. Penguins: Past and Present, Here and There; Yale University Press: New Haven, CT, USA, 1976. [Google Scholar]
  8. Kooyman, G.L.; Cherel, Y.; Le Maho, J.; Croxall, P.; Thorson, P.H.; Ridoux, V.; Kooyman, C.A. Diving behavior and energetics during foraging cycles in King Penguins. Ecol. Monogr. 1992, 62, 143–163. [Google Scholar] [CrossRef]
  9. Cooper, A.; Penny, D. Mass survival of birds across the Cretaceous-Tertiary boundary: Molecular evidence. Science 1997, 275, 1109–1113. [Google Scholar] [CrossRef]
  10. Hedges, S.B.; Parker, P.H.; Sibley, C.G.; Kumar, S. Continental breakup and the ordinal diversification of birds and mammals. Nature 1996, 381, 226–229. [Google Scholar] [CrossRef] [PubMed]
  11. Slack, K.E.; Jones, C.M.; Ando, T.; Harrison, G.L.; Fordyce, R.E.; Arnason, U.; Penny, D. Early penguin fossils, plus mitochondrial genomes, calibrate avian evolution. Mol. Biol. Evol. 2006, 23, 1144–1155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Feduccia, A. Explosive evolution in Tertiary birds and mammals. Science 1995, 267, 637–638. [Google Scholar] [CrossRef] [PubMed]
  13. Feduccia, A. ‘Big bang’ for Tertiary birds? Trends Ecol. Evol. 2003, 18, 172–176. [Google Scholar] [CrossRef]
  14. Gaina, C.; Müller, D.R.; Royer, J.Y.; Stock, J.; Hardebeck, J.; Symonds, P. The tectonic history of the Tasman Sea: A puzzle with 13 pieces. J. Geophys. Res. 1998, 103, 12413–12433. [Google Scholar] [CrossRef] [Green Version]
  15. Ksepka, D.T.; Ando, T. Penguins past, present, and future: Trends in the evolution of the Sphenisciformes. In Living Dinosaurs; John Wiley & Sons: Hoboken, NJ, USA, 2011; pp. 155–186. [Google Scholar]
  16. Fordyce, R.E.; Jones, C.M. Penguin history and new fossil material from New Zealand. In Penguin Biology; Davis, L.S., Darby, J.T., Eds.; Academic Press: San Diego, CA, USA, 1990; pp. 419–446. [Google Scholar]
  17. Mayr, G.; De Pietri, V.L.; Scofield, R.P. A new fossil from the mid-Paleocene of New Zealand reveals an unexpected diversity of world’s oldest penguins. Sci. Nat. 2017, 104, 9. [Google Scholar] [CrossRef]
  18. Mayr, G.; Scofield, R.P.; De Pietri, V.L.; Tennyson, A.J. A Paleocene penguin from New Zealand substantiates multiple origins of gigantism in fossil Sphenisciformes. Nat. Commun. 2017, 8, 1927. [Google Scholar] [CrossRef]
  19. Mayr, G.; De Pietri, V.L.; Love, L.; Mannering, A.A.; Scofield, R.P. A well-preserved new mid-Paleocene penguin (Aves, Sphenisciformes) from the Waipara Greensand in New Zealand. J. Vertebr. Paleontol. 2018, 37, e1398169. [Google Scholar] [CrossRef]
  20. Mayr, G.; De Pietri, V.L.; Love, L.; Mannering, A.; Scofield, R.P. Leg bones of a new penguin species from the Waipara Greensand add to the diversity of very large-sized Sphenisciformes in the Paleocene of New Zealand. Alcheringa 2020, 44, 194–201. [Google Scholar] [CrossRef]
  21. Thomas, D.B.; Fordyce, R.E. Biological plasticity in penguin heat-retention structures. Anat. Rec. 2012, 295, 249–256. [Google Scholar] [CrossRef]
  22. Thomas, D.B.; Ksepka, D.; Fordyce, E. Penguin heat-retention structures evolved in a greenhouse Earth. Biol. Lett. 2011, 7, 461–464. [Google Scholar] [CrossRef] [Green Version]
  23. Acosta Hospitaleche, C.; De Los Reyes, M.; Santillana, S.; Reguero, M. First fossilized skin of a giant penguin from the Eocene of Antarctica. Lethaia 2010, 53, 409–420. [Google Scholar] [CrossRef]
  24. Clarke, J.A.; Ksepka, D.T.; Salas-Gismondi, R.; Altamirano, A.J.; Shawkey, M.D.; D’Alba, L.; Vinther, J.; DeVries, T.J.; Baby, P. Fossil evidence for evolution of the shape and color of penguin feathers. Science 2010, 330, 954–957. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Watanuki, Y.; Burger, A.E. Body mass and dive duration in Alcids and Penguins. Can. J. Zool. 1999, 77, 1838–1999. [Google Scholar] [CrossRef]
  26. Acosta Hospitaleche, C. New giant penguin bones from Antarctica: Systematic and paleobiological significance. C. R. Palevol. 2014, 13, 555–560. [Google Scholar] [CrossRef]
  27. Acosta Hospitaleche, C.; Reguero, M.A.; Santillana, S. Aprosdokitos mikrotero gen. et sp. nov., the tiniest Sphenisciformes that lived in Antarctica during the Paleogene. Neues Jahrb. Geol. Palaontol. Abh. 2017, 283, 25–34. [Google Scholar] [CrossRef]
  28. Clarke, J.A.; Ksepka, D.T.; Stucchi, M.; Urbina, M.; Giannini, N.; Bertelli, S.; Narváez, Y.; Boyd, C.A. Paleogene equatorial penguins challenge the proposed relationship between biogeography, diversity, and Cenozoic climate change. Proc. Natl. Acad. Sci. USA 2007, 104, 11545–11550. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Sallaberry, M.A.; Yury-Yáñez, R.E.; Otero, R.A.; Soto-Acuña, S.; Torres, T. Eocene birds from the western margin of southernmost South America. J. Paleontol. 2010, 84, 1061–1070. [Google Scholar] [CrossRef]
  30. Acosta Hospitaleche, C.; Olivero, E. Re-evaluation of the fossil penguin Palaeeudyptes gunnari from the Eocene Leticia Formation, Argentina: Additional material, systematics and palaeobiology. Alcheringa 2016, 40, 373–382. [Google Scholar] [CrossRef]
  31. Ando, T.; Fordyce, R.E. Evolutionary drivers for flightless, wing-propelled divers in the Northern and Southern Hemispheres. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2014, 400, 50–61. [Google Scholar] [CrossRef]
  32. Gavryushkina, A.; Heath, T.A.; Ksepka, D.T.; Stadler, T.; Welch, D.; Drummond, A.J. Bayesian total-evidence dating reveals the recent crown radiation of penguins. Syst. Biol. 2017, 66, 57–73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Blokland, J.C.; Reid, C.M.; Worthy, T.H.; Tennyson, A.J.; Clarke, J.A.; Scofield, R.P. Chatham Island Paleocene fossils provide insight into the palaeobiology, evolution, and diversity of early penguins (Aves, Sphenisciformes). Palaeontol. Electron. 2019, 22, 78. [Google Scholar] [CrossRef] [Green Version]
  34. Subramanian, S.; Beans-Picón, G.; Swaminathan, S.K.; Millar, C.D.; Lambert, D.M. Evidence for a recent origin of penguins. Biol. Lett. 2013, 9, 20130748. [Google Scholar] [CrossRef]
  35. Vianna, J.A.; Fernandes, F.A.; Frugone, M.J.; Figueiró, H.V.; Pertierra, L.R.; Noll, D.; Bowie, R.C. Genome-wide analyses reveal drivers of penguin diversification. Proc. Natl. Acad. Sci. USA 2020, 117, 22303–22310. [Google Scholar] [CrossRef]
  36. Chávez-Hoffmeister, M.; Briceño, J.D.; Nielsen, S.N. The evolution of seabirds in the Humboldt Current: New clues from the Pliocene of central Chile. PLoS ONE 2014, 9, e90043. [Google Scholar] [CrossRef]
  37. Boessenkool, S.; Austin, J.J.; Worthy, T.H.; Scofield, P.; Cooper, A.; Seddon, P.J.; Waters, J.M. Relict or colonizer? Extinction and range expansion of penguins in southern New Zealand. Proc. R. Soc. B 2009, 276, 815–821. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Thomas, D.B.; Tennyson, A.J.D.; Scofield, R.P.; Heath, T.A.; Pett, W.; Ksepka, D.T. Ancient crested penguin constrains timing of recruitment into seabird hotspot. Proc. R. Soc. B 2020, 287, 20201497. [Google Scholar] [CrossRef] [PubMed]
  39. Paleobiology Database. The Paleobiology Database. Checklist Dataset. 2021. Available online: GBIF.org (accessed on 23 December 2021). [CrossRef]
  40. Acosta Hospitaleche, C.; Canto, J.; Tambussi, C.P. Pingüinos (Aves, Spheniscidae) en Coquimbo (Mioceno medio-Plioceno tardío), Chile y su vinculación con las corrientes oceánicas. Span. J. Palaeont. 2006, 21, 115–121. [Google Scholar] [CrossRef]
  41. Ksepka, D.T.; Clarke, J.A. The basal penguin (Aves: Sphenisciformes) Perudyptes devriesi and a phylogenetic evaluation of the penguin fossil recordBull. Am. Mus. Nat. Hist. 2010, 337, 1–77. [Google Scholar] [CrossRef]
  42. Ksepka, D.T.; Fordyce, R.E.; Ando, T.; Jones, C.M. New fossil penguins (Aves, Sphenisciformes) from the Oligocene of New Zealand reveal the skeletal plan of stem penguins. J. Vertebr. Paleontol. 2012, 32, 235–254. [Google Scholar] [CrossRef]
  43. Degrange, F.J.; Ksepka, D.T.; Tambussi, C.P. Redescription of the oldest crown clade penguin: Cranial osteology, jaw myology, neuroanatomy, and phylogenetic affinities of Madrynornis mirandus. J. Vertebr. Paleontol. 2018, 38, e1445636. [Google Scholar] [CrossRef]
  44. Matzke, N.J. BioGeoBEARS: Biogeography with Bayesian (and Likelihood) Evolutionary Analyses in R Scripts; BioGeoBEARS; The Comprehensive R Archive Network: Berkeley, CA, USA, 2013; Available online: http://CRAN.R-project.org/package (accessed on 23 November 2021).
  45. Matzke, N.J. Probabilistic historical biogeography: New models for founder event speciation, imperfect detection, and fossils allow improved accuracy and model-testing. Front. Biogeogr. 2013, 5, 242–248. [Google Scholar] [CrossRef] [Green Version]
  46. Tambussi, C.P.; Reguero, M.A.; Marenssi, S.A.; Santillana, S.N. Crossvallia unienwillia, a new Spheniscidae (Sphenisciformes, Aves) from the late Paleocene of Antarctica. Geobios 2005, 38, 667–675. [Google Scholar] [CrossRef]
  47. Jadwiszczak, P.; Acosta Hospitaleche, C.; Reguero, M. Redescription of Crossvallia unienwillia: The only Paleocene Antarctic penguin. Ameghiniana 2013, 50, 545–553. [Google Scholar] [CrossRef]
  48. Fordyce, R.E.; Thomas, D.B. Kaiika maxwelli, a new Early Eocene archaic penguin (Sphenisciformes, Aves) from Waihao Valley, South Canterbury, New Zealand. N. Z. J. Geol. Geophys. 2011, 54, 43–51. [Google Scholar] [CrossRef]
  49. Myrcha, A.; Jadwiszczak, P.; Tambussi, C.P.; Noriega, J.I.; Gaździcki, A.; Tatur, A.; del Valle, R.A. Taxonomic revision of Eocene Antarctic penguins based on tarsometatarsal morphology. Pol. Polar Res. 2002, 23, 5–46. [Google Scholar]
  50. Jadwiszczak, P.; Mörs, T. First partial skeleton of Delphinornis larseni Wiman, 1905, a slender-footed penguin from the Eocene of Antarctic Peninsula. Palaeontol. Electron. 2019, 22, 1–31. [Google Scholar] [CrossRef] [Green Version]
  51. Jadwiszczak, P.; Reguero, M.; Mörs, T. A new small-sized penguin from the late Eocene of Seymour Island with additional material of Mesetaornis polaris. GFF 2021, 143, 283–291. [Google Scholar] [CrossRef]
  52. Acosta Hospitaleche, C.; Jadwiszczak, P.; Clarke, J.A.; Cenizo, M. The fossil record of birds from the James Ross Basin, West Antarctica. Adv. Polar Sci. 2019, 30, 251–273. [Google Scholar] [CrossRef]
  53. Jadwiszczak, P. Partial limb skeleton of a “giant penguin” Anthropornis from the Eocene of Antarctic Peninsula. Pol. Polar Res. 2012, 3, 259–274. [Google Scholar] [CrossRef]
  54. Acosta Hospitaleche, C.; Reguero, M.; Scarano, A. Main pathways in the evolution of Antarctic fossil penguins. J. S. Am. Earth Sci. 2013, 43, 101–111. [Google Scholar] [CrossRef]
  55. Ando, T. New Zealand Fossil Penguins: Origin, Pattern, and Process. Ph.D. Thesis, University of Otago, Dunedin, New Zealand, 2007. [Google Scholar]
  56. Park, T.; Fitzgerald, E.M. A review of Australian fossil penguins (Aves: Sphenisciformes). Mem. Mus. Vic. 2012, 69, 309–325. [Google Scholar] [CrossRef] [Green Version]
  57. Tambussi, C.P.; Acosta Hospitaleche, C.; Reguero, M.A.; Marenssi, S.A. Late Eocene penguins from West Antarctica: Systematics and biostratigraphy. Geol. Soc. Spec. Publ. 2006, 258, 145–161. [Google Scholar] [CrossRef]
  58. Jadwiszczak, P.; Acosta Hospitaleche, C. Distinguishing between two Antarctic species of Eocene Palaeeudyptes penguins: A statistical approach using tarsometatarsi. Pol. Polar Res. 2013, 34, 237–252. [Google Scholar] [CrossRef] [Green Version]
  59. Acosta Hospitaleche, C.; Reguero, M. Palaeeudyptes klekowskii, the best-preserved penguin skeleton from the Eocene–Oligocene of Antarctica: Taxonomic and evolutionary remarks. Geobios 2014, 47, 77–85. [Google Scholar] [CrossRef]
  60. Giovanardi, S.; Ksepka, D.T.; Thomas, D.B. A giant Oligocene fossil penguin from the North Island of New Zealand. J. Vertebr. Paleontol. 2021, 41, e1953047. [Google Scholar] [CrossRef]
  61. Acosta Hospitaleche, C. Systematic revision of Arthrodytes Ameghino, 1905 (Aves, Spheniscidae) and its assignment to the Paraptenodytinae. Neues Jahrb. Geol. Paläontol. 2005, 7, 404–414. [Google Scholar] [CrossRef] [PubMed]
  62. Parras, A.; Dix, G.R.; Griffin, M. Sr-isotope chronostratigraphy of Paleogene–Neogene marine deposits: Austral Basin, southern Patagonia (Argentina). J. S. Am. Earth Sci. 2012, 37, 122–135. [Google Scholar] [CrossRef]
  63. Acosta Hospitaleche, C.; Tambussi, C.; Cozzuol, M. Eretiscus tonnii (Simpson) (Aves, Sphenisciformes): Materiales adicionales, status taxonómico y distribución geográfica. Rev. Mus. Argent. Cienc. Nat. 2004, 6, 233–237. [Google Scholar]
  64. Acosta Hospitaleche, C.; Griffin, M.; Asensio, M.; Cione, A.L.; Tambussi, C. Middle Cenozoic penguin remains from the Patagonian Cordillera. Andean Geol. 2013, 40, 490–503. [Google Scholar]
  65. Acosta Hospitaleche, C. Revisión sistemática de Palaeospheniscus biloculata (Simpson) nov. comb.(Aves, Spheniscidae) de la Formación Gaiman (Mioceno Temprano), Chubut, Argentina. Ameghiniana 2007, 44, 417–426. [Google Scholar]
  66. Acosta Hospitaleche, C.; Castro, L.; Tambussi, C.; Scasso, R.A. Palaeospheniscus patagonicus (Aves, Sphenisciformes): New discoveries from the early Miocene of Argentina. J. Paleontol. 2008, 82, 565–575. [Google Scholar] [CrossRef]
  67. Walsh, S.A.; Suárez, M.E. New penguin remains from the Pliocene of northern Chile. Hist. Biol. 2006, 18, 119–130. [Google Scholar] [CrossRef]
  68. Gohlich, U.B. The oldest fossil record of the extant penguin genus Spheniscus-a new species from the Miocene of Peru. Acta Palaeontol. Pol. 2007, 52, 285–298. [Google Scholar]
  69. Acosta Hospitaleche, C.; Paulina-Carabajal, A.; Yury-Yáñez, R. The skull of the Miocene Spheniscus urbinai (Aves, Sphenisciformes): Osteology, brain morphology, and the cranial pneumatic systems. J. Anat. 2021, 239, 151–166. [Google Scholar] [CrossRef] [PubMed]
  70. Acosta Hospitaleche, C.; Tambussi, C.; Donato, M.; Cozzuol, M. A new Miocene penguin from Patagonia and its phylogenetic relationships. Acta Palaeontol. Pol. 2007, 52, 299–314. [Google Scholar]
  71. Viglino, M.; Buono, M.; Acosta Hospitaleche, C.; Cione, A.; Cuitiño, J.; Gaetán, M.; Sterli, J.; Paolucci, F. Vertebrados marinos del Cenozoico. In Relatorio XXI Congreso Geológico Argentino; Geología y Recursos Naturales de la Provincia del Chubut: Puerto Madryn, Argentine, 2021; pp. 335–358. [Google Scholar]
  72. Acosta Hospitaleche, C.; Chávez-Hoffmeister, M.; Fritis, O. Pingüinos fósiles (Pygoscelis calderensis sp. nov.) en la Formación Bahía Inglesa (Mioceno Medio-Plioceno), Chile. Rev. Geol. Chile 2006, 33, 327–338. [Google Scholar] [CrossRef]
  73. Chávez-Hoffmeister, M. Fossil birds of Chile and Antarctic Peninsula. Arq. Mus. Nac. Rio Janeiro 2007, 65, 551–572. [Google Scholar]
  74. Stucchi, M. Los Pingüinos Fósiles de la Formación Pisco (Neógeno). In Proceedings of the 4° European Meeting on the Palaeontology and Stratigraphy of Latin America, Cuadernos del Museo Geominero 8, Perú, Tres Cantos, Madrid, 12–14 September 2007; Díaz–Martínez, E., Rábano, I., Eds.; Instituto Geológico y Minero de España: Madrid, Spain, 2007. [Google Scholar]
  75. Park, T. Redescription of the Miocene penguin Pseudaptenodytes macraei Simpson (Aves: Sphenisciformes) and redefinition of the taxonomic status of? Pseudaptenodytes minor Simpson. Alcheringa. 2014, 38, 450–454. [Google Scholar] [CrossRef]
  76. Simpson, G.G. A new genus of late Tertiary penguin from Langebaanweg, South Africa. Ann. S. Afr. Mus. 1979, 78, 1–9. [Google Scholar]
  77. Ksepka, D.T.; Thomas, D.B. Multiple cenozoic invasions of Africa by penguins (Aves, Sphenisciformes). Proc. R. Soc. B 2012, 279, 1027–1032. [Google Scholar] [CrossRef]
  78. Emslie, S.D.; Guerra Correa, C. A new species of penguin (Spheniscidae: Spheniscus) and other birds from the late Pliocene of Chile. Proc. Biol. Soc. Wash. 2003, 116, 308–316. [Google Scholar]
  79. Marples, B.J. Fossil penguins from the mid-Tertiary of Seymour Island. FIDS Sci. Rep. 1953, 5, 1–15. [Google Scholar]
  80. Heled, J.; Bouckaert, R.R. Looking for trees in the forest: Summary tree from posterior samples. BMC Evol. Biol. 2013, 13, 221. [Google Scholar] [CrossRef] [Green Version]
  81. Chávez-Hoffmeister, M. The humerus and stratigraphic range of Palaeospheniscus (Aves, Sphenisciformes). Ameghiniana 2014, 51, 159–172. [Google Scholar] [CrossRef]
  82. Kinsky, F.C.; Falla, R.A. A subspecific revision of the Australasian Blue Penguin (Eudyptula minor) in the New Zealand area. Rec. Nat. Mus. N 1976, 2, 105–126. [Google Scholar]
  83. Grosser, S.; Burridge, C.P.; Peucker, A.J.; Waters, J.M. Coallescent modelling reveals recent secondary contact of cryptic penguin species. PLoS ONE 2015, 10, e0144966. [Google Scholar] [CrossRef]
  84. Grosser, S.; Scofield, R.P.; Waters, J.M. Multivariate skeletal analyses support a taxonomic distinction between New Zealand and Australian Eudyptula penguins Sphenisciformes: Spheniscidae). Emu 2017, 117, 276–283. [Google Scholar] [CrossRef]
  85. Worthy, T.H.; Grant-Mackie, J.A. Late-Pleistocene avifaunas from Cape Wanbrow, Otago, South Island, New Zealand. J. R. Soc. N. Z. 2003, 33, 427–485. [Google Scholar] [CrossRef]
  86. Benson, R.D. A new species of penguin from the late Miocene of Chile with comments on the stratigraphic range of Palaeospheniscus. Sci. Pub. Sci. Mus. MN 2015, 8, 22. [Google Scholar]
  87. Strogen, D.P.; Seebeck, H.; Nicol, A.; King, P.R. Two-phase Cretaceous–Paleocene rifting in the Taranaki Basin region, New Zealand; implications for Gondwana break-up. J. Geol. Soc. 2017, 174, 929–946. [Google Scholar] [CrossRef]
  88. Storey, B.C.; Granot, R. Tectonic history of Antarctica over the past 200 million years. Geol. Soc. Lond. Mem. 2021, 55, 9–17. [Google Scholar] [CrossRef]
  89. Holbourn, A.; Kuhnt, W.; Frank, M.; Haley, B.A. Changes in Pacific Ocean circulation following the Miocene onset of permanent Antarctic ice cover. Earth Planet Sci. Lett. 2013, 365, 38–50. [Google Scholar] [CrossRef]
  90. Funk, E.R.; Burns, K.J. Biogeographic origins of Darwin’s finches (Thraupidae: Coerebinae). Auk 2018, 135, 561–571. [Google Scholar] [CrossRef] [Green Version]
  91. Burnham, K.P.; Anderson, D.R. Model Selection and Multimodel Inference; Springer: New York, NY, USA, 2002. [Google Scholar]
  92. Ree, R.H.; Sanmartín, I. Conceptual and statistical problems with the DEC+ J model of founder-event speciation and its comparison with DEC via model selection. J. Biogeogr. 2018, 45, 741–749. [Google Scholar] [CrossRef]
  93. Klaus, K.V.; Matzke, N.J. Statistical comparison of trait-dependent biogeographical models indicates that Podocarpaceae dispersal is influenced by both seed cone traits and geographical distance. Syst. Biol. 2020, 69, 61–75. [Google Scholar] [CrossRef] [PubMed]
  94. Harris, R.B.; Birks, S.M.; Leaché, A.D. Incubator birds: Biogeographical origins and evolution of underground nesting in megapodes (Galliformes: Megapodiidae). J. Biogeogr. 2014, 41, 2045–2056. [Google Scholar] [CrossRef]
  95. Van Els, P.; Norambuena, H.V.; Etienne, R.S. From pampa to puna: Biogeography and diversification of a group of Neotropical obligate grassland birds (Anthus: Motacillidae). J. Zool. Syst. Evol. Res. 2019, 57, 485–496. [Google Scholar] [CrossRef]
  96. McCullough, J.M.; Moyle, R.G.; Smith, B.T.; Andersen, M.J. A Laurasian origin for a pantropical bird radiation is supported by genomic and fossil data (Aves: Coraciiformes). Proc. R. Soc. B 2019, 286, 20190122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Oliveros, C.H.; Andersen, M.J.; Hosner, P.A.; Mauck, W.M., III; Sheldon, F.H.; Cracraft, J.; Moyle, R.G. Rapid Laurasian diversification of a pantropical bird family during the Oligocene-Miocene transition. IBIS 2020, 162, 137–152. [Google Scholar] [CrossRef]
  98. Garcia-R, J.C.; Matzke, N.J. Trait-dependent dispersal in rails (Aves: Rallidae): Historical biogeography of a cosmopolitan bird clade. Mol. Phylogenet. Evol. 2021, 159, 107106. [Google Scholar] [CrossRef] [PubMed]
  99. Ding, A.; Pittman, M.; Upchurch, P.; O’Connor, J.M.; Field, D.J.; Xu, X. The biogeography of coelurosaurian theropods and its impact on their evolutionary history. Bull. Am. Mus. Nat. Hist. 2020, 40, 117–157. [Google Scholar]
  100. Cantalapiedra, J.L.; Prado, J.L.; Hernández Fernández, M.; Alberdi, M.T. Decoupled ecomorphological evolution and diversification in Neogene-Quaternary horses. Science 2017, 355, 627–630. [Google Scholar] [CrossRef] [Green Version]
  101. Peucker, A.J.; Dann, P.; Burridge, C.P. Range-wide phylogeography of the little penguin (Eudyptula minor): Evidence of long-distance dispersal. Auk 2009, 126, 397–408. [Google Scholar] [CrossRef]
  102. Baker, A.J.; Pereira, S.L.; Haddrath, O.P.; Edge, K.A. Multiple gene evidence for expansion of extant penguins out of Antarctica due to global cooling. Proc. R. Soc. B 2006, 273, 11–17. [Google Scholar] [CrossRef] [Green Version]
  103. Yu, Y.; Harris, A.J.; Blair, C.; He, X. RASP (Reconstruct Ancestral State in Phylogenies): A tool for historical biogeography. Mol. Phylogenet. Evol. 2015, 87, 46–49. [Google Scholar] [CrossRef]
  104. Mayr, G.; De Pietri, V.L.; Love, L.; Mannering, A.A.; Bevitt, J.J.; Scofield, R.P. First complete wing of a stem group sphenisciform from the Paleocene of New Zealand sheds light on the evolution of the penguin flipper. Diversity 2020, 12, 46. [Google Scholar] [CrossRef] [Green Version]
  105. Stilwell, J.D.; Consoli, C.P. Tectono-stratigraphic history of the Chatham Islands, SW Pacific—The emergence, flooding and reappearance of eastern ‘Zealandia’. Proc. Geol. Assoc. 2012, 123, 170–181. [Google Scholar] [CrossRef]
  106. Tulloch, A.J.; Mortimer, N.; Ireland, T.R.; Waight, T.E.; Maas, R.; Palin, J.; Sahoo, T.; Seebeck, H.; Sagar, M.W.; Barrier, A. Reconnaissance basement geology and tectonics of South Zealandia. Tectonics 2019, 38, 516–551. [Google Scholar] [CrossRef] [Green Version]
  107. Bache, F.; Mortimer, N.; Sutherland, R.; Collot, J.; Rouillard, P.; Stagpoole, V.; Nicol, A. Seismic stratigraphic record of transition from Mesozoic subduction to continental breakup in the Zealandia Sector of eastern Gondwana. Gondwana Res. 2014, 26, 1060–1078. [Google Scholar] [CrossRef]
  108. Rouillard, P.; Collot, J.; Sutherland, R.; Bache, F.; Patriat, M.; Etienne, S.; Maurizot, P. Seismic stratigraphy and paleogeographic evolution of Fairway Basin, Northern Zealandia, Southwest Pacific: From Cretaceous Gondwana breakup to Cenozoic Tonga–Kermadec subduction. Basin Res. 2015, 29, 189–212. [Google Scholar] [CrossRef] [Green Version]
  109. Poole, I.; Cantrill, D.; Utescher, T. A mulitproxy approach to determine Antarctic terrestrial palaeoclimate during the Late Cretaceous and Early Tertiary. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2005, 222, 95–121. [Google Scholar] [CrossRef]
  110. Pross, J.; Contreras, L.; Bijl, P.K.; Greenwood, D.R.; Bohaty, S.M.; Schouten, S.; Brinkhuis, H. Persistent near-tropical warmth on the Antarctic continent during the early Eocene epoch. Nature 2012, 488, 73–77. [Google Scholar] [CrossRef] [PubMed]
  111. Kemp, D.B.; Robinson, S.A.; Crame, J.A.; Francis, J.E.; Ineson, J.; Whittle, R.J.; O’Brien, C. A cool temperate climate on the Antarctic Peninsula through the latest Cretaceous to early Paleogene. Geology 2014, 42, 583–586. [Google Scholar] [CrossRef] [Green Version]
  112. Bijl, P.; Schouten, S.; Slujis, A.; Reichart, G.; Zachos, J.C.; Brinkhuis, H. Early Palaeogene temperature evolution of the southwest Pacific Ocean. Nature 2009, 461, 776–779. [Google Scholar] [CrossRef] [PubMed]
  113. Hollis, C.J.; Taylor, K.W.; Handley, L.; Pancost, R.D.; Huber, M.; Creech, J.B.; Zachos, J.C. Early Paleogene temperature history of the Southwest Pacific Ocean: Reconciling proxies and models. Earth Planet Sci. Lett. 2012, 349, 53–66. [Google Scholar] [CrossRef]
  114. Marenssi, S.A. Eustatically controlled sedimentation recorded by Eocene strata of the James Ross Basin, Antarctica. Geol. Soc. Spec. Publ. 2006, 258, 125–133. [Google Scholar] [CrossRef]
  115. Haidr, N.; Acosta Hospitaleche, C. Feeding habits of Antarctic Eocene penguins from a morphofunctional perspective. Neues Jahrb. Geol. Paläontol. Abh. 2012, 263, 125–131. [Google Scholar] [CrossRef]
  116. Chávez-Hoffmeister, M. Bill disparity and feeding strategies among fossil and modern penguins. Paleobiology 2020, 46, 176–192. [Google Scholar] [CrossRef]
  117. Zachos, J.; Pagani, M.; Sloan, L.; Thomas, E.; Billups, K. Trends, rhythms, and aberrations in global climate 65 Ma to present. Science 2001, 292, 686–693. [Google Scholar] [CrossRef]
  118. Lawver, L.A.; Gahagan, L.M. Evolution of Cenozoic seaways in the circum-Antarctic region. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2003, 198, 11–37. [Google Scholar] [CrossRef]
  119. Houben, A.J.; Bijl, P.K.; Sluijs, A.; Schouten, S.; Brinkhuis, H. Late Eocene Southern Ocean cooling and invigoration of circulation preconditioned Antarctica for full-scale glaciation. Geochem. Geophys. 2019, 20, 2214–2234. [Google Scholar] [CrossRef] [Green Version]
  120. Pandey, M.; Pant, N.C.; Arora, D.; Gupta, R. A review of Antarctic ice sheet fluctuations records during Cenozoic and its cause and effect relation with the climatic conditions. Polar Sci. 2021, 30, 100720. [Google Scholar] [CrossRef]
  121. Acosta Hospitaleche, C.; Stucchi, M. Nuevos restos terciarios de Spheniscidae (Aves, Sphenisciformes) procedentes de la costa del Perú. Span. J. Paleontol. 2005, 20, 1–5. [Google Scholar] [CrossRef]
  122. Ramos, B.; González-Acuña, D.; Loyola, D.E.; Johnson, W.E.; Parker, P.G.; Massaro, M.; Dantas, G.; Marcelo, M.; Vianna, J.A. Landscape genomics: Natural selection drives the evolution of mitogenome in penguins. BMC Genom. 2018, 19, 53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Kooyman, G.L. Evolutionary and ecological aspects of some Antarctic and sub-Antarctic penguin distributions. Oecologia 2001, 130, 485–495. [Google Scholar] [CrossRef]
  124. Dantas, G.P.; Oliveira, L.R.; Santos, A.M.; Flores, M.D.; Melo, D.R.; Simeone, A.; González-Acuña, D.; Luna-Jorquera, G.; Le Bohec, C.; Valdés-Velásquez, A.; et al. Uncovering population structure in the Humboldt penguin (Spheniscus humboldti) along the Pacific coast at South America. PLoS ONE 2019, 14, e0215293. [Google Scholar] [CrossRef] [PubMed]
  125. Glasser, N.F.; Jansson, K.N.; Harrison, S.; Kleman, J. The glacial geomorphology and Pleistocene history of South America between 38 S and 56 S. Quat. Sci. Rev. 2018, 27, 365–390. [Google Scholar] [CrossRef]
  126. Frugone, M.J.; Lowther, A.; Noll, D.; Ramos, B.; Pistorius, P.; Dantas, G.P.M.; Vianna, J.A. Contrasting phylogeographic pattern among Eudyptes penguins around the Southern Ocean. Sci. Rep. 2018, 8, 17481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Updated chronostratigraphic distribution of fossil penguins of the world (n = 65). Follow color key to geographical occurrence (see Table 1 for details): (ak) indicates the species of penguin illustrations. On the right side (bottom-up), the living forms Aptenodytes forsteri, Pygoscelis papua, Spheniscus magellanicus, and Eudyptes chrysolophus are representatives of each genus. Penguins not at scale. Penguin illustration credits: Jacobo Sabogal.
Figure 1. Updated chronostratigraphic distribution of fossil penguins of the world (n = 65). Follow color key to geographical occurrence (see Table 1 for details): (ak) indicates the species of penguin illustrations. On the right side (bottom-up), the living forms Aptenodytes forsteri, Pygoscelis papua, Spheniscus magellanicus, and Eudyptes chrysolophus are representatives of each genus. Penguins not at scale. Penguin illustration credits: Jacobo Sabogal.
Diversity 14 00255 g001
Figure 2. Ancestral range estimation for Sphenisciformes based on results of high percentages for nodes considering the BAYAREALIKE + J scenario and using the six-area regime as shown in map of biogeographic areas powered in BioGeoBEARS. (al) indicates the species of penguin illustrations. Follow the color key for the cases of presence in more than one area. For details see the Supplementary Materials. Penguin illustration credits: Jacobo Sabogal.
Figure 2. Ancestral range estimation for Sphenisciformes based on results of high percentages for nodes considering the BAYAREALIKE + J scenario and using the six-area regime as shown in map of biogeographic areas powered in BioGeoBEARS. (al) indicates the species of penguin illustrations. Follow the color key for the cases of presence in more than one area. For details see the Supplementary Materials. Penguin illustration credits: Jacobo Sabogal.
Diversity 14 00255 g002
Figure 3. Middle-Late Paleocene biogeographical events: origin of Sphenisciformes in New Zealand (Kupoupou stilwelli in the image) and early dispersion to Antarctica, evidenced by the presence of Crossvallia (in the image).
Figure 3. Middle-Late Paleocene biogeographical events: origin of Sphenisciformes in New Zealand (Kupoupou stilwelli in the image) and early dispersion to Antarctica, evidenced by the presence of Crossvallia (in the image).
Diversity 14 00255 g003
Figure 4. Main Eocene biogeographical events: the diversification of diverse Sphenisciformes lineages in Antarctica (i.e., Palaeeudyptes and Delphinornis in the image) and early dispersal and colonization towards South America, evidenced by the presence of Perudyptes and Incayacu (in the image) and Icadyptes.
Figure 4. Main Eocene biogeographical events: the diversification of diverse Sphenisciformes lineages in Antarctica (i.e., Palaeeudyptes and Delphinornis in the image) and early dispersal and colonization towards South America, evidenced by the presence of Perudyptes and Incayacu (in the image) and Icadyptes.
Diversity 14 00255 g004
Figure 5. Main Oligocene biogeographical events: the extinction of Sphenisciformes due to Antarctica cooling; New Zealand as a refuge and center of diversification, evidenced by the presence of many genera and species, such as the Kairuku (in the image).
Figure 5. Main Oligocene biogeographical events: the extinction of Sphenisciformes due to Antarctica cooling; New Zealand as a refuge and center of diversification, evidenced by the presence of many genera and species, such as the Kairuku (in the image).
Diversity 14 00255 g005
Figure 6. Main Miocene biogeographical events: the colonization of lineages from New Zealand to South America due to circum-Antarctic oceanic currents (i.e., Paraptenodytes in the image); diversification and expansion of Spheniscus across South America.; origin and diversification of Megadyptes-Eudyptes clade from New Zealand; diversification of clade Pygoscelis in southern South America.
Figure 6. Main Miocene biogeographical events: the colonization of lineages from New Zealand to South America due to circum-Antarctic oceanic currents (i.e., Paraptenodytes in the image); diversification and expansion of Spheniscus across South America.; origin and diversification of Megadyptes-Eudyptes clade from New Zealand; diversification of clade Pygoscelis in southern South America.
Diversity 14 00255 g006
Figure 7. Main Plio-Pleistocene biogeographical events: Spheniscus dispersion towards South Africa; Pygoscelis and Aptenodytes lineages colonized Antarctica; Eudyptes lineage from New Zealand expanding across circum-Antarctic islands.
Figure 7. Main Plio-Pleistocene biogeographical events: Spheniscus dispersion towards South Africa; Pygoscelis and Aptenodytes lineages colonized Antarctica; Eudyptes lineage from New Zealand expanding across circum-Antarctic islands.
Diversity 14 00255 g007
Table 1. Updated checklist of the fossil penguin species of the world (n = 65), their occurrences, and stratigraphical ranges (SR) (in some cases, only an approximation is provided, corresponding to the age of the level where the fossil was collected).
Table 1. Updated checklist of the fossil penguin species of the world (n = 65), their occurrences, and stratigraphical ranges (SR) (in some cases, only an approximation is provided, corresponding to the age of the level where the fossil was collected).
SpeciesLocationEpochSR (Ma)Reference
Kupoupou stilwelliNew ZealandPaleocene62.5–60[33]
Crossvallia waiparensisNew ZealandPaleocene~61[20]
Muriwaimanu tuatahi * New ZealandPaleocene58–60[20]
Sequiwaimanu rosieaeNew ZealandPaleocene~61[19]
Waimanu manneringi * New ZealandPaleocene60.5–61.6[11]
Crossvalia unienwilliaAntarcticaPaleocene59.2[46,47]
Kumimanu biceaeNew ZealandPaleocene60.5–61.6[17]
Kaiika maxwelliNew ZealandEocene55.8–49.3[48]
Perudyptes devriesi * PeruEocene~42[28,41]
Delphinornis gracilis * AntarcticaEocene41–34[49]
Delphinornis larsenni * AntarcticaEocene53–34[50]
Mesetaornis polaris * AntarcticaEocene53–34[49,51]
Anthropornis grandis * AntarcticaEocene53–34[52,53,54]
Anthropornis nordenskjöldi * AntarcticaEocene53–34[52,54]
Aprosdokitos mikroteroAntarcticaEocene38–34[27]
Marambiornis exilis * AntarcticaEocene53–34[49,54]
Delphinornis arctowskii * AntarcticaEocene38–34[49,54]
Delphinornis wimani *,aAntarcticaEocene53–34[49,54]
Icadyptes salasi * PeruEocene37.2–35.7[28]
Inkayaku paracacensis * PeruEocene37.2–35.7[24]
Marambiornopsis sobraliAntarcticaEocene37.8–41.1[51]
Pachydyptes ponderosus * New ZealandEocene36–30[55]
Pachydyptes simpsoniAustraliaEocene38–36.5[56]
Palaeeudyptes antarcticus * New ZealandEocene–Oligocene38–28[42]
Palaeeudyptes marplesiNew ZealandEocene38–34[42]
Tonniornis mesetaensisAntarcticaEocene37.8–34[54,57]
Tonniornis minimumAntarcticaEocene37.8–34[54,57]
Palaeeudyptes gunnari * AntarcticaEocene55–34[30,54,58]
Palaeeudyptes klekowskii * AntarcticaEocene52–34[54,58,59]
Kairuku waewaeroaNew ZealandOligocene34–27.3[60]
Archaeospheniscus lopdelli * New ZealandOligocene27–25[55]
Archaeospheniscus lowei * New ZealandOligocene27–25[55]
Kairuku grebneffi * New ZealandOligocene27.3–25.2[42]
Kairuku waitaki * New ZealandOligocene27.3–34.5[42]
Korora oliveriNew ZealandOligocene 25–24[55]
Pakudyptes hakatarameaNew ZealandOligocene 25–24[55]
Platydyptes amiesiNew ZealandOligocene26–24[55]
Platydyptes marplesi * New ZealandOligocene–Miocene27–22[55]
Platydyptes novaezealandiae * New ZealandOligocene26–24[55]
Duntroonornis parvus * New ZealandOligocene–Miocene27–24[55]
Paraptenodytes robustusArgentinaOligocene–Miocene25–22[61,62]
Arthrodytes andrewsiArgentinaOligocene–Miocene25–22[61,62]
Eretiscus tonni * ArgentinaMiocene23–20.44[63]
Palaeospheniscus bergi * ArgentinaMiocene23–20.44[64]
Palaeospheniscus biloculata * ArgentinaMiocene23–20.44[65]
Palaeospheniscus patagonicus * ArgentinaMiocene23–20.44[66]
Paraptenodytes antarcticus * ArgentinaMiocene23–20.44[61]
Anthropodyptes gilliAustraliaMiocene21–17.6[67]
Spheniscus muizoni * PeruMiocene13–11[68]
Spheniscus urbinai * Argentina, Chile, PerúMiocene23–5[69]
Madrynornis mirandus * ArgentinaMiocene11.4–9[70,71]
Pygoscelis calderensisChileMiocene~7.6[69,72]
Marplesornis novaezealandiae * New ZealandMiocene–Pliocene12.7–2.4[55]
Spheniscus megaramphus * Chile, PeruMiocene–Pliocene11.6–3.6[73,74]
Pseudaptenodytes macraeiAustraliaMiocene–Pliocene6.2–5[75]
Dege hendeyiSouth AfricaPliocene5.3–3.6[76]
Inguza predemersusSouth AfricaPliocene5[77]
Nucleornis insolitusSouth AfricaPliocene5[77]
Eudyptula calauinaChilePliocene3.6–2.6[36]
Spheniscus chilensisChilePliocene3.6–2.6[78]
Eudyptes atatuNew ZealandPliocene3.3–3[38]
Tereingaornis moisleyiNew ZealandPliocene3–4[55]
Pygoscelis grandis * ChilePliocene5.3–3.6[67]
Pygoscelis tyreeiNew ZealandPliocene–Pleistocene4–2[55]
Aptenodytes ridgeniNew ZealandPliocene–Pleistocene4–2[55]
* Species included in the paleobiogeographical analyses. a We agree with Jadwiszczak [52,78] regarding the prematurity of the new combination Delphinornis wimani [41] for a species that already transferred from Notodyptes to Archaeospheniscus [79]. However, we maintain the new name for this table in accordance with the phylogeny on which we have based our biogeographical analyses [32].
Table 2. Summary of results for all six models evaluated under the six-area regime. Models with +J indicate those allowing for founder effect dispersals. The best-supported model is shown in bold. p is the number of parameters.
Table 2. Summary of results for all six models evaluated under the six-area regime. Models with +J indicate those allowing for founder effect dispersals. The best-supported model is shown in bold. p is the number of parameters.
ModelLnLpAICcAICc wt.
DEC−139.922842.6 × 10−6
DEC + J−1333272.58 × 10−4
DIVALIKE−151.42307.12.5 × 10−11
DIVALIKE + J−143.732941.8 × 10−8
BAYAREALIKE−150.22304.68.9 × 10−11
BAYAREALIKE + J−125.93258.31.00
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pelegrín, J.S.; Acosta Hospitaleche, C. Evolutionary and Biogeographical History of Penguins (Sphenisciformes): Review of the Dispersal Patterns and Adaptations in a Geologic and Paleoecological Context. Diversity 2022, 14, 255. https://doi.org/10.3390/d14040255

AMA Style

Pelegrín JS, Acosta Hospitaleche C. Evolutionary and Biogeographical History of Penguins (Sphenisciformes): Review of the Dispersal Patterns and Adaptations in a Geologic and Paleoecological Context. Diversity. 2022; 14(4):255. https://doi.org/10.3390/d14040255

Chicago/Turabian Style

Pelegrín, Jonathan S., and Carolina Acosta Hospitaleche. 2022. "Evolutionary and Biogeographical History of Penguins (Sphenisciformes): Review of the Dispersal Patterns and Adaptations in a Geologic and Paleoecological Context" Diversity 14, no. 4: 255. https://doi.org/10.3390/d14040255

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

Article Metrics

Back to TopTop