Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2017 Jul 3;114(29):E5864–E5870. doi: 10.1073/pnas.1704632114

Phylogenomics reveals rapid, simultaneous diversification of three major clades of Gondwanan frogs at the Cretaceous–Paleogene boundary

Yan-Jie Feng a, David C Blackburn b, Dan Liang a, David M Hillis c, David B Wake d,1, David C Cannatella c,1, Peng Zhang a,1
PMCID: PMC5530686  PMID: 28673970

Significance

Frogs are the dominant component of semiaquatic vertebrate faunas. How frogs originated and diversified has long attracted the attention of evolutionary biologists. Here, we recover their evolutionary history by extensive sampling of genes and species and present a hypothesis for frog evolution. In contrast to prior conclusions that the major frog clades were established in the Mesozoic, we find that ∼88% of living frogs originated from three principal lineages that arose at the end of the Mesozoic, coincident with the Cretaceous–Paleogene (K–Pg) mass extinction event that decimated nonavian dinosaurs 66 Mya. The K–Pg extinction events played a pivotal role in shaping the current diversity and geographic distribution of modern frogs.

Keywords: amphibia, Anura, nuclear genes, phylogeny, divergence time

Abstract

Frogs (Anura) are one of the most diverse groups of vertebrates and comprise nearly 90% of living amphibian species. Their worldwide distribution and diverse biology make them well-suited for assessing fundamental questions in evolution, ecology, and conservation. However, despite their scientific importance, the evolutionary history and tempo of frog diversification remain poorly understood. By using a molecular dataset of unprecedented size, including 88-kb characters from 95 nuclear genes of 156 frog species, in conjunction with 20 fossil-based calibrations, our analyses result in the most strongly supported phylogeny of all major frog lineages and provide a timescale of frog evolution that suggests much younger divergence times than suggested by earlier studies. Unexpectedly, our divergence-time analyses show that three species-rich clades (Hyloidea, Microhylidae, and Natatanura), which together comprise ∼88% of extant anuran species, simultaneously underwent rapid diversification at the Cretaceous–Paleogene (K–Pg) boundary (KPB). Moreover, anuran families and subfamilies containing arboreal species originated near or after the KPB. These results suggest that the K–Pg mass extinction may have triggered explosive radiations of frogs by creating new ecological opportunities. This phylogeny also reveals relationships such as Microhylidae being sister to all other ranoid frogs and African continental lineages of Natatanura forming a clade that is sister to a clade of Eurasian, Indian, Melanesian, and Malagasy lineages. Biogeographical analyses suggest that the ancestral area of modern frogs was Africa, and their current distribution is largely associated with the breakup of Pangaea and subsequent Gondwanan fragmentation.


A robust, reliable phylogeny is essential to understand the role of macroevolutionary processes in generating biodiversity. However, resolution of evolutionary relationships among certain groups has been persistently difficult because of sparse genotypic and phenotypic data. Frogs (Anura) are one such example; they are one of the most diverse groups of tetrapods, and currently comprise 6,775 described species, 446 genera, and 55 families (1) that are well represented on all continents. They exhibit great adaptive diversity within a highly constrained phenotype estimated to be 200 My old. Evolutionary convergence in body form, life history, and behavioral traits is widespread in frogs, including forms reflecting different microhabitat use by arboreal, aquatic, and fossorial species. These features make frogs a challenging but fascinating model for addressing fundamental questions of morphological, developmental, and biogeographical evolution. However, despite intensive molecular phylogenetic studies (27), areas of uncertainty and disagreement persist among clades that are crucial for interpreting broad-scale macroevolutionary patterns. In addition, a general consensus on divergence times of the major anuran lineages is also lacking (7, 8).

The poor resolution for many nodes in anuran phylogeny is likely a result of the small number of molecular markers traditionally used for these analyses. Previous large-scale studies used 6 genes (∼4,700 nt) (4), 5 genes (∼3,800 nt) (5), 12 genes (6) with ∼12,000 nt of GenBank data (but with ∼80% missing data), and whole mitochondrial genomes (∼11,000 nt) (7). In the larger datasets (e.g., ref. 6), most data (>50%) are from the 12S and 16S mitochondrial ribosomal genes. The limited amount of data also causes a wide range of estimates of divergence times for many nodes in the tree. For example, age estimates for the last common ancestor of extant Neobatrachia, often referred to as “modern frogs” and containing 95% of extant anuran species, span ∼100 Mya (5, 711). Furthermore, divergences time estimates among the earliest neobatrachian clades, such as the Heleophrynidae, Myobatrachidae, Calyptocephalellidae, Nasikabatrachidae, and Sooglossidae, range from the Late Jurassic to early Cretaceous (∼150–100 Mya) and have wide CIs (5, 711). In addition to these species-poor groups of neobatrachians, there are two species-rich clades: Ranoidea (39% of extant anuran species, mostly Old World) and Hyloidea (54%; mostly New World). The estimated ages of each clade range from the Late Jurassic to the end of the Cretaceous, spanning ∼100 My, and relationships of family-level taxa within each clade remain poorly resolved.

In this study, we increased gene sampling by using a recently developed nuclear marker toolkit (12). Our new data include ∼88,000 nt of aligned sequences from 95 nuclear protein-coding genes covering 164 species (156 anuran species and 8 outgroups) from 44 of 55 frog families; to our knowledge, this is the largest source of new data for anuran phylogenetics. In addition, we enlarged this dataset to a total of 301 anuran species by incorporating previously published RAG1 and CXCR4 sequences so that all 55 extant frog families were included. Our goal was to propose a robust hypothesis of phylogenetic relationships and divergence times of the major lineages. Our results resolve previously intractable relationships, generate divergence times with narrow CIs, and provide perspectives on the evolutionary history and historical biogeography of frogs.

Results and Discussion

Data Characteristics.

We assembled a de novo 164-species dataset by using 95 nuclear genes (Table S1) and 88,302 nt from 156 frog species and 8 outgroups; this matrix is 89.6% complete. To increase coverage of anuran families, we added sequences of RAG1 and CXCR4 from GenBank of 145 additional anuran species. This 309-species dataset contains 88,386 nt and is 48.2% complete. The 164-species and 309-species matrices are available from the Dryad Digital Repository (doi:10.5061/dryad.12546).

Table S1.

Descriptive statistics for the 95 loci used in this study

No. Gene name Length, bp Variable sites Parsimony informative sites Relative composition variability GC of total mean, % GC of third position mean, % Substitution model (BIC)
1 ADNP 822 86 507 0.0445 47.6 45.1 GTR+I+G
2 ANKRD50 996 61 484 0.0261 42.1 32.2 GTR+I+G
3 ARID2 873 83 587 0.0342 44.0 32.8 HKY+I+G
4 ARSI 852 44 485 0.0431 44.2 41.9 GTR+I+G
5 B3GALT1 780 49 332 0.0402 40.1 41.9 HKY+I+G
6 BPTF 546 29 258 0.0980 54.4 70.1 SYM+I+G
7 CAND1 1,176 50 514 0.0238 42.6 35.3 GTR+I+G
8 CASR 750 39 346 0.0515 46.5 64.0 GTR+I+G
9 CELSR3 1,281 91 693 0.0682 48.5 52.6 GTR+I+G
10 CHST1 693 47 391 0.0417 46.5 47.6 GTR+I+G
11 CILP 1,158 92 661 0.0325 41.6 32.7 GTR+I+G
12 CPT2 792 63 546 0.0588 47.5 52.6 SYM+I+G
13 CXCR4 651 51 379 0.0347 47.3 60.2 HKY+I+G
14 DBC1 810 32 383 0.0713 45.9 50.4 GTR+I+G
15 DCHS1 363 41 173 0.0566 49.6 46.5 GTR+I+G
16 DET1 711 39 329 0.0397 42.9 42.7 SYM+I+G
17 DISP1 1,062 53 510 0.0344 44.0 51.1 GTR+I+G
18 DISP2 981 98 653 0.0494 42.8 43.3 GTR+I+G
19 DMXL1 1,002 80 653 0.0316 43.1 37.4 GTR+I+G
20 DNAH3 957 73 487 0.0293 41.1 37.5 HKY+I+G
21 DOLK 741 66 489 0.0504 49.0 57.8 GTR+I+G
22 DOPEY1 600 49 303 0.0343 42.1 49.6 GTR+I+G
23 DSEL 1,305 77 763 0.0211 39.2 31.3 GTR+I+G
24 ENC1 1,098 53 604 0.0308 45.7 44.1 GTR+I+G
25 EVPL 1,014 111 664 0.0404 42.0 43.1 GTR+I+G
26 EXOC8 1,143 67 658 0.0463 40.9 39.3 GTR+I+G
27 EXTL3 1,269 114 622 0.0386 46.8 49.9 GTR+I+G
28 FAT1 1,545 137 1071 0.0235 38.7 30.0 GTR+I+G
29 FAT4 756 41 467 0.0290 41.7 35.6 HKY+I+G
30 FEM1B 996 60 552 0.0553 49.2 51.4 GTR+I+G
31 FICD 525 25 271 0.0364 45.0 43.5 GTR+I+G
32 FILIP1 810 61 492 0.0328 37.1 31.7 GTR+I+G
33 FLRT3 1,059 148 559 0.0252 42.5 42.0 GTR+I+G
34 FREM2 1,152 76 688 0.0356 40.7 36.1 GTR+I+G
35 FUT9 771 58 357 0.0250 40.5 42.8 HKY+I+G
36 FZD4 762 37 329 0.0382 45.9 53.3 GTR+I+G
37 GGPS1 414 19 212 0.0366 36.4 31.1 GTR+I+G
38 GLCE 438 26 220 0.0433 48.8 50.3 HKY+I+G
39 GPER 519 30 260 0.0439 44.5 58.2 GTR+I+G
40 GRIN3A 651 33 309 0.0337 41.1 33.4 GTR+I+G
41 GRM2 687 46 374 0.0831 54.6 68.3 GTR+I+G
42 HYP 1,299 370 653 0.0414 47.4 52.2 SYM+I+G
43 IRS1 1,020 48 496 0.0391 50.9 48.9 SYM+I+G
44 KBTBD2 1,125 94 638 0.0204 44.4 42.8 GTR+I+G
45 KCNF1 765 38 374 0.0396 46.4 58.2 SYM+I+G
46 KIAA2013 537 46 300 0.0820 53.8 69.3 HKY+I+G
47 LCT 666 62 450 0.0453 45.3 44.3 GTR+I+G
48 LIG4 1,056 87 618 0.0262 38.8 37.3 GTR+I+G
49 LINGO1 1,137 65 594 0.0335 46.0 45.6 HKY+I+G
50 LINGO2 1,332 108 805 0.0266 40.4 39.9 GTR+I+G
51 LPHN2 582 38 252 0.0361 48.3 50.9 SYM+I+G
52 LRRC8D 1,155 62 554 0.0325 40.8 41.4 HKY+I+G
53 LRRN1 837 48 424 0.0563 51.0 56.5 HKY+I+G
54 LRRN3 1,113 100 607 0.0288 37.8 38.4 HKY+I+G
55 LRRTM4 1,155 60 590 0.0248 45.1 51.5 GTR+I+G
56 MB21D2 1,029 42 422 0.0260 44.0 43.8 HKY+I+G
57 MED1 627 45 329 0.0434 46.5 40.7 GTR+I+G
58 MED13 525 39 277 0.0278 43.5 42.5 GTR+I+G
59 MGAT4C 792 51 400 0.0297 36.6 33.9 HKY+I+G
60 MIOS 972 53 464 0.0865 53.0 63.2 GTR+I+G
61 MSH6 1,350 101 821 0.0426 44.2 44.1 GTR+I+G
62 MYCBP2 1,068 80 551 0.0216 42.1 34.8 GTR+I+G
63 NHS 1,044 88 639 0.0248 42.1 32.8 HKY+I+G
64 NTN1 585 22 238 0.0904 52.2 66.8 HKY+I+G
65 P2RY1 744 41 342 0.0311 39.7 48.9 GTR+I+G
66 PANX2 735 49 362 0.0429 37.8 37.2 GTR+I+G
67 PCDH1 1,479 117 774 0.0277 45.5 42.7 GTR+I+G
68 PCDH10 849 53 487 0.0582 60.3 82.8 GTR+I+G
69 PCLO 906 43 480 0.0256 39.1 24.4 GTR+I+G
70 PDP1 1,071 60 491 0.0391 46.1 44.2 GTR+I+G
71 PIK3CG 978 90 572 0.0258 38.6 35.6 GTR+I+G
72 PPL 1,389 141 998 0.0418 44.9 51.3 GTR+I+G
73 RAG1 1425 84 763 0.0331 44.3 43.9 HKY+I+G
74 RAG2 906 62 653 0.0363 43.0 44.6 GTR+I+G
75 RERE 456 28 216 0.0336 52.0 43.7 SYM+I+G
76 REV3L 675 52 512 0.0327 39.9 28.2 GTR+I+G
77 ROR2 972 78 489 0.0274 43.5 40.3 GTR+I+G
78 RP2 510 41 272 0.0393 45.6 51.7 HKY+I+G
79 SACS 1,146 56 579 0.0252 40.2 34.3 GTR+I+G
80 SALL1 1,497 96 862 0.0223 43.1 35.0 GTR+I+G
81 SETBP1 795 64 443 0.0353 43.7 42.3 GTR+I+G
82 SH3BP4 1,158 79 717 0.0403 44.9 43.8 GTR+I+G
83 SLITRK1 1,185 103 636 0.0714 46.9 55.4 GTR+I+G
84 SOCS5 999 55 504 0.0460 49.0 47.2 GTR+I+G
85 SPEN 948 80 577 0.0503 41.4 39.8 GTR+I+G
86 STON2 921 74 602 0.0506 47.6 55.1 SYM+I+G
87 SVEP1 840 81 566 0.0294 41.3 27.2 HKY+I+G
88 TTN 999 250 575 0.0232 41.1 28.0 GTR+G
89 VCPIP1 1,125 77 578 0.0592 46.8 48.7 GTR+I+G
90 VPS18 975 81 586 0.0651 42.8 49.4 GTR+I+G
91 WFIKKN2 1,155 82 699 0.0617 47.0 53.6 GTR+I+G
92 ZBED4 1077 92 518 0.0352 40.2 37.3 GTR+I+G
93 ZEB1 819 55 484 0.0270 42.5 33.4 HKY+I+G
94 ZFPM2 1,251 119 652 0.0304 45.2 41.4 GTR+I+G
95 ZHX2 993 78 649 0.0422 44.5 45.2 GTR+I+G

Higher-Level Phylogenetic Relationships of Frogs.

Maximum-likelihood (ML) and Bayesian analyses of concatenated genes in the 164-species dataset produced identical trees except for two nodes with low support (Figs. S1 and S2). The ASTRAL species tree differed from the ML tree at eight poorly supported nodes (Fig. S3). Overall support is high: 94% of 155 nodes within frogs have a bootstrap value (BS) ≥70% (Fig. S1), and 97% have Bayesian posterior probabilities (BPPs) ≥0.95 (Fig. S2). The ASTRAL species-tree method produced BSs ≥70% at 84% of the nodes (Fig. S3). Although ML analysis of the 309-species dataset has weaker support (78% of nodes have BSs ≥70%; Fig. S4), the basic topology of the ML tree is similar. Therefore, we used the ML tree as our primary hypothesis (Fig. 1A) for estimating the chronogram.

Fig. 1.

Fig. 1.

Time-calibrated phylogenetic tree of frogs and the pattern of net diversification rate across time. (A) Evolutionary chronogram based on 95% nuclear genes and 20 fossil age constraints. Gray bars represent the 95% credibility interval of divergence time estimates. Divergence time estimates and corresponding 95% credibility intervals for all nodes are provided in Table S2. Note that the initial diversification of the three major frog clades: Hyloidea (blue), Microhylidae (purple), and Natatanura (green) took place simultaneously near the KPB (dashed red line). (B) Rate-through-time plot of extant frogs indicates an increase in diversification rate at the end of the Cretaceous.

Archaeobatrachian relationships are identical to those of most recent studies (refs. 2, 5, 6, 13; but see refs. 4, 10), including an analysis of mitogenomes of 90 anuran species (7). Relationships among the early branches of Neobatrachia are identical to the mitogenomic phylogeny (7). We corroborate the placement of Heleophrynidae, which is from extreme southern Africa, as the sister taxon to all other neobatrachian frogs (Fig. 1A), as found by some authors (47, 9), but not by others (10, 11). We find that Sooglossidae (known only from the Seychelles Islands) is the sister taxon of Ranoidea (BS = 100%, BPP = 1.0; Figs. S1–S3) in the 164-species analyses, and that Sooglossidae + Nasikabatrachidae (known only from the Western Ghats of peninsular India) form the Sooglossoidea, which is the sister group of Ranoidea (BS = 100%; Fig. S4) in the 309-species analyses. Sooglossoidea is placed as the sister taxon of all other neobatrachians (11), all neobatrachian frogs to the exclusion of Heleophrynidae (6), Hyloidea + Myobatrachidae + Calyptocephalellidae (4), and Ranoidea (3, 5, 7), but no previous studies recovered the placement of Sooglossoidea with strong statistical support, including a mitogenome phylogeny (7).

Taxa within the superfamily Ranoidea have been consistently grouped into three clades (47): Microhylidae, Afrobatrachia (i.e., epifamily Brevicipitoidae, which includes Brevicipitidae, Hemisotidae, Hyperoliidae, and Arthroleptidae), and Natatanura (Fig. 1A). We found that Microhylidae is the sister group of Afrobatrachia + Natatanura (BS = 72%, PP = 1.0; Figs. S1 and S2); in contrast, many previous studies placed Microhylidae as the sister group of Afrobatrachia (46, 14) or of Natatanura (7). Relationships within the Afrobatrachia mirror those found in other studies (57, 14).

Natatanura is a large clade of extant anurans (24% of species) and mainly found in the Old World. Our ML and Bayesian topologies of Natatanura are identical. All nodes in the Bayesian tree have a BPP of 1.0, and only three nodes in the ML tree have BSs <90%. The 309-species topology is identical, but with low support among the deeper branches, likely because of missing data. Notably, we found that endemic African continental lineages (Conrauidae, Odontobatrachidae, Petropedetidae, Phrynobatrachidae, Ptychadenidae, Pyxicephalidae) form a clade that is the sister group to the clade of the remaining North American, Eurasian, Melanesian, and Malagasy lineages (Ceratobatrachidae, Dicroglossidae, Mantellidae, Rhacophoridae, and Ranidae; Fig. 1A and Figs. S1–S4). This African clade has low bootstrap support (56%) but high Bayesian support (1.0). The clade of the remaining non-African families is strongly supported (BS = 100%), and the internal branches are strongly supported (BS = 100%, BPP = 1.0), although they are short. In other studies, this group of African lineages is not monophyletic (6, 7, 10, 14, 15). The phylogenetic position of the continental African lineages has important biogeographic significance (as detailed later).

Relationships among the subfamilies of Microhylidae (one of the largest anuran families, including 8.8% of all species), which have significant radiations on most continents and the large islands Madagascar and New Guinea, have proven difficult to resolve (4, 6, 7, 1519). In contrast, our Bayesian tree has strong support; all 10 of the deepest nodes have a BPP of 1.0 (Fig. S2). The ML topology is identical, but 3 of 10 nodes have a BS <90% (Fig. S1). ML analysis of our 309-species dataset (Fig. S4) recovered the same topology, but support is weaker. Nonetheless, our tree is more strongly supported than others except for a phylogenomic analysis (19) of 66 anchored loci for 48 taxa and 7 Sanger-sequenced loci for 142 taxa.

Significantly, our study resolves relationships among one of the most diverse clades in the anuran phylogeny: Hyloidea, which contains 54% of extant anuran species. ML and Bayesian analyses strongly support 53 of 55 of nodes in Hyloidea (BS > 80% and PP = 1.0; Figs. S1 and S2). This is particularly unexpected because even the most species- and character-rich studies of hyloids (57) have recovered poorly resolved topologies. Our expanded 309-species topology (Fig. S4) is similar to our primary tree, even though many nodes have lower support, which is consistent with the higher degree of missing data for this dataset of more species.

Two arrangements within hyloids are noteworthy. The deepest divergences in Hyloidea are among southern South American taxa: Rhinodermatidae in the temperate beech forests of Chile, and Alsodidae and Batrachylidae in Patagonia (Figs. S1–S4). Similar relationships were previously reported (5, 20), albeit with weak support and a smaller sample of species. These relationships support a southern South American origin of Hyloidea. Moreover, most previous studies (4, 6, 7, 21) supported Terrarana, a large New World tropical clade (15% of extant anuran species), rather than southern South America groups, as the sister group of all other hyloids. This difference in the placement of Terrarana may result from long-branch attraction of mtDNA sequences, because Terrarana species apparently have higher rates of mtDNA evolution than other hyloid lineages (22).

In summary, our analyses corroborate many of the deeper neobatrachian nodes inferred by other studies, but with greater support (all BS values = 100%). Furthermore, we find strong support for many shallower nodes that are weakly supported in other studies; only 4 of 155 nodes in our Bayesian tree have a posterior probability <1.0 and only 9 of 155 nodes have bootstrap support <75%. Relationships among the deepest branches in the Hyloidea, Microhylidae, and Natatanura, for which previous studies have conflicting results, are now strongly supported in Hyloidea and Natatanura, and most nodes are well-supported in Microhylidae. Some clades that were not well-supported in the mitogenomic phylogeny (7) are rejected in this study. Our findings include the position of southern South American taxa as the earliest branches of hyloids, and the identification of a clade of endemic African taxa as the sister group of all other Natatanura.

A New Timescale for Extinction and Diversification.

Our extensive sample of nuclear loci produced not only a strongly supported phylogeny for frogs but also younger divergence time estimates with smaller CIs than previous studies. The estimated divergence times in the 164-species and the 309-species data sets are similar (Figs. S5 and S6); the average time deviation among comparable nodes between the two analyses is ∼5%. Following a thorough literature review, we selected 20 fossil calibration points, of which 13 are within crown Anura (Fig. S7). Jackknife removal of each fossil resulted in highly congruent sets of date estimates. The overall average time differences were less than 1% except for the two outlier fossils (the salamander Chunerpeton and the salientian Triadobatrachus), and even the two outliers only led to overall average time differences of 3–5.4% (Fig. S8).

For this discussion, we focus on the 164-species tree (Fig. 1A). Overall, our divergence times are notably younger than those found by most other studies (5, 10, 23, 24) (Fig. 1A and Table S2), and especially much younger than those in the mitogenome phylogeny (7). For example, we estimate the last common ancestor of crown-group Anura to be during the Upper Triassic at 210.0 Mya (95% CI, 199.1–220.4 Mya), and the age of crown Neobatrachia in the Lower Cretaceous at 142.1 Mya (95% CI, 132.8–149.8 Mya). In contrast, many other estimates place the age of crown Anura between the Permian and Middle Triassic and the origin of Neobatrachia between Upper Triassic and Upper Jurassic (Fig. S9).

Table S2.

Detailed node information of the concatenated analyses and divergence analyses

Node BS PP Age, Mya 95% CI
1 N/A 1.0 433.4 416.3–455.2
2 N/A 1.0 422 410.8–427.5
3 100 1.0 348.4 343.0–351.5
4 100 1.0 319 316.9–322.7
5 100 1.0 252.2 246.9–259.7
6 100 1.0 297.9 287.7–309.1
7 100 1.0 271.7 266.8–274.7
8 100 1.0 153 130.5–166.3
9 100 1.0 210 199.1–220.4
10 100 1.0 193.8 178.6–207.0
11 100 1.0 194.1 183.2–204.1
12 100 1.0 138.5 122.9–158.6
13 100 1.0 102.2 74.8–125.5
14 100 1.0 42.8 33.0–54.7
15 100 1.0 10 7.4–13.3
16 100 1.0 182.5 172.0–192.7
17 100 1.0 159.4 146.7–171.4
18 100 1.0 117.6 104.2–130.9
19 100 1.0 11.6 8.4–15.5
20 100 1.0 104.1 91.7–117.7
21 100 1.0 57.9 45.3–69.9
22 100 1.0 45.3 34.2–57.9
23 100 1.0 172.6 162.5–182.6
24 100 1.0 128 115.6–141.5
25 100 1.0 41.9 32.1–54.1
26 100 1.0 9.3 6.8–12.3
27 100 1.0 112.3 99.9–125.6
28 100 1.0 74.4 66.2–86.2
29 100 1.0 50.6 44.3–57.9
30 100 1.0 24.8 20.5–29.8
31 100 1.0 17 13.1–21.6
32 100 1.0 32.4 27.1–38.0
33 100 1.0 18.7 15.2–22.9
34 100 1.0 15.5 12.1–19.6
35 100 1.0 142.1 132.8–149.8
36 100 1.0 130 121.1–137.8
37 100 1.0 119.9 110.7–129.1
38 100 1.0 99.5 88.6–111.3
39 100 1.0 72.6 63.4–84.8
40 100 1.0 65.7 56.2–77.6
41 100 1.0 72.6 67.4–78.5
42 100 1.0 61.5 54.3–68.3
43 100 1.0 68 63.2–73.4
44 100 1.0 61.8 56.5–67.9
45 100 1.0 20.1 14.7–27.0
46 100 1.0 26.7 21.2–33.5
47 100 1.0 13.4 9.8–17.7
48 100 1.0 66.4 61.8–71.7
49 100 1.0 4.9 3.5–6.7
50 100 1.0 64.8 60.3–70.0
51 98 1.0 63.9 59.5–69.1
52 100 1.0 18.2 13.1–24.6
53 94 1.0 62.6 58.2–67.9
54 100 1.0 43.9 35.5–51.9
55 100 1.0 19.2 14.2–25.2
56 100 1.0 58.9 54.4–64.2
57 100 1.0 47.5 42.0–53.4
58 100 1.0 24.5 20.0–29.7
59 100 1.0 16.7 12.8–21.3
60 100 1.0 49.5 44.9–54.4
61 100 1.0 32.3 26.6–38.3
62 100 1.0 25.2 19.6–31.0
63 100 1.0 46.2 41.6–50.9
64 99 1.0 43.9 39.4–48.6
65 100 1.0 41.1 36.8–45.8
66 100 1.0 24.7 20.0–29.9
67 100 1.0 35.3 30.6–40.2
68 100 1.0 63.6 59.2–68.9
69 100 1.0 34.6 28.7–40.8
70 100 1.0 28.9 23.1–35.4
71 67 1.0 62.1 57.6–67.4
72 100 1.0 48.5 43.2–53.7
73 100 1.0 45.1 40.0–50.4
74 100 1.0 30.2 24.5–36.0
75 100 1.0 43.2 38.2–48.4
76 91 0.84 39.2 34.0–44.6
77 81 0.84 40.7 35.6–45.9
78 81 1.0 59.9 55.5–64.9
79 36 1.0 58.1 53.8–63.1
80 100 1.0 32 24.8–40.0
81 100 1.0 52.8 48.2–58.0
82 100 1.0 36.2 29.6–42.5
83 100 1.0 39 34.0–44.4
84 100 1.0 25.8 21.1–30.8
85 100 1.0 6.2 4.5–8.4
86 100 1.0 48 43.1–53.2
87 100 1.0 35.4 30.8–40.5
88 100 1.0 25.4 22.4–28.9
89 100 1.0 21 18.5–24.0
90 100 1.0 18 15.4–20.9
91 100 1.0 17.1 14.6–20.0
92 87 0.99 16.4 13.9–19.4
93 100 1.0 19.4 16.9–22.5
94 100 1.0 16.5 14.0–19.4
95 100 1.0 12.3 9.7–15.2
96 100 1.0 122.9 114.1–131.0
97 100 1.0 100.9 92.8–108.6
98 100 1.0 67.1 61.0–72.9
99 39 1.0 63.6 57.7–69.2
100 100 1.0 59.6 53.9–65.3
101 100 1.0 48.1 42.1–53.8
102 100 1.0 27.9 23.0–32.8
103 100 1.0 21 16.6–25.9
104 40 1.0 62 56.1–67.4
105 100 1.0 56.8 50.8–62.6
106 100 1.0 32.3 25.2–39.9
107 57 NR 30.3 23.3–38.1
108 100 1.0 47.7 41.3–54.3
109 100 1.0 7.2 5.2–9.8
110 35 1.0 60.9 55.1–66.2
111 100 1.0 54.8 49.2–60.4
112 99 1.0 52.7 47.2–58.2
113 100 1.0 45.9 40.1–51.5
114 100 1.0 15 11.1–20.1
115 100 1.0 20.4 17.6–23.5
116 100 1.0 16.9 14.1–20.0
117 42 NR 19.7 16.9–22.8
118 100 1.0 16.8 14.3–19.8
119 100 1.0 14.2 11.9–16.9
120 100 1.0 12.5 10.1–15.2
121 72 1.0 96.6 88.6–104.1
122 100 1.0 82 73.7–89.7
123 100 1.0 53.8 45.2–62.8
124 100 1.0 34.3 26.7–42.7
125 100 1.0 63.9 56.7–70.9
126 100 1.0 39.9 31.9–48.0
127 100 1.0 48.9 43.3–55.1
128 92 1.0 46.2 40.7–51.9
129 94 1.0 42.5 36.9–48.3
130 100 1.0 36.9 31.4–42.6
131 100 1.0 34.8 29.2–40.4
132 100 1.0 20.9 16.1–26.5
133 100 1.0 64.5 59.1–69.8
134 56 1.0 63.1 57.7–68.4
135 64 1.0 61 55.5–66.2
136 100 1.0 54.5 48.3–60.1
137 97 1.0 61.3 55.8–66.5
138 80 1.0 58.3 52.7–63.9
139 100 1.0 47 40.1–53.6
140 100 1.0 7.2 5.3–9.6
141 100 1.0 62.7 57.3–67.8
142 100 1.0 39.4 31.2–47.1
143 100 1.0 61.2 56.0–66.4
144 100 1.0 45.9 39.6–52.2
145 91 1.0 42.8 36.8–49.0
146 100 1.0 26.6 21.2–32.8
147 100 1.0 59.1 54.1–64.3
148 100 1.0 53.7 48.6–59.0
149 100 1.0 42.7 35.9–48.9
150 100 1.0 42.5 37.3–48.1
151 100 1.0 26.2 22.0–31.1
152 100 1.0 22.5 18.1–27.2
153 100 1.0 30.6 27.1–34.3
154 100 1.0 26.9 24.0–30.2
155 100 1.0 18.8 14.8–22.9
156 100 1.0 24.4 21.6–27.5
157 100 1.0 17.7 14.9–20.7
158 100 1.0 14.4 11.5–17.4
159 100 1.0 22.2 19.5–25.3
160 100 1.0 16.4 14.2–18.9
161 100 1.0 11.4 8.7–14.1
162 100 1.0 12.6 10.6–14.7
163 100 1.0 6.5 5.1–8.3

The node number corresponds to Fig. S5. BS represents ML bootstrap support value in Fig. S1. PP represents Bayesian posterior probability in Fig. S2. N/A, not applicable.

A striking pattern is the synchronous origin of three species-rich neobatrachian clades—Hyloidea, Microhylidae, and Natatanura—at the Cretaceous–Paleogene (K–Pg) boundary (KPB;Fig. 1A, dashed red line). The diversification synchronicity of the three frog clades still existed when all calibration constraints in frogs were excluded (Fig. S10), suggesting that this pattern is unlikely the result of the choice of calibration points. The K–Pg boundary (KPB), dated precisely at 66 Mya (25), marks one of Earth's great extinctions, largely attributed to the impact of the Chicxulub bolide. However, other factors such as climate warming and the Indian Deccan volcanism, possibly related to the bolide impact, likely contributed to this extinction event (26). Although a near-KPB origin was separately reported for Hyloidea (5, 8, 10, 20, 27) and Microhylidae (10, 16) (Fig. S9), our study found that these three major clades originated at the same time, very near the KPB. The contemporaneous origin of the three large clades is highlighted in our results by the narrow 95% CIs for each and their overlap with the KPB. Nine of the 10 deepest nodes of Hyloidea have relatively narrow CIs that overlap the KPB (9.6–14.0 My; Fig. 1A and Table S2). Similarly, within Microhylidae, the CIs of the four deepest nodes encompass the KPB, and the CIs are similarly narrow (11.1–11.9 My). Finally, in Natatanura, the CIs of the six deepest nodes all overlap the KPB, also with similarly narrow CIs (10.4–10.7 My). The use of internal calibration points for these clades is unlikely to be driving this pattern because there was only one internal calibration point (within the Natatanura) used within these three clades.

There are several reasons why our study may have produced younger divergence times compared with previous studies. First, we included diverse outgroups, including lungfish, coelacanth, salamanders, caecilians, and amniotes. Second, we used a new set of calibration points rather than uncritically recycling calibration points from other studies; these calibration points have been carefully reexamined based on the fossil record. The choice of calibration points has significant impact on divergence time estimation for frogs and may be an important reason that older divergence times were obtained in previous studies (27). Third, our phylogenetic estimate is based on a much larger dataset (88 kb; 62% of sites are variable) derived solely from 95 nuclear loci (rather than primarily from mitochondrial loci) that exceeds other studies by at least sevenfold. We attribute the greater precision (i.e., smaller CIs) to the large amount of information within our dataset. A similar increase in precision was found for plethodontid salamanders by using the same nuclear marker toolkit (28).

Rate-through-time analysis indicates that a surge in net diversification rates of frogs occurred immediately following the KPB (Fig. 1B), which suggests a clear impact of the KPB extinction on frog diversification. The rapidity of the diversification is reflected in the short branches of the deepest nodes in each clade. In Hyloidea, all but one branch connecting the 10 deepest nodes are <4.6 My in duration. In Microhylidae, the branches connecting the four deepest nodes are <4.0 My long. In Natatanura, the branches connecting the six deepest nodes are <2.1 My. Even though these branches are short, they are strongly supported. Another line of evidence demonstrates three mass extinctions preceding explosive speciation. The branches subtending the Hyloidea, Natatanura, and Microhylidae are long: 47.3, 32.1, and 33.8 My, respectively (Fig. 1A and Table S2). The lack of other extant lineages originating from these three stem lineages corroborates our suggestion that an extinction event simultaneously decimated these lineages near the KPB.

Even though the fossil record indicates mass extinction at the KPB for several taxa, including birds, squamates, and mammals, and subsequent radiation post-KPB (2931), molecular data often indicate that many clades had diversified, at least in the sense of lineage splitting, before the KPB event (3236). Unlike nonavian dinosaurs, whose demise is well documented in the rocks, the fossil record of frogs is so far largely uninformative about survival/extinction across the KPB (3740). In the present study, the three parallel combinations of precise node ages overlapping the KPB, preceded by a long branch and followed by very short but strongly supported branches, indicate three extinctions followed by rapid divergence. These three radiations, which are coincident with the late Cretaceous extinction, account for 88% of extant frog species. This molecular phylogenetic perspective strongly suggests that frogs experienced a major extinction at the KPB.

Historical Biogeography of Frogs.

We performed a biogeographic analysis on the 309-species data set using BioGeoBEARS (41). The DEC+J model performs significantly better than the DEC model (Akaike information criterion; Table S3), indicating the importance of the J parameter, which models long-distance or “jump” dispersal. Our interpretation is that dispersal into a new area is accompanied by near-instantaneous speciation. The practical effect of adding the J parameter to the model is that ancestral ranges often comprise one area rather than several.

Table S3.

Models and parameters of ancestral range estimation of frogs

Model LnL Parameters d e j AIC w
DEC −257.30 2 0.13 0.02 518.6 5.0 × 10−22
DEC+J −207.26 3 0.02 1.0 × 10−12 0.04 420.5 1

Models include dispersal-extinction cladogenesis (DEC), and the same model allowing for founder event speciation (+J). AIC, Akaike Information Criterion; d, rate of range expansion by adding an area; e, rate of range reduction through extirpation in an area; j, relative per-event weight of jump dispersal at cladogenesis; ln L, log-likelihood; w, Akaike weights. The best-fit model is the DEC+J model.

The most recent common ancestor (MRCA) of extant frogs was distributed in Eurasia + North America + Australia or North America + Australia (Fig. 2 and Fig. S11), a Pangaean origin. In contrast, neobatrachians (i.e., modern frogs) originated in Gondwanaland, most likely in Africa (relative probability, 93.6%; Fig. 2). The major neobatrachian lineages are also Gondwanan in origin: Hyloidea in South America and Ranoidea, Afrobatrachia, and Natatanura in Africa (relative probability, >90%; Fig. 2).

Fig. 2.

Fig. 2.

Ancestral-area estimates for 69 terminal taxa (families, subfamilies, and genera) of extant frogs using the DEC+J model in BioGeoBEARS. Circles on nodes represent the set of possible ancestral areas, and the color is associated with the area legends. The probabilities are given next to circles for the most probable ancestral area. Circles without values indicate that the probability of the ancestral area is >99%. Three important landmass breakup events are indicated: (A) the break-up of Pangea with division into Laurasia and Gondwana in the Middle Jurassic coincident with the origin of neobatrachian frogs; (B) the separation of Africa and South America in the Early Cretaceous coincident with the divergence of Ranoidea and Hyloidea as well as between the African and New World pipids; and (C) the separation of the Seychelles and India in the Late Cretaceous coincident with the divergence between the Sooglossidae and Nasikabatrachidae. Anuran taxa that contain at least some arboreal species are indicated in green and a tree icon (we do not imply that the last common ancestor of each of these families was arboreal). Note that all clades containing arboreal frogs originated after the KPB.

Breakup of land masses may be associated with three divergence events in the evolution of frogs. The first of these was the split of neobatrachians from ancestral groups (Fig. 2A), dated at 172–181 Mya, which agrees with the breakup of Pangaea into Laurasia and Gondwana in the Early Jurassic (∼180 Mya) (42). This breakup event, which isolated the MRCA of Neobatrachia in Gondwana, is associated with the initial diversification of the clade. The second breakup event caused the split between the two major lineages of neobatrachians, Procoela (Hyloidea and Myobatrachoidea, mostly South American) and Diplasiocoela (Ranoidea + Sooglossoidea, mostly African), as well as the split between the African and South American pipids (Fig. 2B). Spreading of the South Atlantic Ocean sea floor started in the Early Cretaceous (135 Mya), but the final physical separation between Africa and South America took place approximately 105 Mya (42). Our estimates for timing of these divergences, 125–130 Mya and ∼120 Mya, respectively, are highly congruent with this rifting process. The third breakup event took place when the Sooglossidae (Seychelles) and the Nasikabatrachidae (India) diverged (Fig. 2C). The Seychelles/India land mass continued to exist until 66 Mya, when new rifting severed the Seychelles from India (43). We estimate Sooglossidae split from Nasikabatrachidae at ∼66 Mya, which is considerably younger than the previous estimates (77–130 Mya) (10, 11, 23), in congruence with the geological event.

The MRCA of Hyloidea and of Myobatrachidae + Calyptocephalellidae occurred in South America (Fig. 2). The divergence between Myobatrachidae (Australia) and Calyptocephalellidae (South America) and the split between Phyllomedusinae (South America) and Pelodryadinae (Australia and New Guinea) took place ∼100 Mya and ∼50 Mya, respectively (Fig. 2). During this period, South America and Australia were distantly separated but connected intermittently via Antarctica (42). From the late Cretaceous to the early Tertiary, the Earth experienced a period of global warming (44). Climate data from plant fossils, sediments, and geochemical indicators show that the mean annual temperature of the Antarctic Peninsula region was 10–20 °C from 100 to 50 Mya (45), which is sufficiently warm to allow dispersal of frogs through this region. Thus, the origin of the Australian myobatrachids and pelodryadine hylids is most likely explained by dispersal from South America to Australia through Antarctica, and later extinction in Antarctica because of the formation of ice sheets. The role of Antarctica in dispersal of frogs among Gondwanan land masses has received little attention (but see ref. 46), but our time-calibrated phylogeny predicts that paleontological research in the Cretaceous and early Cenozoic of Antarctica could shed light on the early evolution and dispersal, especially of hyloids.

The nearly worldwide distribution of Microhylidae presents a longstanding and challenging biogeographic puzzle. Vicariant origin and long-distance oceanic dispersal have been proposed to explain the current distribution of this family (10, 15, 16, 18, 23). Which scenario dominates interpretation depends largely on the time and location of origin for the Microhylidae. Our analyses place the initial divergence of the Microhylidae at ∼66 Mya, with a relative probability of 79.5% for an African origin (Fig. 2). By this time, Gondwana was already highly fragmented and Africa was separated from South America and Madagascar by ocean. Accordingly, overseas dispersal of the major microhylid lineages on Gondwanan landmasses is required.

The origin and diversification of the second large clade, Natatanura, is controversial. Two hypotheses have been proposed. The Out-of-Africa hypothesis (47), argues that the origin of the Natatanura lies in Africa, with subsequent dispersal to other continents. The Out-of-India hypothesis (48) postulates that the ancestor of Natatanura originated on the Indian plate and was confined there until the plate collided with Asia approximately 55 Mya. We find that Natatanura originated in Africa (relative probability, 92.5%; Fig. 2). Considering that Natatanura is the sister group of Afrobatrachia (apparently of African origin) and the basal split within the Natatanura phylogeny separates all African continental lineages from other Asian, Indian, and Madagascar lineages (Fig. 2), an African origin seems more likely. The endemic Indian natatanurans (Nyctibatrachidae, Micrixalidae, and Ranixalidae) are nested within Asian lineages, and the divergences between them and their closest Asian relatives occurred between 55 and 60 Mya (Fig. 2), which is consistent with the timing of the India–Asia collision (∼55 Mya) (49). Intriguingly, the divergence between the Asian tree frogs (Rhacophoridae) and the Malagasy mantellids also happened during this period (Fig. 2), implying that the Indian plate served as a stepping stone for the long-distance dispersal from Asia to Madagascar.

The three long branches of similar duration leading to Hyloidea, Microhylidae, and Natatanura suggest that the K–Pg mass extinction event may have triggered explosive radiations by emptying ecological space. Notably, all terminal taxa (family and subfamily ranks) with arboreal species originate after the K–Pg boundary (Fig. 2). The rebounding of forests after the massive loss of vegetation (50) at the KPB may have provided new ecological opportunities for the subsequent radiation of largely arboreal groups such as Hylidae, Centrolenidae, and Rhacophoridae. The observation that no lineages of frogs originating before the KPB (archaeobatrachians, Heleophrynidae, Myobatrachoidea, and Sooglossoidea) have truly arboreal species, and that all origins of arboreality (e.g., within hyloids or natatanurans) follow the KPB, supports the hypothesis that the K–Pg mass extinction shaped the current diversity of frogs.

Materials and Methods

Taxon Sampling and Data Collection.

Our sampling included 156 frog species from 44 of the 55 recognized families (1). Eight outgroup species, including two salamanders, one caecilian, one bird, one crocodile, one mammal, and two lobe-finned fishes, were used in all phylogenetic analyses. Total DNA was isolated from frozen or ethanol-preserved tissues (liver or muscle) by proteinase K digestion followed by standard salt extraction protocol. Previously published PCR primers and PCR protocols (12) were used to amplify 95 unlinked nuclear protein-coding genes (including RAG1 and CXCR4) from the DNA extracts in 96-well plates. The amplification products were sequenced using a next-generation sequencing (NGS) strategy as described by Feng et al. (51). Briefly, all amplification products from a single species were pooled together and purified. The amplification product pool of a sample was then randomly sheared to small fragments (200–500 bp), ends were repaired, and a species-specific barcode linker was added. All indexed amplification product pools were mixed together. A sequencing library was constructed with the pooled DNA by using the TruSeq DNA Sample Preparation kit and sequenced on a Illumina HiSEq.2500 sequencer. Approximately 3 GB of Illumina HiSeq paired-end 90-bp reads were obtained. These reads were bioinformatically sorted by barcode sequence and assembled into consensus sequences. NGS protocol of library construction and bioinformatic analyses have been described previously (51). All sequences were compared by using BLAST against GenBank to ensure that the target sequences were amplified. The remaining sequences were further checked for frame shift or stop codons. GenBank accession numbers for the new sequences are given in Dataset S1.

GenBank Data.

To provide a comprehensive phylogeny at the family level, we retrieved RAG1 and CXCR4 sequences from GenBank for 145 additional species of frogs (Dataset S1). These data plus our new sequences comprise a combined dataset that contains 301 frog species from all recognized anuran families.

Phylogenetic Analyses.

The 95 nuclear protein coding genes were aligned by using the ClustalW algorithm as implemented in MEGA v6 (52). Ambiguously aligned regions were culled using GBlocks v.0.91b (53) with the “codon” model (−t = c), smaller block (−b4 = 3), and all gaps allowed (−b5 = a). All refined alignments were then concatenated into a concatenated supermatrix. Two supermatrices were built: 164-species dataset (156 frogs and 8 outgroups) and 309-species dataset (301 frogs and 8 outgroups). PartitionFinder v.1.1.1 (54) was used to select models and partitioning schemes for the two supermatrices according to the Bayesian information criterion. A total of 285 data partitions were selected as the best partitioning scheme that corresponded to the three separate codon positions for each of the 95 genes (Table S4). The ML tree was estimated by using RAxML version 8.0 (55) with the GTR + Γ + I model assigned to each partition. Support for nodes in the ML tree was assessed with a rapid bootstrap analysis (option –f a) with 1,000 replicates. The Bayesian tree was inferred using MrBayes 3.2 (56) using the models and partitions identified by PartitionFinder. Two Markov chain Monte Carlo (MCMC) runs were performed with one cold chain and three heated chains (temperature set to 0.1) for 50 million generations and sampled every 1,000 generations. Chain stationarity was visualized by plotting likelihoods against the generation number by using TRACER v1.6 (beast.bio.ed.ac.uk/Tracer). The effective sample sizes were greater than 200 for all parameters after the first 10% of generations were discarded. Species tree analysis without gene concatenation was performed by using the Accurate Species TRee ALgorithm (ASTRAL) (57) under the coalescent model. The individual input gene trees were inferred from partitioned ML analyses by using RAxML with the same GTR + Γ + I model assigned to each codon position of each gene. The species tree analysis was conducted by using ASTRAL under the multilocus bootstrapping option with 200 replicates (−r = 200).

Table S4.

Comparisons of partitioning schemes using AIC, AICc, and BIC

Partitions Rank No. of parameters Ln L AIC AICc BIC
285p 1 2,989 −2,869,254.946 5,743,956.247 5,744,203.285 5,772,548.75
190p 2 2,124 −2,875,870.896 5,755,641.612 5,755,762.151 5,775,929.993
3p 3 357 −2,891,347.478 5,783,408.955 5,783,411.863 5,786,760.486
2p 4 346 −2,895,102.496 5,790,896.993 5,790,899.724 5,794,145.256
95p 5 1,277 −2,919,155.093 5,840,791.627 5,840,830.889 5,852,852.715
1p 6 335 −2,931,059.583 5,862,789.166 5,862,791.726 5,865,934.16

The Ln L and number of parameters are represented with results from BIC.

Divergence Time Analyses.

Divergence time estimation was conducted by using the program MCMCTREE in the PAML package (58). The ML topology was used as the reference tree. Twenty calibration points were used to calibrate the clock (Fig. S7). The ML estimates of branch lengths for each of the 95 nuclear genes were obtained by using BASEML (in PAML) programs under the GTR + Γ model. Based on the mean estimate from the 95 genes using a strict molecular clock with a 450-Mya root age (the divergence between Latimeria and Protopterus; ref. 59), the prior for the overall substitution rate (rgene gamma) was set at G (1, 11.96, 1). The prior for the rate-drift parameter (sigma2 gamma) was set at G (1, 4.5, 1). The 20 calibration points were specified with soft boundaries by using 2.5% tail probabilities above and below their limits; this is a built-in function of MCMCTREE. The independent rate model (clock = 2 in MCMCTREE) was used to specify the rate priors for internal nodes. The MCMC run was first executed for 10,000,000 generations as burn-in, then sampled every 1,000 generations until a total of 10,000 samples was collected. Two MCMC runs using random seeds were compared for convergence, and similar results were found.

Rate-Through-Time Analysis.

We investigated the diversification tempo of frogs using the program Bayesian Analysis of Macroevolutionary Mixtures (BAMM), v2.5 (60). The 309-species chronogram was used as the input tree. To account for incomplete taxon sampling, the time tree was pruned to family level, and the sampling fraction of each family was calculated based on the number of species following AmphibiaWeb. The BAMM analysis was run for 100 million generations at a temperature increment parameter of 0.01 and sampled event data every 1,000 generations. The first 20% samples were removed as burn-in. The rate-through-time plot was summarized and visualized by using BAMMtools (61) from the remaining 80% event data samples.

Biogeographic Analyses.

Based on the current distribution pattern of frogs, we defined seven biogeographic areas: Africa, Eurasia (Europe and Asia with exception of Indian plate), India (including Sri Lanka and the Seychelles), Madagascar, North America (northern Mexico, United States, and Canada), South America, and Australia (Australia, New Zealand, and New Guinea; Fig. 2). Connectivity of these biogeographic areas was modeled with three dispersal probability categories: 0.01 for well-separated areas, 0.5 for moderately separate areas, and 1.0 for well-connected areas. Area connectivity and dispersal probability were modeled in seven time slices: 0–30 Mya, 30–66 Mya, 66–90 Mya, 90–120 Mya, 120–160 Mya, 160–200 Mya, and 200–270 Mya (Table S5).

Table S5.

BioGeoBEARS dispersal multipliers for seven slices

graphic file with name pnas.1704632114st05.jpg

Dispersal rates were assigned as 1 for neighboring areas connected by land bridge, 0.5 for neighboring but not connected areas, and 0.01 for nonneighboring areas or areas separated by large ocean barriers. A, Africa; E, Eurasia (Europe and Asia with exception of Indian plate); I, India-Seychelles; M, Madagascar; N, North America; O, Australia (Australia, New Zealand, and New Guinea); S, South America. The paleogeography maps are based on ref. 42 and from the website jan.ucc.nau.edu/rcb7/index.html.

Biogeographic analyses were performed by using BioGeoBEARS (41). We used the 309-species chronogram generated by our divergence time analyses as the input phylogeny. The current distribution of each frog species was assigned based on data from AmphibiaWeb (Table S6). The maximum number of ancestral areas allowed at each node was set to four. We compared the DEC and DEC+J models to determine the influence of founder-event dispersal on biogeographic patterns. The AIC criterion selected the DEC+J as the best-fitting model (Table S3), and this was subsequently used to infer the most likely biogeographic history of anurans.

Table S6.

Taxonomic information and geographical distribution used for biogeographical analysis

No. Species Taxonomy Distribution
1 Allophryne ruthveni Allophrynidae South America
2 Alsodes gargola Alsodidae South America
3 Eupsophus calcaratus Alsodidae South America
4 Alytes obstetricans Alytidae Eurasia
5 Discoglossus pictus Alytidae Eurasia
6 Arthroleptis poecilonotus Arthroleptidae Africa
7 Arthroleptis variabilis Arthroleptidae Africa
8 Astylosternus diadematus Arthroleptidae Africa
9 Leptodactylodon ovatus Arthroleptidae Africa
10 Leptopelis kivuensis Arthroleptidae Africa
11 Leptopelis parkeri Arthroleptidae Africa
12 Nyctibates corrugatus Arthroleptidae Africa
13 Scotobleps gabonicus Arthroleptidae Africa
14 Trichobatrachus robustus Arthroleptidae Africa
15 Ascaphus truei Ascaphidae North America
16 Atelognathus reverberii Batrachylidae South America
17 Batrachyla leptopus Batrachylidae South America
18 Batrachyla taeniata Batrachylidae South America
19 Barbourula busuangensis Bombinatoridae Eurasia
20 Bombina fortinuptialis Bombinatoridae Eurasia
21 Bombina orientalis Bombinatoridae Eurasia
22 Brachycephalus ephippium Brachycephalidae South America
23 Ischnocnema guentheri Brachycephalidae South America
24 Breviceps macrops Brevicipitidae Africa
25 Breviceps mossambicus Brevicipitidae Africa
26 Callulina kreffti Brevicipitidae Africa
27 Callulina laphami Brevicipitidae Africa
28 Probreviceps durirostris Brevicipitidae Africa
29 Spelaeophryne methneri Brevicipitidae Africa
30 Sclerophrys brauni Bufonidae Africa
31 Sclerophrys maculata Bufonidae Africa
32 Mertensophryne micranotis Bufonidae Africa
33 Nectophrynoides tornieri Bufonidae Africa
34 Schismaderma carens Bufonidae Africa
35 Bufo bufo Bufonidae Eurasia
36 Bufo gargarizans Bufonidae Eurasia
37 Bufo japonicus Bufonidae Eurasia
38 Bufotes viridis Bufonidae Eurasia
39 Epidalea calamita Bufonidae Eurasia
40 Ingerophrynus divergens Bufonidae Eurasia
41 Ingerophrynus galeatus Bufonidae Eurasia
42 Leptophryne borbonica Bufonidae Eurasia
43 Rentapia hosii Bufonidae Eurasia
44 Phrynoidis aspera Bufonidae Eurasia
45 Phrynoidis juxtaspera Bufonidae Eurasia
46 Duttaphrynus melanostictus Bufonidae Eurasia, India-Seychelles
47 Duttaphrynus stomaticus Bufonidae Eurasia, India-Seychelles
48 Adenomus kelaartii Bufonidae India-Seychelles
49 Amazophrynella minuta Bufonidae South America
50 Atelopus peruensis Bufonidae South America
51 Incilius luetkenii Bufonidae South America
52 Incilius nebulifer Bufonidae South America
53 Melanophryniscus stelzneri Bufonidae South America
54 Nannophryne cophotis Bufonidae South America
55 Nannophryne variegata Bufonidae South America
56 Peltophryne lemur Bufonidae South America
57 Peltophryne longinasus Bufonidae South America
58 Peltophryne peltocephala Bufonidae South America
59 Rhaebo glaberrimus Bufonidae South America
60 Rhaebo nasicus Bufonidae South America
61 Rhinella marina Bufonidae South America
62 Rhinella ocellata Bufonidae South America
63 Anaxyrus canorus Bufonidae North America
64 Anaxyrus punctatus Bufonidae North America
65 Calyptocephalella gayi Calyptocephalellidae South America
66 Centrolene bacatum Centrolenidae South America
67 Centrolene daidaleum Centrolenidae South America
68 Chimerella mariaelenae Centrolenidae South America
69 Cochranella granulosa Centrolenidae South America
70 Espadarana prosoblepon Centrolenidae South America
71 Hyalinobatrachium aureoguttatum Centrolenidae South America
72 Hyalinobatrachium colymbiphyllum Centrolenidae South America
73 Hyalinobatrachium ibama Centrolenidae South America
74 Ikakogi tayrona Centrolenidae South America
75 Nymphargus grandisonae Centrolenidae South America
76 Rulyrana adiazeta Centrolenidae South America
77 Rulyrana flavopunctata Centrolenidae South America
78 Sachatamia ilex Centrolenidae South America
79 Teratohyla spinosa Centrolenidae South America
80 Vitreorana helenae Centrolenidae South America
81 Platymantis hazelae Ceratobatrachidae Eurasia
82 Cornufer pelewensis Ceratobatrachidae Eurasia
83 Ceratophrys cornuta Ceratophryidae South America
84 Ceratophrys ornata Ceratophryidae South America
85 Lepidobatrachus laevis Ceratophryidae South America
86 Lepidobatrachus sp. Ceratophryidae South America
87 Ceuthomantis smaragdinus Ceuthomantidae South America
88 Conraua crassipes Conrauidae Africa
89 Craugastor fitzingeri Craugastoridae South America
90 Craugastor podiciferus Craugastoridae South America
91 Craugastor augusti Craugastoridae North America
92 Thoropa taophora Cycloramphidae South America
93 Allobates femoralis Dendrobatidae South America
94 Hyloxalus jacobuspetersi Dendrobatidae South America
95 Dendrobates auratus Dendrobatidae South America
96 Epipedobates tricolor Dendrobatidae South America
97 Phyllobates vittatus Dendrobatidae South America
98 Ranitomeya imitator Dendrobatidae South America
99 Euphlyctis cyanophlyctis Dicroglossidae: Dicroglossinae Eurasia
100 Fejervarya limnocharis Dicroglossidae: Dicroglossinae Eurasia
101 Fejervarya multistriata Dicroglossidae: Dicroglossinae Eurasia
102 Hoplobatrachus tigerinus Dicroglossidae: Dicroglossinae Eurasia
103 Limnonectes fujianensis Dicroglossidae: Dicroglossinae Eurasia
104 Limnonectes laticeps Dicroglossidae: Dicroglossinae Eurasia
105 Limnonectes magnus Dicroglossidae: Dicroglossinae Eurasia
106 Quasipaa spinosa Dicroglossidae: Dicroglossinae Eurasia
107 Fejervarya granosa Dicroglossidae: Dicroglossinae India-Seychelles
108 Liurana xizangensis Dicroglossidae: Occidozyginae Eurasia
109 Occidozyga laevis Dicroglossidae: Occidozyginae Eurasia
110 Occidozyga lima Dicroglossidae: Occidozyginae Eurasia
111 Diasporus diastema Eleutherodactylidae South America
112 Eleutherodactylus coqui Eleutherodactylidae South America
113 Eleutherodactylus planirostris Eleutherodactylidae South America
114 Eleutherodactylus marnockii Eleutherodactylidae North America
115 Heleophryne purcelli Heleophrynidae Africa
116 Cryptobatrachus boulengeri Hemiphractidae South America
117 Flectonotus fitzgeraldi Hemiphractidae South America
118 Gastrotheca pseustes Hemiphractidae South America
119 Gastrotheca weinlandii Hemiphractidae South America
120 Hemiphractus bubalus Hemiphractidae South America
121 Stefania ginesi Hemiphractidae South America
122 Hemisus marmoratus Hemisotidae Africa
123 Hyla chinensis Hylidae: Hylinae Eurasia
124 Aplastodiscus perviridis Hylidae: Hylinae South America
125 Dendropsophus parviceps Hylidae: Hylinae South America
126 Hyloscirtus lindae Hylidae: Hylinae South America
127 Hypsiboas fasciatus Hylidae: Hylinae South America
128 Osteocephalus taurinus Hylidae: Hylinae South America
129 Pseudis paradoxa Hylidae: Hylinae South America
130 Scinax ruber Hylidae: Hylinae South America
131 Trachycephalus typhonius Hylidae: Hylinae South America
132 Acris crepitans Hylidae: Hylinae North America
133 Hyla arenicolor Hylidae: Hylinae North America
134 Hyla cinerea Hylidae: Hylinae North America
135 Cyclorana maini Hylidae: Pelodryadinae Australia
136 Litoria caerulea Hylidae: Pelodryadinae Australia
137 Nyctimystes kubori Hylidae: Pelodryadinae Australia
138 Nyctimystes pulcher Hylidae: Pelodryadinae Australia
139 Agalychnis callidryas Hylidae: Phyllomedusinae South America
140 Agalychnis lemur Hylidae: Phyllomedusinae South America
141 Agalychnis saltator Hylidae: Phyllomedusinae South America
142 Cruziohyla calcarifer Hylidae: Phyllomedusinae South America
143 Phyllomedusa hypochondrialis Hylidae: Phyllomedusinae South America
144 Phyllomedusa tomopterna Hylidae: Phyllomedusinae South America
145 Hylodes nasus Hylodidae South America
146 Afrixalus dorsalis Hyperoliidae Africa
147 Hyperolius bolifambae Hyperoliidae Africa
148 Hyperolius viridiflavus Hyperoliidae Africa
149 Phlyctimantis boulengeri Hyperoliidae Africa
150 Leiopelma hochstetteri Leiopelmatidae Australia
151 Physalaemus pustulosus Leptodactylidae South America
152 Leptodactylus albilabris Leptodactylidae South America
153 Leptodactylus latrans Leptodactylidae South America
154 Leptodactylus melanonotus Leptodactylidae South America
155 Lithodytes lineatus Leptodactylidae South America
156 Physalaemus cuvieri Leptodactylidae South America
157 Pleurodema thaul Leptodactylidae South America
158 Pleurodema somuncurensis Leptodactylidae South America
159 Boophis madagascariensis Mantellidae: Boophinae Madagascar
160 Boophis xerophilus Mantellidae: Boophinae Madagascar
161 Aglyptodactylus madagascariensis Mantellidae: Laliostominae Madagascar
162 Blommersia wittei Mantellidae: Mantellinae Madagascar
163 Mantella madagascariensis Mantellidae: Mantellinae Madagascar
164 Brachytarsophrys feae Megophryidae Eurasia
165 Leptobrachium chapaense Megophryidae Eurasia
166 Leptolalax alpinus Megophryidae Eurasia
167 Megophrys nasuta Megophryidae Eurasia
168 Ophryophryne microstoma Megophryidae Eurasia
169 Oreolalax jingdongensis Megophryidae Eurasia
170 Scutiger gongshanensis Megophryidae Eurasia
171 Xenophrys omeimontis Megophryidae Eurasia
172 Micrixalus sp. Micrixalidae India-Seychelles
173 Gastrophrynoides immaculatus Microhylidae: Asterophryinae Eurasia
174 Callulops wilhelmanus Microhylidae: Asterophryinae Australia
175 Cophixalus cheesmanae Microhylidae: Asterophryinae Australia
176 Cophixalus cryptotympanum Microhylidae: Asterophryinae Australia
177 Cophixalus sp. A Microhylidae: Asterophryinae Australia
178 Cophixalus sp. B Microhylidae: Asterophryinae Australia
179 Hylophorbus rufescens Microhylidae: Asterophryinae Australia
180 Liophryne schlaginhaufeni Microhylidae: Asterophryinae Australia
181 Mantophryne lateralis Microhylidae: Asterophryinae Australia
182 Oreophryne sp. A Microhylidae: Asterophryinae Australia
183 Oreophryne sp. B Microhylidae: Asterophryinae Australia
184 Xenorhina obesa Microhylidae: Asterophryinae Australia
185 Xenorhina sp. Microhylidae: Asterophryinae Australia
186 Anodonthyla boulengerii Microhylidae: Cophylinae Madagascar
187 Platypelis tuberifera Microhylidae: Cophylinae Madagascar
188 Plethodontohyla inguinalis Microhylidae: Cophylinae Madagascar
189 Stumpffia pygmaea Microhylidae: Cophylinae Madagascar
190 Dyscophus antongilii Microhylidae: Dyscophinae Madagascar
191 Chiasmocleis ventrimaculata Microhylidae: Gastrophryninae South America
192 Ctenophryne geayi Microhylidae: Gastrophryninae South America
193 Elachistocleis ovalis Microhylidae: Gastrophryninae South America
194 Stereocyclops incrassatus Microhylidae: Gastrophryninae South America
195 Gastrophryne olivacea Microhylidae: Gastrophryninae North America
196 Hoplophryne rogersi Microhylidae: Hoplophryninae Africa
197 Kalophrynus interlineatus Microhylidae: Kalophryninae Eurasia
198 Kalophrynus pleurostigma Microhylidae: Kalophryninae Eurasia
199 Melanobatrachus indicus Microhylidae: Melanobatrachinae India-Seychelles
200 Calluella guttulata Microhylidae: Microhylinae Eurasia
201 Chaperina fusca Microhylidae: Microhylinae Eurasia
202 Kaloula conjuncta Microhylidae: Microhylinae Eurasia
203 Kaloula pulchra Microhylidae: Microhylinae Eurasia
204 Metaphrynella pollicaris Microhylidae: Microhylinae Eurasia
205 Metaphrynella sundana Microhylidae: Microhylinae Eurasia
206 Microhyla annectens Microhylidae: Microhylinae Eurasia
207 Microhyla heymonsi Microhylidae: Microhylinae Eurasia
208 Microhyla marmorata Microhylidae: Microhylinae Eurasia
209 Microhyla okinavensis Microhylidae: Microhylinae Eurasia
210 Micryletta inornata Microhylidae: Microhylinae Eurasia
211 Phrynella pulchra Microhylidae: Microhylinae Eurasia
212 Uperodon montanus Microhylidae: Microhylinae India-Seychelles
213 Uperodon variegatus Microhylidae: Microhylinae India-Seychelles
214 Uperodon systoma Microhylidae: Microhylinae India-Seychelles
215 Otophryne pyburni Microhylidae: Otophryninae South America
216 Synapturanus sp. Microhylidae: Otophryninae South America
217 Phrynomantis microps Microhylidae: Phrynomerinae Africa
218 Paradoxophyla palmata Microhylidae: Scaphiophryninae Madagascar
219 Scaphiophryne boribory Microhylidae: Scaphiophryninae Madagascar
220 Scaphiophryne madagascariensis Microhylidae: Scaphiophryninae Madagascar
221 Scaphiophryne marmorata Microhylidae: Scaphiophryninae Madagascar
222 Limnodynastes salmini Myobatrachidae: Limnodynastinae Australia
223 Mixophyes coggeri Myobatrachidae: Limnodynastinae Australia
223 Mixophyes coggeri Myobatrachidae: Limnodynastinae Australia
224 Crinia signifera Myobatrachidae: Myobatrachinae Australia
225 Myobatrachus gouldii Myobatrachidae: Myobatrachinae Australia
226 Uperoleia laevigata Myobatrachidae: Myobatrachinae Australia
227 Nasikabatrachus sahyadrensis Nasikabatrachidae India-Seychelles
228 Lankanectes corrugatus Nyctibatrachidae India-Seychelles
229 Nyctibatrachus sp. Nyctibatrachidae India-Seychelles
230 Odontobatrachus natator Odontobatrachidae Africa
231 Odontophrynus occidentalis Odontophrynidae South America
232 Proceratophrys boiei Odontophrynidae South America
233 Pelobates cultripes Pelobatidae Eurasia
234 Pelobates syriacus Pelobatidae Eurasia
235 Pelodytes ibericus Pelodytidae Eurasia
236 Petropedetes euskircheni Petropedetidae Africa
237 Petropedetes parkeri Petropedetidae Africa
238 Phrynobatrachus africanus Phrynobatrachidae Africa
239 Phrynobatrachus krefftii Phrynobatrachidae Africa
240 Phrynobatrachus natalensis Phrynobatrachidae Africa
241 Hymenochirus boettgeri Pipidae Africa
242 Pseudhymenochirus merlini Pipidae Africa
243 Xenopus epitropicalis Pipidae Africa
244 Xenopus kobeli Pipidae Africa
245 Pipa parva Pipidae South America
246 Pipa pipa Pipidae South America
247 Ptychadena cooperi Ptychadenidae Africa
248 Ptychadena mascareniensis Ptychadenidae Africa
249 Ptychadena oxyrhynchus Ptychadenidae Africa
250 Amietia lubrica Pyxicephalidae: Cacosterninae Africa
251 Strongylopus grayii Pyxicephalidae: Cacosterninae Africa
252 Aubria subsigillata Pyxicephalidae: Pyxicephalinae Africa
253 Pyxicephalus edulis Pyxicephalidae: Pyxicephalinae Africa
254 Amolops loloensis Ranidae Eurasia
255 Amolops ricketti Ranidae Eurasia
256 Babina chapaensis Ranidae Eurasia
257 Babina pleuraden Ranidae Eurasia
258 Sylvirana guentheri Ranidae Eurasia
259 Papurana latouchii Ranidae Eurasia
260 Hydrophylax leptoglossa Ranidae Eurasia
261 Chalcorana macrops Ranidae Eurasia
262 Papurana sp. Ranidae Eurasia
263 Meristogenys kinabaluensis Ranidae Eurasia
264 Odorrana hosii Ranidae Eurasia
265 Odorrana schmackeri Ranidae Eurasia
266 Pelophylax nigromaculatus Ranidae Eurasia
267 Rana amurensis Ranidae Eurasia
268 Rana chensinensis Ranidae Eurasia
269 Rana japonica Ranidae Eurasia
270 Rana temporaria Ranidae Eurasia
271 Sanguirana luzonensis Ranidae Eurasia
272 Staurois latopalmatus Ranidae Eurasia
273 Rana catesbeiana Ranidae North America
274 Rana draytonii Ranidae North America
275 Rana pipiens Ranidae North America
276 Rana virgatipes Ranidae North America
277 Rana berlandieri Ranidae North America, South America
278 Indirana sp. A Ranixalidae India-Seychelles
279 Indirana sp. B Ranixalidae India-Seychelles
280 Buergeria buergeri Rhacophoridae: Buergeriinae Eurasia
281 Buergeria oxycephala Rhacophoridae: Buergeriinae Eurasia
282 Kurixalus odontotarsus Rhacophoridae: Rhacophorinae Eurasia
283 Polypedates megacephalus Rhacophoridae: Rhacophorinae Eurasia
284 Rhacophorus dennysi Rhacophoridae: Rhacophorinae Eurasia
285 Pseudophilautus wynaadensis Rhacophoridae: Rhacophorinae India-Seychelles
286 Insuetophrynus acarpicus Rhinodermatidae South America
287 Rhinoderma darwinii Rhinodermatidae South America
288 Rhinophrynus dorsalis Rhinophrynidae North America
289 Scaphiopus couchii Scaphiopodidae North America
290 Scaphiopus holbrookii Scaphiopodidae North America
291 Spea intermontana Scaphiopodidae North America
292 Spea multiplicata Scaphiopodidae North America
293 Sooglossus thomasseti Sooglossidae India-Seychelles
294 Barycholos pulcher Strabomantidae South America
295 Hypodactylus brunneus Strabomantidae South America
296 Phrynopus bracki Strabomantidae South America
297 Pristimantis thymelensis Strabomantidae South America
298 Strabomantis biporcatus Strabomantidae South America
299 Strabomantis sulcatus Strabomantidae South America
300 Batrachophrynus macrostomus Telmatobiidae South America
301 Telmatobius vellardi Telmatobiidae South America
302 Andrias davidianus Cryptobranchidae Eurasia
303 Batrachuperus yenyuanensis Hynobiidae Eurasia

Supplementary Material

Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File

Acknowledgments

This work was supported by National Natural Science Foundation of China Grants 31672266 and 31372172 (to P.Z.), National Youth Talent Support Program Grant W02070133 (to P.Z.), National Science Fund for Excellent Young Scholars of China Grant 31322049 (to P.Z.), and National Science Foundation Grant DEB-1202609 (to D.C.B.) and DEB-1441652 (to D.B.W.).

Footnotes

The authors declare no conflict of interest.

Data deposition: The sequences reported in this paper have been deposited in the GenBank database. For a list of accession numbers, see Dataset S1.

This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1704632114/-/DCSupplemental.

References

  • 1.University of California, Berkeley 2017 AmphibiaWeb: Information on amphibian biology and conservation. Available at amphibiaweb.org. Accessed June 14, 2017.
  • 2.Roelants K, Bossuyt F. Archaeobatrachian paraphyly and Pangaean diversification of crown-group frogs. Syst Biol. 2005;54:111–126. doi: 10.1080/10635150590905894. [DOI] [PubMed] [Google Scholar]
  • 3.San Mauro D, Vences M, Alcobendas M, Zardoya R, Meyer A. Initial diversification of living amphibians predated the breakup of Pangaea. Am Nat. 2005;165:590–599. doi: 10.1086/429523. [DOI] [PubMed] [Google Scholar]
  • 4.Frost DR, et al. The amphibian tree of life. Bull Am Mus Nat Hist. 2006;297:1–370. [Google Scholar]
  • 5.Roelants K, et al. Global patterns of diversification in the history of modern amphibians. Proc Natl Acad Sci USA. 2007;104:887–892. doi: 10.1073/pnas.0608378104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Pyron RA, Wiens JJ. A large-scale phylogeny of Amphibia including over 2800 species, and a revised classification of extant frogs, salamanders, and caecilians. Mol Phylogenet Evol. 2011;61:543–583. doi: 10.1016/j.ympev.2011.06.012. [DOI] [PubMed] [Google Scholar]
  • 7.Zhang P, et al. Efficient sequencing of anuran mtDNAs and a mitogenomic exploration of the phylogeny and evolution of frogs. Mol Biol Evol. 2013;30:1899–1915. doi: 10.1093/molbev/mst091. [DOI] [PubMed] [Google Scholar]
  • 8.Bossuyt F, Roelants K. Anura. In: Hedges SB, Kumar S, editors. The Timetree of Life. Oxford Univ Press; New York: 2009. pp. 357–364. [Google Scholar]
  • 9.Wiens JJ. Re-evolution of lost mandibular teeth in frogs after more than 200 million years, and re-evaluating Dollo’s law. Evolution. 2011;65:1283–1296. doi: 10.1111/j.1558-5646.2011.01221.x. [DOI] [PubMed] [Google Scholar]
  • 10.Frazão A, da Silva HR, Russo CA. The Gondwana breakup and the history of the Atlantic and Indian oceans unveils two new clades for early neobatrachian diversification. PLoS ONE. 2015;10:e0143926. doi: 10.1371/journal.pone.0143926. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Biju SD, Bossuyt F. New frog family from India reveals an ancient biogeographical link with the Seychelles. Nature. 2003;425:711–714. doi: 10.1038/nature02019. [DOI] [PubMed] [Google Scholar]
  • 12.Shen XX, Liang D, Feng YJ, Chen MY, Zhang P. A versatile and highly efficient toolkit including 102 nuclear markers for vertebrate phylogenomics, tested by resolving the higher level relationships of the Caudata. Mol Biol Evol. 2013;30:2235–2248. doi: 10.1093/molbev/mst122. [DOI] [PubMed] [Google Scholar]
  • 13.Irisarri I, Vences M, San Mauro D, Glaw F, Zardoya R. Reversal to air-driven sound production revealed by a molecular phylogeny of tongueless frogs, family Pipidae. BMC Evol Biol. 2011;11:114. doi: 10.1186/1471-2148-11-114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Bossuyt F, Brown RM, Hillis DM, Cannatella DC, Milinkovitch MC. Phylogeny and biogeography of a cosmopolitan frog radiation: Late cretaceous diversification resulted in continent-scale endemism in the family Ranidae. Syst Biol. 2006;55:579–594. doi: 10.1080/10635150600812551. [DOI] [PubMed] [Google Scholar]
  • 15.Van Bocxlaer I, Roelants K, Biju SD, Nagaraju J, Bossuyt F. Late Cretaceous vicariance in Gondwanan amphibians. PLoS ONE. 2006;1:e74. doi: 10.1371/journal.pone.0000074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.van der Meijden A, et al. Nuclear gene phylogeny of narrow-mouthed toads (family: Microhylidae) and a discussion of competing hypotheses concerning their biogeographical origins. Mol Phylogenet Evol. 2007;44:1017–1030. doi: 10.1016/j.ympev.2007.02.008. [DOI] [PubMed] [Google Scholar]
  • 17.Kurabayashi A, et al. From Antarctica or Asia? New colonization scenario for Australian-New Guinean narrow mouth toads suggested from the findings on a mysterious genus Gastrophrynoides. BMC Evol Biol. 2011;11:175. doi: 10.1186/1471-2148-11-175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.de Sá RO, et al. Molecular phylogeny of microhylid frogs (Anura: Microhylidae) with emphasis on relationships among New World genera. BMC Evol Biol. 2012;12:241. doi: 10.1186/1471-2148-12-241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Peloso PLV, et al. The impact of anchored phylogenomics and taxon sampling on phylogenetic inference in narrow-mouthed frogs (Anura, Microhylidae) Cladistics. 2016;32:113–140. doi: 10.1111/cla.12118. [DOI] [PubMed] [Google Scholar]
  • 20.Heinicke MP, et al. A new frog family (Anura: Terrarana) from South America and an expanded direct-developing clade revealed by molecular phylogeny. Zootaxa. 2009;2211:1–35. [Google Scholar]
  • 21.Darst CR, Cannatella DC. Novel relationships among hyloid frogs inferred from 12S and 16S mitochondrial DNA sequences. Mol Phylogenet Evol. 2004;31:462–475. doi: 10.1016/j.ympev.2003.09.003. [DOI] [PubMed] [Google Scholar]
  • 22.Hedges SB, Duellman WE, Heinicke MP. New World direct-developing frogs (Anura: Terrarana): Molecular phylogeny, classification, biogeography, and conservation. Zootaxa. 2008;1737:1–182. [Google Scholar]
  • 23.Pyron RA. Biogeographic analysis reveals ancient continental vicariance and recent oceanic dispersal in amphibians. Syst Biol. 2014;63:779–797. doi: 10.1093/sysbio/syu042. [DOI] [PubMed] [Google Scholar]
  • 24.Pyron RA. Divergence time estimation using fossils as terminal taxa and the origins of Lissamphibia. Syst Biol. 2011;60:466–481. doi: 10.1093/sysbio/syr047. [DOI] [PubMed] [Google Scholar]
  • 25.Renne PR, et al. Time scales of critical events around the Cretaceous-Paleogene boundary. Science. 2013;339:684–687. doi: 10.1126/science.1230492. [DOI] [PubMed] [Google Scholar]
  • 26.Keller G. Deccan volcanism, the Chicxulub impact, and the end-Cretaceous mass extinction: Coincidence? Cause and effect? Geol Soc Am. 2014;505:57–89. [Google Scholar]
  • 27.Duellman WE, Marion AB, Hedges SB. Phylogenetics, classification, and biogeography of the treefrogs (Amphibia: Anura: Arboranae) Zootaxa. 2016;4104:1–109. doi: 10.11646/zootaxa.4104.1.1. [DOI] [PubMed] [Google Scholar]
  • 28.Shen XX, et al. Enlarged multilocus data set provides surprisingly younger time of origin for the Plethodontidae, the largest family of salamanders. Syst Biol. 2016;65:66–81. doi: 10.1093/sysbio/syv061. [DOI] [PubMed] [Google Scholar]
  • 29.Longrich NR, Bhullar BAS, Gauthier JA. Mass extinction of lizards and snakes at the Cretaceous-Paleogene boundary. Proc Natl Acad Sci USA. 2012;109:21396–21401. doi: 10.1073/pnas.1211526110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Longrich NR, Tokaryk T, Field DJ. Mass extinction of birds at the Cretaceous-Paleogene (K-Pg) boundary. Proc Natl Acad Sci USA. 2011;108:15253–15257. doi: 10.1073/pnas.1110395108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.O’Leary MA, et al. The placental mammal ancestor and the post-K-Pg radiation of placentals. Science. 2013;339:662–667. doi: 10.1126/science.1229237. [DOI] [PubMed] [Google Scholar]
  • 32.Cooper A, Penny D. Mass survival of birds across the Cretaceous-Tertiary boundary: Molecular evidence. Science. 1997;275:1109–1113. doi: 10.1126/science.275.5303.1109. [DOI] [PubMed] [Google Scholar]
  • 33.Benton MJ. The origins of modern biodiversity on land. Philos Trans R Soc Lond B Biol Sci. 2010;365:3667–3679. doi: 10.1098/rstb.2010.0269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Meredith RW, et al. Impacts of the Cretaceous Terrestrial Revolution and KPg extinction on mammal diversification. Science. 2011;334:521–524. doi: 10.1126/science.1211028. [DOI] [PubMed] [Google Scholar]
  • 35.dos Reis M, et al. Phylogenomic datasets provide both precision and accuracy in estimating the timescale of placental mammal phylogeny. Proc R Soc B Biol Sci. 2012;279:3491–3500. doi: 10.1098/rspb.2012.0683. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.dos Reis M, Donoghue PCJ, Yang Z. Neither phylogenomic nor palaeontological data support a Palaeogene origin of placental mammals. Biol Lett. 2014;10:20131003. doi: 10.1098/rsbl.2013.1003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Blain HA, Canudo JI, Cuenca-Bescós G, López-Martínez N. Amphibians and squamate reptiles from the latest Maastrichtian (Upper Cretaceous) of Blasi 2 (Huesca, Spain) Cretac Res. 2010;31:433–446. [Google Scholar]
  • 38.Fastovsky DE, Bercovici A. The Hell Creek Formation and its contribution to the Cretaceous–Paleogene extinction: A short primer. Cretac Res. 2016;57:368–390. [Google Scholar]
  • 39.Marjanović D, Laurin M. An updated paleontological timetree of lissamphibians, with comments on the anatomy of Jurassic crown-group salamanders (Urodela) Hist Biol. 2014;26:535–550. [Google Scholar]
  • 40.Mercier GK, Demar DG, Wilson GP. 2016. Anurans, caudates, and albanerpetontids (Lissamphibia) reveal differential patterns of turnover and extinctions during the end-Cretaceous mass extinction, northeastern Montana, USA. 76th Annual Meeting of the Society of Vertebrate Paleontology, October 26–29, 2016 (Salt Lake City), p 188.
  • 41.Matzke NJ. 2013 BioGeoBEARS: Biogeography with Bayesian and likelihood evolutionary analysis in R scripts. Available at cran.r-project.org/web/packages/BioGeoBEARS/. Accessed February 26, 2016.
  • 42.Blakey RC. Gondwana paleogeography from assembly to breakup—a 500 m.y. odyssey. GSA Special Papers. 2008;441:1–28. [Google Scholar]
  • 43.Plummer PS, Belle ER. Mesozoic tectono-stratigraphic evolution of the Seychelles microcontinent. Sediment Geol. 1995;96:73–91. [Google Scholar]
  • 44.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. doi: 10.1126/science.1059412. [DOI] [PubMed] [Google Scholar]
  • 45.Francis JE, Poole I. Cretaceous and early Tertiary climates of Antarctica: Evidence from fossil wood. Palaeogeogr Palaeoclimatol Palaeoecol. 2002;182:47–64. [Google Scholar]
  • 46.Goin CJ, Goin OB. Antarctica, isostacy, and the origin of frogs. Q J Florida Acad Sci. 1973;35:113–129. [Google Scholar]
  • 47.Savage JM. The geographic distribution of frogs: Patterns and predictions. In: Vial JL, editor. Evolutionary Biology of the Anurans: Contemporary Research on Major Problems. Univ Missouri Press; Columbia, MO: 1973. pp. 351–445. [Google Scholar]
  • 48.Bossuyt F, Milinkovitch MC. Amphibians as indicators of early tertiary “out-of-India” dispersal of vertebrates. Science. 2001;292:93–95. doi: 10.1126/science.1058875. [DOI] [PubMed] [Google Scholar]
  • 49.Ruddiman W. Early uplift in Tibet? Nature. 1998;394:723–725. [Google Scholar]
  • 50.Vajda V, Raine JI, Hollis CJ. Indication of global deforestation at the Cretaceous-Tertiary boundary by New Zealand fern spike. Science. 2001;294:1700–1702. doi: 10.1126/science.1064706. [DOI] [PubMed] [Google Scholar]
  • 51.Feng YJ, Liu QF, Chen MY, Liang D, Zhang P. Parallel tagged amplicon sequencing of relatively long PCR products using the Illumina HiSeq platform and transcriptome assembly. Mol Ecol Resour. 2016;16:91–102. doi: 10.1111/1755-0998.12429. [DOI] [PubMed] [Google Scholar]
  • 52.Tamura K, Stecher G, Peterson D, Filipski A, Kumar S. MEGA6: Molecular evolutionary genetics analysis version 6.0. Mol Biol Evol. 2013;30:2725–2729. doi: 10.1093/molbev/mst197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Castresana J. Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Mol Biol Evol. 2000;17:540–552. doi: 10.1093/oxfordjournals.molbev.a026334. [DOI] [PubMed] [Google Scholar]
  • 54.Lanfear R, Calcott B, Ho SYW, Guindon S. PartitionFinder: Combined selection of partitioning schemes and substitution models for phylogenetic analyses. Mol Biol Evol. 2012;29:1695–1701. doi: 10.1093/molbev/mss020. [DOI] [PubMed] [Google Scholar]
  • 55.Stamatakis A. RAxML version 8: A tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics. 2014;30:1312–1313. doi: 10.1093/bioinformatics/btu033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Ronquist F, et al. MrBayes 3.2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Syst Biol. 2012;61:539–542. doi: 10.1093/sysbio/sys029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Mirarab S, et al. ASTRAL: Genome-scale coalescent-based species tree estimation. Bioinformatics. 2014;30:i541–i548. doi: 10.1093/bioinformatics/btu462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Yang Z. PAML 4: Phylogenetic analysis by maximum likelihood. Mol Biol Evol. 2007;24:1586–1591. doi: 10.1093/molbev/msm088. [DOI] [PubMed] [Google Scholar]
  • 59.Benton MJ, et al. Constraints on the timescale of animal evolutionary history. Paleontol Electron. 2015;18:1–106. [Google Scholar]
  • 60.Rabosky DL. Automatic detection of key innovations, rate shifts, and diversity-dependence on phylogenetic trees. PLoS ONE. 2014;9:e89543. doi: 10.1371/journal.pone.0089543. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Rabosky DL, et al. BAMMtools: An R package for the analysis of evolutionary dynamics on phylogenetic trees. Methods Ecol Evol. 2014;5:701–707. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File
Supplementary File

Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES