Open Access
How to translate text using browser tools
13 March 2013 Phylogeny and Evolutionary Patterns of South American Octodontoid Rodents
Diego H. Verzi, A. Itatí Olivares, Cecilia C. Morgan
Author Affiliations +
Abstract

Octodontoidea is the most diverse clade of hystricognath rodents, and is richly recorded in South America since at least the Oligocene. A parsimony-based morphological phylogenetic analysis of a wide range of extant and extinct octodontoids recovered three major clades, here recognised as Echimyidae, Octodontidae, and Abrocomidae. Taxa previously assigned to Echimyidae or Octodontoidea incertae sedis are here interpreted for the first time as early representatives of Ctenomyinae (Octodontidae), Octodontinae or Abrocomidae. Based on our results, we estimate the divergence of octodontoid families and subfamilies to have occurred during the Late Oligocene, which is consistent with molecular estimates, but older than previous inferences based on the fossil record. Contrary to previous suggestions, we show the first appearances of modern members of Abrocomidae, Octodontinae and Ctenomyinae to be distinctly decoupled from the origin of these clades, with different stages in the evolutionary history of octodontoids seemingly following distinct phases of palaeoenvironmental change. Depending on the phylogenetic pattern, fossils from the stage of differentiation bear evolutionary information that may not be provided by crown groups, thus highlighting the unique and important contribution of fossils to our understanding of macroevolutionary patterns.

Introduction

Octodontoidea is the most diverse clade of hystricognath rodents (Woods and Kilpatrick 2005). In the Recent South American fauna, it includes the fossorial to subterranean Octodontidae (degus, coruros, and tuco-tucos), the terrestrial and scansorial Abrocomidae (chinchilla rats), and the arboreal, semiaquatic, and terrestrial to fossorial Echimyidae (spiny rats and coypus) (Redford and Eisenberg 1992; Eisenberg and Redford 1999; Emmons and Feer 1999). In particular, the families Echimyidae (including Myocastor, sometimes included in a family of its own; Table 1) and Octodontidae (including Ctenomyinae, sometimes considered a family in their own right; Table 1, and discussion in Verzi 2001) comprise more than 60% of the extant species of South American hystricognaths (Woods and Kilpatrick 2005; Upham and Patterson 2012) and have the richest fossil record of the entire suborder (McKenna and Bell 1997).

The abundant fossil record of octodontoids extends at least as far back as the Oligocene (Wood and Patterson 1959; Patterson and Wood 1982; Vucetich et al. 1999, 2010b) and possibly even the Eocene (Frailey and Campbell 2004; Antoine et al. 2012). While the monophyly of Octodontoidea is not disputed (Opazo 2005; Honeycutt 2009; Vilela et al. 2009; Upham and Patterson 2012), the phylogenetic relationships of most living and fossil octodontoids, and in particular those from the Oligocene-Middle Miocene, remain poorly understood (e.g., Patterson and Wood 1982; Vucetich and Kramarz 2003; Vucetich et al. 1999, 2010a). Most of these early octodontoids have been assigned to Echimyidae owing to their lophate, low-crowned molars (Carvalho and Salles 2004). The latter are retained by extant members of this family, which inhabit mainly forested areas in northern South America (Reig 1986). By contrast, Octodontidae and Abrocomidae are thought to have originated during the Late Miocene, based on the first appearance of species with simplified, hypsodont molars — which are characteristic of their extant representatives — in Argentina and Bolivia (Rovereto 1914; Reig 1989; McKenna and Bell 1997; Vucetich et al. 1999). However, this taxonomic arrangement has been disputed by several authors (Winge 1941; Wood and Patterson 1959; Patterson and Wood 1982; Verzi 2002; Verzi et al. 2011).

Here, we reevaluate the evolutionary history of octodontoids through a comprehensive phylogenetic analysis including a wide range of extant and extinct taxa. In addition, we provide a new set of intra-clade divergence time estimates based on the fossil record.

Institutional abbreviations.—IMCN-CM, Instituto y Museo de Ciencias Naturales, San Juan, Argentina; MACN, Museo Argentino de Ciencias Naturales “Bernardino Rivadavia”, Buenos Aires, Argentina; MLP, Museo de La Plata, La Plata, Argentina; MMP, Museo de Ciencias Naturales de Mar del Plata “Lorenzo Scaglia”, Mar del Plata, Argentina; MPEF, Museo Paleontológico Egidio Feruglio, Trelew, Argentina.

Other abbreviations.—MSM*, Stratigraphic Manhattan Measure; TBR, tree bisection reconnection; t1-3, times of origin (t1) and differentiation (t2) of a clade, and of emergence of its crown group (t3).

Material and methods

Our phylogenetic analysis is based on 73 cranio-mandibular and dental characters scored for 53 extant and extinct Octodontoid genera (SOM 1, Supplementary Online Material available at  http://app.pan.pl/SOM/appXX-Verzi_etal_SOM.pdf). The character matrix was generated through first-hand revision of nearly 2000 specimens representing all of the South American octodontoid families and subfamilies, plus the extant Antillean Capromys (Capromyidae) (SOM 2). The extant cavioid Dasyprocta (Dasyproctidae) was used as outgroup. The informal name “†Acaremys group” is here used to refer to the closely related genera †Sciamys and †Acaremys (Scott 1905; Vucetich and Kramarz 2003), whose contents and independence remain in need of revision. We analysed our matrix using the “traditional search” option of TNT v. 1.1 (Goloboff et al. 2008a, b), using 10 000 random stepwise-addition replicates and tree bisection reconnection (TBR) branch swapping, saving 100 trees per replicate. In addition, we performed an extra round of TBR on the optimal trees to increase the chance of finding all topologies of minimum length (Bertelli and Giannini 2005). All characters were equally weighted and (except for character 49; SOM 1) treated as non-additive. Zero-length branches were collapsed if they lacked support under any of the most parsimonious reconstructions (Coddington and Scharff 1994). Branch support was calculated in the form of absolute and relative Bremer indices (Bremer 1994; Goloboff and Farris 2001).

The modified Stratigraphic Manhattan Measure (MSM*; Pol and Norell 2001) was used to assess the fit of the resulting phylogeny with the stratigraphic record. Minimum divergence date estimates for individual clades were based on the oldest, reliably dated fossil confidently assigned to either of the two lineages arising from each branching point (SOM 3) (Benton and Donoghue 2007). We calculated the MSM* both for the entire tree and for the two major octodontoid clades, which are similar in size and hence comparable (Pol et al. 2004).

Results

Phylogenetic relationships.—Our analysis resulted in three most parsimonious trees of 139 steps (CI = 0.65, RI = 0.88; SOM 1), and reveal three major clades (Fig. 1): extant Echimyidae (node A), as well as some of the extinct genera traditionally assigned to this family (Table 1); Octodontinae and Ctenomyinae (node J), plus a range of genera previously included in other taxa (Fig. 1, Table 1); and †Spaniomys and Abrocomidae (node T).

Node A (Echimyidae) has low support, but no character conflict (Fig. 1), and is diagnosed by (i) the lacrimal foramen opening into the maxilla (character state 10-1; Fig. 2); (ii) the presence of a continuous rim (without a suture) formed by the maxilla around the foramen into the lacrimal canal (character state 11-1); (iii) the lateral process of the supraoccipital extending ventrally below the level of the mastoid process (character state 38-1; Fig. 3); and (iv) the rotation of the distal portion of the paroccipital process, resulting in the posterolateral or posterior orientation of its external margin (character state 41-1). Within Echimyidae, there are two major subclades. The first of these (node I) includes the extant fossorial eumysopines Carterodon, Clyomys, and Euryzygomatomys, as well as the extinct †Theridomysops and †Dicolpomys, and is supported by details of the morphology of the zygomatic arch (character states 21-1, 24-1, and 25-1; not preserved in †Theridomysops and †Dicolpomys). The second subclade (node B) comprises the terrestrial spiny rats Thrichomys-†Pampamys-†Eumysops and Hoplomys + Proechimys- Trinomys, the semiaquatic coypu Myocastor, the spiny tree-rat Mesomys, and the arboreal echimyines-dactylomyines, all of which are united by the shape of their pterygoid fossae (character state 33-1).

Within the second echimyid subclade (node B), Myocastor is sister to the arboreal Mesomys + echimyines-dactylomyines (node C), as indicated by the presence of a thick lower margin of the posterior process of the squamosal (character state 34-1) and a short lower incisor (character state 49-1). The clade comprising Mesomys + echimyines- dactylomyines (node D) is supported by the position of the sphenopalatine fissure posteroventral to the lacrimal foramen (character state 18-1), the presence of a wide maxilla dorsal to this fissure (character state 19-1), and the dorsal or anterodorsal orientation of the anterior margin of the alisphenoid (character state 30-1). Within this clade, the adelphomyine †Maruchito clusters with the echimyines Echimys + Phyllomys owing to the presence of a transversely oriented posteroflexus crossing the entire occlusal surface of M1–M2 (node E; character state 57-1). A further group (node F) is formed by Capromys, †Deseadomys, dactylomyines, and the adelphomyine assemblage †Adelphomys-†Stichomys- †Eodelphomys-†Paradelphomys-†Xylechimys (node G), and supported by the subvertical orientation of the lateral crest of the mandible (character state 48-1; Fig. 4). In two of the three most parsimonious trees (SOM 1), the extant hutia Capromys clusters with the dactylomyines Dactylomys + Kannabateomys, based on the close juxtaposition of the margins of the maxillary fossae (character state 4-1) and the lateral margins of the incisive foramina (character state 5-1).

Fig. 1.

Strict consensus of the three most parsimonious trees (139 steps; CI = 0.65, RI = 0.88) resulting from the phylogenetic analysis. Numbers above and below branches represent absolute and relative Bremer support, respectively. A to T represent nodes discussed in the text. Black background indicates major taxa and their corresponding nodes as recognised in this study (cf. Table 1). Dashed lines indicate extinct taxa.

f01_757.jpg

Table 1.

Taxonomic arrangements for studied taxa. Comments: 1, tentatively included; 2, including †Sciamys; 3, including Pipanacoctomys; 4, invalid synonym of Thrichomys; 5, including †Abrocoma antiqua; 6, including Tympanoctomys; 7, including Aconaemys.

t01_757.gif

Fig. 2.

Morphology of the orbital region of octodontoids in lateral view. A. Octomys mimax Thomas, 1920, Recent, IMCN-CM 24. B. Thrichomys sp., Recent, MMP 150-USB542. C. Abrocoma sp., Recent, MLP 2038. D. †Spaniomys sp. from the Santa Cruz Formation, late Early Miocene, southern Argentina, MACN-A 4184. Dotted line in photographs shows the orientation of the nasolacrimal canal. Scale bars 5 mm.

f02_757.jpg

Node J (Octodontidae) is poorly supported and based on (i) the paroccipital process terminating dorsal to the level of the ventral portion of the auditory bulla (character state 39-1; Fig. 3); (ii) a ventrally deflected origin of the masseteric crest of the mandible (character state 45-1; Fig. 4); (iii) the formation of a lobe by the anterior lophids of Dp4 (character state 60-1); and (iv) the presence of a reduced metalophulid II on m1-2 (character state 68-1; Fig. 5). Within this broader clade, the grouping of †Protadelphomys-†Willidewu-† Sallamys, †Chasicomys and †Chasichimys with traditional ctenomyines (node K) is supported by the absence of the mesolophid on m1-2 (character state 67-1; Fig. 5) and the mastoid bulla terminating posterior to the level of the root of the paroccipital process (character state 44-1; Fig. 6). †Chasichimys shares with traditional ctenomyines a high basal portion (lacking flexids) of the molar crown (node M; character state 70-1). The latter, including the extant tuco-tuco Ctenomys (node N), are united by molars with a simplified occlusal morphology (character states 59-3 and 69-2).

Next to Ctenomyinae, †Caviocricetus + †Acarechimys- †Neophanomys, the †Acaremys group and traditional octodontines form a clade (node O) supported by the connection of the labial end of the mesolophule to the medial wall of the metacone on M1 (character state 54-1, not comparable for traditional octodontines; Fig. 7). The †Acaremys group is sister to traditional octodontines, and shares with them figure- eight shaped adult m1-2 (node P; character state 69-1). The traditional octodontines (node Q) includes fossorial and subterranean genera, and are supported by the presence of a premaxillary septum with divergent posterior ends (character state 1-1). Within this clade, Octodontomys + †Pseudoplataeomys innominatus forms the sister group to the remaining genera. The desert-specialists Octomys + †Abalosia-Tympanoctomys are united in a strongly supported clade (node S) diagnosed by the presence of an apophysis on the posterior portion of the maxilla posterior to the M3 alveolus (character state 27-2), a narrow posterior process of the squamosal reaching the posterior margin of epitympanic recess (character state 36-1), and the orientation of both the tip of the lateral process of the supraoccipital (character state 37-1) and the root of the paroccipital process (character state 43-1). The mesic-adapted Octodon + Spalacopus + Aconaemys-†Pithanotomys share the presence of an alisphenoid contacting the maxilla anteriorly (node R; character state 29-1). †Pithanotomys is sister to the rock rat Aconaemys, as indicated by a dorsally projecting jugal in the antorbital zygomatic bar (character state 20-1), and a pointed anterior face of Dp4 with well-defined secondary folds (character-state 58-1).

Fig. 3.

Lateral (A1, B1) and posterolateral (A2, B2) views of the temporal region of Recent octodontoids. A. Octomys mimax Thomas, 1920, IMCN-CM 24. B. Proechimys poliopus Osgood, 1914, MLP 22.II.00.7. Dotted line indicates the margin of the supraoccipital. Scale bars 5 mm.

f03_757.jpg

Finally, node T (†Spaniomys + Abrocomidae) is well-supported, and based on the contribution of the lacrimal to the posterior border of the foramen into the nasolacrimal canal (character state 12-1), a subhorizontally oriented nasolacrimal canal (character state 13-1; Fig. 2), a shortened lower incisor (character state 49-1), and anteriorly concave lophids on m1-3 (character state 63-1). The extinct abrocomid †Abrocoma antiqua is sister to the extant chinchilla rats Abrocoma + Cuscomys, with which it shares an extremely short lower incisor (character state 49-2), as well as a pointed anterior face of Dp4 with weak or no secondary folds (character state 58-3).

Divergence times.—The overall fit of our phylogenetic hypotheses with the stratigraphic record is low (MSM* score of 0.07 for the three optimal trees), implying the existence of substantial ghost lineages. We obtained similarly low values for echimyids (MSM*: 0.11-0.12), in which most extant lineages are inferred to originate in the Oligocene but are only recorded from the Late Miocene or Pliocene onwards. By contrast, our phylogenetic hypothesis for octodontids fits the stratigraphic record somewhat better (MSM*: 0.23). The MSM* scores for all trees are statistically significant (P = 0.001).

Estimated divergence times suggest Late Oligocene origins for the main octodontoid clades (Fig. 8; SOM 3). The potential antiquity of †Eodelphomys (Late Eocene?—Early Oligocene?) implies that some of these splits may have occurred even earlier; however, the age of this taxon is disputed (Frailey and Campbell 2004; Shockey et al. 2004) and hence was not taken into account here. Our results place the initial divergence of abrocomids (recorded since the Early Miocene), crown echimyids, and the main octodontid lineages in the Late Oligocene. By contrast, branching within modern (i.e., euhypsodont) octodontines and ctenomyines occurred at different times during the Late Miocene, with octodontines starting to diversify slightly earlier. The desert-adapted and mesic-adapted octodontines likely arose later during the Pliocene (Fig. 8). Finally, also during the Pliocene, the most recent branching event within ctenomyines gave rise to the extant genus Ctenomys.

Discussion

Phylogenetic relationships.—The clade defined by node A corresponds to extant Echimyidae (Galewski et al. 2005; Woods and Kilpatrick 2005), but excludes some of the Late Oligocene-Middle Miocene taxa referred to this family in previous studies (Fig. 1, Table 1). The relationships among echimyids are mostly, but not strictly, in agreement with previous morphological (Emmons 2005; Olivares et al. 2012; Candela and Rasia 2012) and molecular phylogenies (Galewski et al. 2005; Upham and Patterson 2012; Fabre et al. 2013). In particular, our results resemble earlier studies in recovering a major subclade comprising fossorial echimyids (Fig. 1: node I) in addition to a further clade comprising terrestrial, semiaquatic, and arboreal forms (Fig. 1: node B, Fig. 8; see also Galewski et al. 2005: fig. 3; Upham and Patterson 2012: fig. 3; Fabre et al. 2013: fig. 1). Likewise, our results do not support the monophyly of most of the traditionally recognised subfamilies.

Fig. 4.

Lateral view of the mandible of octodontoids. A. Kannabateomys amblyonyx (Wagner, 1845), Recent, MACN-Ma 15457. B. Tympanoctomys barrerae (Lawrence, 1941), Recent, MLP 2050. C. †Protadelphomys sp. from the Sarmiento Formation, Early Miocene, southern Argentina, CNP Pv 89-21a. Dotted line shows the anterior lower margin of the masseteric crest. Scale bars 5 mm.

f04_757.jpg

Fig. 5.

Occlusal morphology of the left lower molars (right inverted in A–C) of octodontoids. A. m2 of †Caviocricetus lucasi Vucetich and Verzi, 1996 from the Sarmiento Formation, Early Miocene, southern Argentina, MPEF 5076. B. m1 of †Willidewu esteparius Vucetich and Verzi, 1991 from the Sarmiento Formation, Early Miocene, southern Argentina, MPEF 5034. C. m3 of †Acaremys group from the Santa Cruz Formation, late Early Miocene, southern Argentina, MLP 15-XII-13-151. D. m1 of †Acaremys group from the Santa Cruz Formation, late Early Miocene, southern Argentina, MLP 15-393. Scale bars 2.5 mm.

f05_757.jpg

The clade comprising the fossorial echimyids, reco gnised by Emmons (2005) as the tribe Euryzygomatomyini, radiated in the Pampas of southern South America during the Late Miocene—Pliocene, alongside the terrestrial †Pampamys and †Eumysops (Fig. 8; Reig 1989; Olivares et al. 2012). Verzi et al. (1995) suggested all of these southern echimyids to form a clade, based on their shared possession of molars with a simplified occlusal morphology. However, this notion is contradicted by the present results (see also Olivares et al. 2012).

The position of Myocastor within Echimyidae agrees with the conclusions of several previous studies (Table 1; Emmons 2005; Galewski et al. 2005; Upham and Patterson 2012; Fabre et al. 2013). The precise relationships of this genus remain controversial (cf. Fig. 1; Candela and Rasia 2012; Upham and Patterson 2012), but it is worth noting that forcing Myocastor into a basal position as sister to all of the remaining echimyids (as in Candela and Rasia 2012) would require four extra steps.

The monophyly of the Late Oligocene—Early Miocene Patagonian subfamily †Adelphomyinae in its original sense (Patterson and Pascual 1968), is only partially supported by our results (Fig. 1: node G). In addition, the clade recovered here is not consistent with the more inclusive definition of the subfamily provided by Vucetich et al. (2010a: 215; Table 1). †Maruchito, previously included in the Adelphomyinae by these authors, is here grouped with extant echimyines, as proposed by Candela and Rasia (2012; see also Emmons and Vucetich 1998).

The extant Capromys is currently included in a family of its own (Capromyidae), which represents an intra-Caribbean radiation (Woods and Kilpatrick 2005). However, consistent with our results, recent molecular analyses found Capromys to be nested within Echimyidae (Upham and Patterson 2012: fig. 4; Fabre et al. 2013). In addition, a sister relationship between Capromys and dactylomyines, as indicated in two of our three most parsimonious trees (SOM 1), was previously proposed by Reig (1986: 409). Nevertheless, wider sampling of the Caribbean radiation of octodontoids is necessary to settle the question of the phylogenetic affinities of capromyids (Table 1; Upham and Patterson 2012).

Rather than elevating Ctenomyinae and Octodontinae to family level, we prefer to retain the name Octodontidae for node J (Patterson and Wood 1982: 523; Verzi 2001, and literature therein) to avoid the creation of a new higher taxon. According to our results, Ctenomyinae (Fig. 1: node K) comprises both traditional (euhypsodont) ctenomyines and genera that were previously classified as Echimyidae or as Octodontoidea with uncertain affinities (Table 1). With the single exception of the Late Miocene †Chasichimys (Verzi 1999), these genera are here assigned to the ctenomyines for the first time. Although the novel, unorthodox position of †Sallamys, †Protadelphomys, and †Willidewu as basal ctenomyines is poorly supported (Fig. 1), forcing these genera into a more basal position, either as sister to echimyids + octodontids or as sister to echimyids only, would require 5 additional steps.

Fig. 6.

Ventral view of the basicranial region of octodontids. A. Octodon sp., Recent, MLP 12.VII.88.2. B. Ctenomys maulinus Philippi, 1872, Recent, MLP 1.X.01.4. C. †Protadelphomys latus Ameghino, 1902 from the Sarmiento Formation, Early Miocene, southern Argentina, MPEF 5006. Scale bars 5 mm.

f06_757.jpg

Chasichimys has been suggested to have given rise to eumysopine echimyids (Table 1; Pascual 1967). Its position as sister to the euhypsodont ctenomyines in our phylogeny is poorly supported (Fig. 1), but placing it as the most basal octodontid or echimyid would require an additional 3 or 6 steps, respectively. The presumably related †Chasicomys was originally described as an octodontid (Pascual 1967) linked to modern octodontines (Reig 1989: 263). Given the low support for its inclusion in Ctenomyinae (Fig. 1: node L) and the fact that only one extra step is needed to move †Chasicomys to the base of Octodontidae, the position of this genus should thus be regarded as particularly tentative.

We consider the clade defined by node O to represent Octodontinae. Similar to ctenomyines, this clade groups traditional (euhypsodont) octodontines with genera here related to them for the first time (except for the †Acaremys group; Table 1). The interpretation of †Caviocricetus + †Acarechimys-†Neophanomys as stem Octodontinae is weakly supported (Fig. 1); however, forcing this clade into a basal position with respect to either Octodontidae + Echimyidae (as suggested by Vucetich et al. 2010a), or Echimyidae (as suggested by Patterson and Wood 1982), would each require 4 extra steps.

The topology of crown Octodontinae (node Q) is consistent with molecular phylogenies, except for the position of Octodontomys, which is usually placed closer to the mesic- adapted clade comprising Octodon + Spalacopus + Aconaemys-† Pithanotomys (Gallardo and Kirsch 2001; Honeycutt et al. 2003; Opazo 2005; Upham and Patterson 2012). Within the latter (Fig. 8), a close relationship between the extinct Pliocene †Pithanotomys and the extant Aconaemys was previously proposed by Reig (1986, 1989).

Spaniomys has previously been interpreted as either a myocastorine or adelphomyine echimyid (Table 1) (Ameghino 1889; Scott 1905). Its relationship with abrocomids, as reported here, might represent the first evidence of the pre-Late Miocene history of this family.

Delimitation and dating of higher taxa.—Three successive stages can be recognised in the evolutionary history of any given clade, referred to as t1, t2, and t3 by Hennig (1965: fig. 4): (i) its origin, i.e., divergence from its sister clade; (ii) acquisition of the apomorphies that characterise its extant members (modernisation); and (iii) the origin of the last common ancestor of the living representatives (Fig. 9). Recognition of these stages in the fossil record is essential in achieving an adequate understanding of the evolutionary history of a clade.

Octodontines, ctenomyines, and abrocomids have traditionally been recognised based on their hypsodont (especially euhypsodont) molars bearing simplified occlusal surfaces, which characterise their Late Miocene to Recent representatives (Fig. 8). The interpretation of the first record of these derived morphologies as an indicator of clade origin has yielded ages younger than 10 Ma for each of these taxa (e.g., Reig 1989; Verzi 1999; Vucetich et al. 1999). Including stem representatives provides an alternative way of defining and dating clades (Fig. 9). Although more unstable, as stem members are often poorly preserved and/or share few apomorphies with their corresponding crown-group (Briggs and Fortey 2005), this definition has the advantage of taking into account the deep history of a lineage, while at the same time avoiding paraphyletic taxa and the need for new names (Patterson 1993a, b; Donoghue 2005).

Fig. 7.

Occlusal morphology of the right upper molars (left inverted in B) of octodontids. A. DP4-M1 of †Caviocricetus lucasi Vucetich and Verzi, 1996 from the Sarmiento Formation, Early Miocene, southern Argentina, MPEF 505. B. M1 of †Acaremys group from the Santa Cruz Formation, late Early Miocene, southern Argentina, MLP 15-197. Scale bars 1 mm.

f07_757.jpg

Fig. 8.

Temporal ranges and divergence times of octodontoids mapped on to the strict consensus tree. Occlusal figures of the left m1 or m2 are illustrated next to the corresponding genus (when two figures are presented, the one to the right is ontogenetically more derived). Light gray background, clades adapted to xerophytic forest or open environments, first recorded during the Late Miocene global cooling and drying event; darker gray background, the desert-adapted octodontine clade first recorded during the Late Pliocene (ca. 2.5 Ma) global cooling and drying pulse; H, the modernisation stage represented by the acquisition of euhypsodont molars (black occlusal figures); asterisks, crown-groups. Timescale after Gradstein et al. (2008); isotopic curve after Zachos et al. (2008); palaeoclimatic events after Vrba et al. (1995), Verzi and Quintana (2005), Zachos et al. (2008), and Arakaki et al. (2011).

f08_757.jpg

Table 2.

Comparison of estimated ages of origin of total-groups (t1) and crown-groups (t3) with that of the modernisation stage (morphological differentiation, t2) of the studied taxa (see Fig. 8). Values from this study are fossil-based estimates and represent minimum ages. Comments: 1, original chronology updated on the basis of current available information (SOM 3); 2, data as averaged in Honeycutt (2009); 3, age estimation based on partial sampling; 4, calculated using 55 million years as maximum age for Caviomorpha. See references for calibration points and molecular data.

t02_757.gif

Following this line of reasoning, our reinterpretation of the phylogeny of octodontoids suggests that all of its major subclades are at least Late Oligocene in age (Fig. 8, Table 2), with the superfamily as a whole potentially being even older (Frailey and Campbell 2004; Vucetich et al. 2010b; Antoine et al. 2012). Both the times of origin of the major clades and the initial diversification of their respective crown groups are at least partially consistent with molecular age estimates (Table 2). In the case of Abrocomidae, Ctenomyinae and Octodontinae, our proposed dates exceed previous fossil-based estimates (e.g., Reig 1989; Vucetich et al. 1999; Cook et al. 2000; Vucetich and Kramarz 2003). This is largely a result of our different interpretation of the higher taxa, which regards modern (i.e., euhypsodont) abrocomids, ctenomyines and octodontines as stages of differentiation characterised by the acquisition of the morphology defining the extant species (t2 of Hennig 1965: fig. 4; apomorphy-based clades sensu De Queiroz and Gauthier 1990, 1992; see also Sereno 2005). These stages are decoupled from the divergences that originally separated these families and subfamilies (Fig. 9; Steiper and Young 2008).

A stage of modernisation (t2) distinct from the origin of the clade (t1) is not recognisable within Echimyidae or any of the subordinate clades of this family with an early fossil record. Echimyids are morphologically conservative, especially with regards to their dental morphology (Fabre et al. 2013) and, with the exception of the Antillean capromyids (if these are indeed echimyids), never acquired euhypsodonty (Fig. 8). This likely led previous authors to include early fossil octodontoids with lophate, rooted molars within echimyids (see above and Table 1), even though this morphology may in fact represent the plesiomorphic condition for Octodontoidea as a whole.

Morphological differentiation and Cenozoic climatic changes.—The diverging evolutionary patterns of octodontids and abrocomids on the one hand and echimyids on the other may reflect different responses to Cenozoic climatic change. In the palaeoclimatic history of the Cenozoic, arid climates and open habitats are clearly derived, and arose more recently than environmental conditions similar to those of present-day tropical forests (e.g., Janis 1993; Partridge et al. 1995; Janis et al. 2000; Zachos et al. 2001). Echimyids responded to Cenozoic environmental changes (including diastrophic events) mostly by tracking their original habitats. Thus, their extant representatives primarily inhabit Amazonian and Atlantic forests (Fabre et al. 2013), with only a few species having colonised more open areas (Pascual 1967; Hoffstetter 1986; Verzi et al. 1994; Olivares et al. 2012; Upham and Patterson 2012).

The fossil record of echimyids from southern South America is an impoverished, marginal sample of the astonishing diversity achieved by this group in the tropical and subtropical areas of northern South America. Fossils linked to the extant arboreal species are present in Patagonia only until the Middle Miocene, coinciding with the persistence of forests with tropical elements at this latitude (Fig. 8; Palazzesi and Barreda 2007). By contrast, all of the known Late Miocene fossils are related to the few groups adapted to open areas (Fig. 8; Verzi et al. 1994; Olivares et al. 2012). Local studies illustrate the Late Miocene decline of echimyids in southern South America (Verzi et al. 2011), which led to their current absence in the area. It is to be expected that taxa which respond to environmental change by tracking their original habitats through local extinctions and distribution drift (Vrba 1992) will be morphologically more conservative than those that adapt to their new environment, irrespective of speciation rates (Verzi 2002). This might explain why the origin and differentiation of echimyids are not recognisable as decoupled events, and furthermore may account for the morphological redundancy of some highly species-rich, unexpectedly ancient clades of living echimyids, such as Proechimys-Trinomys (Verzi 2002; see Lara et al. 1996: 410; Da Silva and Patton 1998; Fabre et al. 2013).

Fig. 9.

Graphical representation of stem-, crown-, and total-groups, and related times of origin (t1 and t3) and morphological differentiation (t2) (modified from Hennig 1965: fig. 4 and Sereno 2005: fig. 1). Note that in clade A (exemplified by Ctenomyinae and Abrocomidae), t2 and t3 are decoupled while in clade B (exemplified by Octodontinae) they are coincident.

f09_757.jpg

By contrast, abrocomids, octodontines, and ctenomyines show stages of differentiation that are distinctly decoupled from their respective origins (Fig. 9, Table 2), with the hierarchy of these stages following that of concurrent palaeoenvironmental changes. A derived, hypsodont dental morphology appeared in all of these groups by the Late Miocene, likely in response to increased aridity (Vucetich and Verzi 1999; Verzi 1999; Verzi et al. 2004) and the subsequent rise of extensive open biomes in southern South America (Fig. 8; Pascual and Ortiz Jaureguizar 1990; Rabassa et al. 2005; Palazzesi and Barreda 2007; Le Roux 2012). Available data show the global nature of this Late Miocene cooling and drying, resulting in the expansion of open environments worldwide (e.g., Janis 1993; Vrba et al. 1995; Zachos et al. 2001, 2008; Arakaki et al. 2011). The radiation of hypsodont species and the extinction of lineages with primitive molars (Verzi et al. 2011: fig. 8) marked the beginning of the stage of modernisation of these clades. The subsequent appearance of desert specialists among hypsodont octodontines coincided with a profound global Late Pliocene cooling and drying event (Fig. 8; Verzi 2001; Verzi and Quintana 2005 and references therein). Together, these observations suggest that the extratropics may have acted as a cradle of evolutionary novelties through fixation of new adaptations to newly emerging open habitats (Verzi 2002).

The importance of recognising morphological differentiation in the fossil record.—Hennig (1965: 114) noted that the delimitation of the stage of morphological differentiation in the history of a clade (t2) depends on subjective criteria concerning the interpretation of the emergence of particular “types” or “Baupläne”. Indeed, whereas the origin of a clade (t1) and of its crown-group (t3) can be determined using both molecular and palaeontological evidence, the time of morphological differentiation (i.e., modernisation, t2) can only be informed by fossils, and is constrained by biases affecting the minimum ages provided by the fossil record (Benton and Donoghue 2007). Nevertheless, as demonstrated here, such differentiation stages can yield important evolutionary information regarding environmentally driven changes in morphology that occurred before the emergence of the crown group. This highlights the unique contribution of fossils to the appreciation of the true shape of trees (Helgen 2011) and our understanding of macroevolutionary patterns.

Acknowledgements

We thank Federico Anaya (Universidad Autónoma Tomás Frías, Potosí, Peru), Eileen Lacey, James Patton (both Museum of Vertebrate Zoology, Berkeley, USA), João de Oliveira (Museu Nacional, Universidade Federal do Rio de Janeiro, Argentina), and Guiomar Vucetich (Museo de La Plata, Argentina) for actively facilitating access to materials; Diego Pol and María Pérez (both Museo Paleontológico Egidio Feruglio, Trelew, Argentina) for their valuable assistance with methodological aspects of the cladistic analysis; Diego Pol for providing the TNT script for MSM*, and for his valuable comments on an earlier version of this manuscript; James Zachos (Earth and Planetary Sciences Department, University of California Santa Cruz, USA) for allowing use of his original isotope dataset; and Fernando Ballejo (Universidad Nacional de La Plata, Argentina) for drawing the illustrations in Fig. 5. We are especially grateful to Philip Cox (Hull York Medical School, University of York, UK) and Laurent Marivaux (Laboratoire de Paléontologie, Institut des Sciences de l´Évolution de Montpellier, Université Montpellier 2, France) for their critical reviews and Felix Marx (National Museum of Nature and Science, Tsukuba, Japan) for his thorough editorial work, all of which helped to improve this paper. Access to the software TNT was available thanks to the sponsorship of the Willi Hennig Society. This paper is a contribution to projects ANPCyT PICT 2007-01744 and CONICET PIP 0270.

References

1.

F. Ameghino 1889. Contribución al conocimiento de los mamíferos fósiles de la República Argentina. Actas Academia Nacional de Ciencias, Córdoba 6: 1–1027. Google Scholar

2.

F. Ameghino 1902. Première contribution à la connaissance de la faune mammalogique des couches à Colpodon. Boletín de la Academia Nacional de Ciencias en Córdoba 17: 71–138. Google Scholar

3.

P.-O Antoine , L. Marivaux , D.A. Croft , G. Billet , M. Ganerød , C. Jaramillo , T. Martin , M.J. Orliac , J. Tejada , A.J. Altamirano , F. Duranthon , G. Fanjat , S. Rousse , and R.S. Gismondi 2012. Middle Eocene rodents from Peruvian Amazonia reveal the pattern and timing of caviomorph origins and biogeography. Proceedings of the Royal Society B 279: 1319–1326. Google Scholar

4.

M. Arakaki , P.-A. Christin , R. Nyffeler , A. Lendel , U. Eggli , R.M. Ogburn , E. Spriggs , M.J. Moore , and E.J. Edwards 2011. Contemporaneous and recent radiations of the world's major succulent plant lineages. Proceedings of the National Academy of Sciences 108: 8379–8384. Google Scholar

5.

M.J. Benton and P.C.J. Donoghue 2007. Paleontological evidence to date the tree of life. Molecular Biology and Evolution 24: 26–53. Google Scholar

6.

S. Bertelli and N.P. Giannini 2005. A phylogeny of extant penguins (Aves: Sphenisciformes) combining morphology and mitochondrial sequences. Cladistics 21: 209–239. Google Scholar

7.

K. Bremer 1994. Branch support and tree stability. Cladistics 10: 295–304. Google Scholar

8.

D.E.G. Briggs and R.A. Fortey 2005. Wonderful strife: systematics, stem groups, and the phylogenetic signal of the Cambrian radiation. Paleobiology 31: 94–112. Google Scholar

9.

A.M. Candela and L. Rasia 2012. Tooth morphology of Echimyidae (Rodentia, Caviomorpha): homology assessments, fossils, and evolution. Zoological Journal of the Linnean Society 164: 451–480. Google Scholar

10.

G.A.S. Carvalho and O.L. Salles 2004. Relationships among extant and fossil echimyids (Rodentia: Hystricognathi). Zoological Journal of the Linnean Society 142: 445–477. Google Scholar

11.

J.A. Coddington and N. Scharff 1994. Problems with zero-length branches. Cladistics 10: 415–423. Google Scholar

12.

J.A. Cook , E.P. Lessa , and E.A. Hadly 2000. Paleontology, phylogenetic patterns, and macroevolutionary processes in subterranean rodents. In : A.E. Lacey , J.L. Patton , and G.N. Cameron (eds.), Life Underground. The Biology of Subterranean Rodents , 332–369. The University of Chicago Press, Chicago. Google Scholar

13.

M.N. Da Silva and J.L. Patton 1998. Molecular phylogeography and the evolution and conservation of Amazonian mammals. Molecular Ecology 7: 475–486. Google Scholar

14.

K. De Queiroz and J. Gauthier 1990. Phylogeny as a central principle in taxonomy: phylogenetics definition of taxon names. Systematic Zoology 39: 307–322. Google Scholar

15.

K. De Queiroz and J. Gauthier 1992. Phylogenetic taxonomy. Annual Review of Ecology and Systematics 23: 449–480. Google Scholar

16.

P.C.J. Donoghue 2005. Saving the stem group—a contradiction in terms? Paleobiology 31: 553–558. Google Scholar

17.

J.F. Eisenberg and K.H. Redford 1999. Mammals of the Neotropics. The Central Neotropics: Ecuador, Peru, Bolivia, Brazil . 624 pp. University of Chicago Press, Chicago. Google Scholar

18.

L.H. Emmons 2005. A revision of the genera of arboreal Echimyidae (Rodentia: Echimyidae, Echimyinae), with descriptions of two new genera. In : E.A. Lacey and P. Myers (eds.), Mammalian Diversification: from Chromosomes to Phylogeography (a Celebration of the Career of James L. Patton), 247–309. University of California Press, Berkeley. Google Scholar

19.

L.H. Emmons and F. Feer 1999. Neotropical Rainforest Mammals: a Field Guide. 281 pp. University of Chicago Press, Chicago. Google Scholar

20.

L.H. Emmons and M.G. Vucetich 1998. The Identity of Winge's Lasiuromys villosus and the description of a new genus of echimyid rodent (Rodentia: Echimyidae). American Museum Novitates 3223: 1–12. Google Scholar

21.

P.-H. Fabre , T. Galewski , M. Tilak , and E.J.P. Douzery 2013. Diversification of South American spiny rats (Echimyidae): a multigene phylogenetic approach. Zoologica Scripta 42 (2): 117–134. Google Scholar

22.

C.D. Frailey and K.E. Campbell Jr. 2004. Paleogene rodents from Amazonian Peru: the Santa Rosa Local Fauna. In : K.E. Campbell Jr.The Paleogene Mammalian Fauna of Santa Rosa, Amazonian Peru. Natural History Museum of Los Angeles County, Science Series 40: 71–130. Google Scholar

23.

T. Galewski , J.F. Mauffrey , Y.L.R. Leite , J.L. Patton , and E.J.P. Douzery 2005. Ecomorphological diversification among South American spiny rats (Rodentia: Echimyidae): a phylogenetic and chronological approach. Molecular Phylogenetics and Evolution 34: 601–615. Google Scholar

24.

M.H. Gallardo and J.A.W. Kirsch 2001. Molecular relationships among Octodontidae (Mammalia: Rodentia: Caviomorpha). Journal of Mammalian Evolution 8: 73–89. Google Scholar

25.

P. Goloboff and J. Farris 2001. Methods for quick consensus estimation. Cladistics 17: 26–34. Google Scholar

26.

P.A. Goloboff , J.S. Farris , and K. Nixon 2008a. TNT: Tree Analysis Using New Technology, Version 1.1. Available at:  http://www.zmuc.dk/public/phylogeny/tntGoogle Scholar

27.

P.A. Goloboff , J S. Farris , and K. Nixon 2008b. TNT, a free program for phylogenetic analysis. Cladistics 24: 774–786. Google Scholar

28.

F.M. Gradstein , J.G. Ogg , and M. Van Kranendonk 2008. On the Geologic Time Scale 2008. Newsletters on Stratigraphy 43: 5–13. Google Scholar

29.

K.M. Helgen 2011. The Mammal Family Tree. Science 334: 458–459. Google Scholar

30.

W. Hennig 1965. Phylogenetic systematics. Annual Review of Entomology 10: 97–116. Google Scholar

31.

R. Hoffstetter 1986. High Andean mammalian faunas during the Plio-Pleistocene. In : F. Vuilleumier and M. Monasterio (eds.), High Altitude Tropical Biogeography , 218–245. Oxford University Press, Oxford. Google Scholar

32.

R.L. Honeycutt 2009. Rodents (Rodentia). In : S.B. Hedges and S. Kumar (eds.), The Timetree of Life , 490–494. Oxford University Press, New York. Google Scholar

33.

R.L. Honeycutt , D.L. Rowe , and M.H. Gallardo 2003. Molecular systematics of the South American caviomorph rodents: relationships among species and genera in the family Octodontidae. Molecular Phylogenetics and Evolution 26: 476–489. Google Scholar

34.

C.M. Janis 1993. Tertiary mammal evolution in the context of changing climates, vegetation, and tectonic events. Annual Review of Ecology and Systematics 24: 467–500. Google Scholar

35.

C.M. Janis , J. Damuth , and J.M. Theodor 2000. Miocene ungulates and terrestrial primary productivity: Where have all the browsers gone? Proceedings of the National Academy of Sciences 97: 7899–7904. Google Scholar

36.

M.C. Lara , J.L. Patton , and M.N. Da Silva 1996. The simultaneous diversification of South American echimyid rodents (Hystricognathi) based on complete cytochrome b sequences. Molecular Phylogenetics and Evolution 5: 403–413. Google Scholar

37.

B. Lawrence 1941. A new species of Octomys from Argentina. Proceedings of the New England Zoölogical Club 18: 43–46. Google Scholar

38.

J.P. Le Roux 2012. A review of Tertiary climate changes in southern South America and the Antarctic Peninsula. Part 2: continental conditions. Sedimentary Geology 247–248: 21–38. Google Scholar

39.

M.C. McKenna and S.K. Bell 1997. Classification of Mammals above the Species Level. 631 pp. Columbia University Press, New York. Google Scholar

40.

A.I. Olivares , D.H. Verzi , M.G. Vucetich , and C.I. Montalvo 2012. Phylogenetic affinities of the late Miocene echimyid †Pampamys and the age of Thrichomys (Rodentia, Hystricognathi). Journal of Mammalogy 92: 76–86. Google Scholar

41.

J.C. Opazo 2005. A molecular timescale for caviomorph rodents (Mammalia, Hystricognathi). Molecular Phylogenetics and Evolution 37: 932–937. Google Scholar

42.

W.H. Osgood 1914. Four new mammals from Venezuela. Publication of Field Museum of Natural History, Zoological Series 10: 135–141. Google Scholar

43.

L. Palazzesi and V. Barreda 2007. Major vegetation trends in the Tertiary of Patagonia (Argentina): a qualitative paleoclimatic approach based on palynological evidence. Flora 202: 328–337. Google Scholar

44.

T.C. Partridge , G.C. Bond , C.J.H. Hartnady , P.B. deMenocal , and W.F. Ruddiman 1995. Climatic effects of Late Neogene tectonism and volcanism. In : E.S. Vrba , G.H. Denton , T.C. Partridge , and L.H. Burckle (eds.), Palaeoclimate and Evolution with Emphasis on Human Origins , 8–23. Yale University Press, New Haven. Google Scholar

45.

R. Pascual 1967. Los roedores Octodontoidea (Caviomorpha) de la Formación Arroyo Chasicó (Plioceno inferior) de la Provincia de Buenos Aires. Revista del Museo de La Plata, Paleontología 5: 259–282. Google Scholar

46.

R. Pascual and E. Ortiz Jaureguizar 1990. Evolving climates and Mammal faunas in Cenozoic South America. Journal of Human Evolution 19: 23–60. Google Scholar

47.

B. Patterson and R. Pascual 1968. New echimyid rodents from the Oligocene of Patagonia, and a synopsis of the family. Museum of Comparative Zoology, Breviora 301: 1–14. Google Scholar

48.

B. Patterson and A.E. Wood 1982. Rodents from the Deseadan Oligocene of Bolivia and the relationships of the Caviomorpha. Bulletin of the Museum of Comparative Zoology 149: 371–543. Google Scholar

49.

C. Patterson 1993a. Bird or dinosaur? Nature 365: 21–22. Google Scholar

50.

C. Patterson 1993b. Naming names. Nature 366: 518. Google Scholar

51.

R.A. Philippi 1872. Drei neue Nager aus Chile. Zeitschrift für die Gesammten Naturwissenschaften 6: 442–447. Google Scholar

52.

D. Pol and M.A. Norell 2001. Comments on the Manhattan Stratigraphic Measure. Cladistics 17: 285–289. Google Scholar

53.

D. Pol , M.A. Norell , and M.E. Siddall 2004. Measures of stratigraphic fit to phylogeny and their sensitivity to tree size, tree shape, and scale. Cladistics 20: 64–75. Google Scholar

54.

J. Rabassa , A. Coronato , and M. Salemme 2005. Chronology of the Late Cenozoic Patagonian glaciations and their correlation with biostratigraphic units of the Pampean Region (Argentina). Journal of South American Earth Sciences 20: 81–103. Google Scholar

55.

K.H. Redford and J.F. Eisenberg 1992. Mammals of the Neotropics. The Southern Cone: Chile, Argentina, Uruguay, Paraguay . 430 pp. Uni versity of Chicago Press, Chicago. Google Scholar

56.

O.A. Reig 1986. Diversity patterns and differentiation of high Andean rodents. In : F. Vuilleumier and M. Monasterio (eds.), High Altitude Tropical Biogeography , 404–439. Oxford University Press, Oxford. Google Scholar

57.

O.A. Reig 1989. Karyotypic repatterning as one triggering factor in cases of explosive speciation. In : A. Fontdevila (ed.), Evolutionary Biology of Transient Unstable Populations , 246–289. Springer-Verlag, Berlin. Google Scholar

58.

C. Rovereto 1914. Los estratos araucanos y sus fósiles. Anales del Museo Nacional de Historia Natural de Buenos Aires 25: 1–247. Google Scholar

59.

D.L. Rowe , K.A. Dunn , R.M. Adkins and R.L. Honeycutt 2010. Mole cular clocks keep dispersal hypotheses afloat: evidence for trans-Atlantic rafting by rodents. Journal of Biogeography 37: 305–324. Google Scholar

60.

W.B. Scott 1905. Mammalia of the Santa Cruz beds, 3: Glires. In : W.B. Scott (ed.), Reports of the Princeton University Expeditions to Patagonia, 1896-1899, Palaeontology 5 , 384–490. The New Era Printing Company, Lancaster. Google Scholar

61.

P.C. Sereno 2005. The logical basis of phylogenetic taxonomy. Systematic Biology 54: 595–619. Google Scholar

62.

B.J. Shockey , R. Hitz , and M. Bond 2004. Paleogene notoungulates from the Amazon Basin of Peru. In : K.E. Campbell Jr. (ed.), The Paleogene Mammalian Fauna of Santa Rosa, Amazonian Peru. Natural History Museum of Los Angeles County, Science Series 40: 61–69. Google Scholar

63.

G.G. Simpson 1945. The principles of classification and a classification of mammals. Bulletin of the American Museum of Natural History 85: 1–350. Google Scholar

64.

M.E. Steiper and N.M. Young 2008. Timing primate evolution: lessons from the discordance between molecular and paleontological estimates. Evolutionary Anthropology 17: 179–188. Google Scholar

65.

O. Thomas 1920. On mammals from near Tinogasta, Catamarca. Annals and Magazine of Natural History, Series 9 6: 117–119. Google Scholar

66.

N.S. Upham and B.D. Patterson 2012. Diversification and biogeography of the Neotropical caviomorph lineage Octodontoidea (Rodentia: Hystricognathi). Molecular Phylogenetics and Evolution 63: 417–429. Google Scholar

67.

D.H. Verzi 1999. The dental evidence on the differentiation of the ctenomyine rodents (Caviomorpha, Octodontidae, Ctenomyinae). Acta Theriologica 44: 263–282. Google Scholar

68.

D.H. Verzi 2001. Phylogenetic position of Abalosia and the evolution of the extant Octodontinae (Rodentia, Caviomorpha, Octodontidae). Acta Theriologica 46: 243–268. Google Scholar

69.

D.H. Verzi 2002. Patrones de evolución morfológica en Ctenomyinae (Rodentia, Octodontidae). Mastozoología Neotropical 9: 309–328. Google Scholar

70.

D.H. Verzi and C.A. Quintana 2005. The caviomorph rodents from the San Andrés Formation, east-central Argentina, and global Late Pliocene climatic change. Palaeogeography, Palaeoclimatology, Palaeoecology 219: 303–320. Google Scholar

71.

D.H. Verzi , E.C. Vieytes , and C.I. Montalvo 2004. Dental evolution in Xenodontomys and first notice on secondary acquisition of radial enamel in rodents (Rodentia, Caviomorpha, Octodontidae). Geobios 37: 795–806. Google Scholar

72.

D.H. Verzi , E.C. Vieytes , and C.I. Montalvo 2011. Dental evolution in Neophanomys (Rodentia, Octodontidae) from the late Miocene of central Argentina. Geobios 44: 621–633. Google Scholar

73.

D.H. Verzi , M.G. Vucetich , and C.I. Montalvo 1994. Octodontid-like Echimyidae (Rodentia): an upper Miocene episode in the radiation of the family. Palaeovertebrata 23: 199–210. Google Scholar

74.

D.H. Verzi , M.G. Vucetich , and C.I. Montalvo 1995. Un nuevo Eumysopinae (Rodentia, Echimyidae) de Mioceno tardío de la Provincia de La Pampa y consideraciones sobre la historia de la subfamilia. Ameghiniana 32: 191–195. Google Scholar

75.

R.V. Vilela , T. Machado , K. Ventura , V. Fagundes , M.J. Silva , and Y. Yonenaga-Yassuda 2009. The taxonomic status of the endangered thinspined porcupine, Chaetomys subspinosus (Olfers, 1818), based on molecular and karyologic data. BMC Evolutionary Biology 9: 29. Google Scholar

76.

E.S. Vrba 1992. Mammals as a key to evolutionary theory. Journal of Mammalogy 73: 1–28. Google Scholar

77.

E.S. Vrba , G.H. Denton , T.C. Partridge , and L.H. Burckle (eds.). 1995. Paleoclimate and Evolution with Emphasis on Human Origins. 547 pp. Yale University Press, New Haven. Google Scholar

78.

M.G. Vucetich and A.G. Kramarz 2003. New Miocene rodents from Patagonia (Argentina) and their bearing on the early radiation of the octodontoids (Hystricognathi). Journal of Vertebrate Paleontology 23: 435–444. Google Scholar

79.

M.G. Vucetich and D.H. Verzi 1991. Un nuevo Echimyidae (Rodentia, Hystricognathi) de la Edad Colhuehuapense de Patagonia y consideraciones sobre la sistemática de la familia. Ameghiniana 28: 67–74. Google Scholar

80.

M.G. Vucetich and D.H. Verzi 1996. A peculiar octodontoid (Rodentia, Caviomorpha) with terraced molars from the lower Miocene of Patagonia (Argentina). Journal of Vertebrate Paleontology 16: 297–302. Google Scholar

81.

M.G. Vucetich and D.H. Verzi 1999. Changes in diversity and distribution of the caviomorph rodents during the Late Cenozoic in Southern South America. Quaternary of South America and Antarctic Peninsula 12: 207–223. Google Scholar

82.

M.G. Vucetich , A.G. Kramarz , and A.M. Candela 2010a. Colhuehuapian rodents from Gran Barranca and other Patagonian localities: the state of the art. In : R.H. Madden , A.A. Carlini , M.G. Vucetich , and R.F. Kay (eds.), The Paleontology of Gran Barranca: Evolution and Environmental Change through the Middle Cenozoic of Patagonia , 206–219. Cambridge University Press, New York. Google Scholar

83.

M.G. Vucetich , D.H. Verzi , and J.-L. Hartenberger , 1999. Review and analysis of the radiation of the South American Hystricognathi (Mammalia, Rodentia). Comptes Rendus de L'Academie des Sciences, Série IIa/Sciences de la Terre et des Planètes. Paléontologie 329: 763–769. Google Scholar

84.

M.G. Vucetich , E.C. Vieytes , M.E. Pérez , and A.A. Carlini 2010b. The rodents from La Cantera and the early evolution of caviomorphs in South America. In : R.H. Madden , A.A. Carlini , M.G. Vucetich , and R.F. Kay (eds.), The Paleontology of Gran Barranca: Evolution and Environmental Change through the Middle Cenozoic of Patagonia , 193–205. Cambridge University Press, New York. Google Scholar

85.

J.A. Wagner 1845. Diagnosen einiger neuen Arten von Nagern und Handflülern. Archiv für Naturgeschichte 11: 145–149. Google Scholar

86.

H. Winge 1941. The interrelationships of the mammalian genera. In : A.D.S. Jensen , R. Spräck , and H. Volsøe (eds.), Vol. 2, Rodentia, Carnivora, Primates (translated from Danish by Deichmann E, Allen GM), 1-376. CA Reitzels Forlag, Copenhagen. Google Scholar

87.

A.E. Wood 1955. A revised classification of the rodents. Journal of Mammalogy 36: 165–187. Google Scholar

88.

A.E. Wood and B. Patterson 1959. Rodents of the Deseadan Oligocene of Patagonia and the beginnings of South American rodent evolution. Bulletin of the Museum of Comparative Zoology 120: 279–428. Google Scholar

89.

C.A. Woods 1984. Hystricognath rodents. In : S. Anderson and J.K. Jones (eds.), Orders and Families of Recent Mammals of the World , 389–446. John Wiley & Sons, Inc., New York. Google Scholar

90.

C.A. Woods and C.W. Kilpatrick 2005. Infraorder Hystricognathi Brandt, 1855. In : D.E. Wilson and D.M. Reeder (eds.), Mammal Species of the World , 1538–1600. Johns Hopkins University Press, Baltimore. Google Scholar

91.

J.C. Zachos , G.R. Dickens , and R.E. Zeebe 2008. An early Cenozoic perspective on greenhouse warming and carbon-cycle dynamics. Nature 451: 279–283. Google Scholar

92.

J. Zachos , M. Pagani , L. Sloan , E. Thomas , and K. Billups 2001. Trends, rhythms, and aberrations in global climate 65 Ma to present. Science 292: 686–693. Google Scholar
© 2014 D.H. Verzi et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Diego H. Verzi, A. Itatí Olivares, and Cecilia C. Morgan "Phylogeny and Evolutionary Patterns of South American Octodontoid Rodents," Acta Palaeontologica Polonica 59(4), 757-769, (13 March 2013). https://doi.org/10.4202/app.2012.0135
Received: 9 November 2012; Accepted: 11 March 2013; Published: 13 March 2013
KEYWORDS
Cenozoic
divergence dates
evolution
Hystricognathi
Mammalia
Octodontoidea
phylogeny
Back to Top