Introduction

advertisement
1
Journal of Biogeography
SUPPORTING INFORMATION
Diversity versus disparity and the role of ecological opportunity in a continental bird
radiation
Manuel Schweizer, Stefan T. Hertwig and Ole Seehausen
Appendix S1 Phylogenetic analyses and species sampled.
Materials and methods
Nucleotide sequences of 133 of 164 described species were collated from GenBank. Our
sampling thus included more than 81% of all described Arini species and all of the 29 recognized genera following the taxonomy of Forshaw (2010), except the monotypic genus Ognorhynchus (Table S1). We moreover included four taxa of the African Psittacini (Psittacus
erithacus and three species of the genus Poicephalus), the sister group of Arini (Schweizer et
al., 2010, 2011). Melopsittacus undulatus was used as an outgroup. We used partial sequences of the comparatively slowly evolving nuclear exon RAG-1 (Groth & Barrowclough,
1999) and of the faster evolving mitochondrial genes COI (cytochrome c oxidase subunit I),
cytb (cytochrome b) and ND2 (NADH dehydrogenase subunit 2) (Table S1). We used BEAST
1.7.5 (Drummond & Rambaut, 2007) to simultaneously estimate divergence times and establish a phylogenetic hypothesis for Arini. We applied a relaxed molecular clock with an uncorrelated lognormal distribution of branch lengths and a Yule tree prior. We used PARTITIONFINDER 1.0.1 (Lanfear et al., 2012) to select the best-fitting partitioning scheme and models
of nuclear evolution using the greedy algorithm with unlinked branch lengths corresponding
to separate models with varying base frequencies, rate matrix, shape parameters and proportion of invariable sites for the different genes and/or their codon positions. The Bayesian
information criterion (BIC) was used as optimality criterion for model selection. We incorporated a secondary calibration point in our dating analyses based on the results of Schweizer et
al. (2011): a mean estimate of 35 Ma (millions of years ago) for the split between Arini and
Psittacini (based on a normal distribution with a standard deviation of 6.5 Ma). The 95% confidence interval of this normal distribution included the 95% highest posterior density (HPD)
interval for this split as estimated in Schweizer et al. (2011). Default prior distributions were
chosen for all other parameters and we ran three independent chains of Markov chain Monte
Carlo (MCMC) with 25 million generations sampled every 1000 generations. TRACER 1.5
(Rambaut & Drummond, 2007) was used to confirm appropriate burn-in, confirm that the
effective sample sizes of the posterior distribution for each run were adequate, and to assess
convergence among runs by comparing the likelihoods and posterior distributions of all
parameters. The resulting trees and 95% posterior distribution of each estimated node were
analysed with FIGTREE 1.2.1 (Rambaut, 2008). The outgroup and taxa of the African Psittacini were pruned from the trees before further analyses.
We moreover employed a maximum-likelihood search using RAXML 7.0.4 (Stamatakis, 2006) on the web server with 100 rapid bootstrap inferences (Stamatakis et al., 2008),
with all free model parameters estimated by the software (substitution rates, gamma shape
2
parameter and base frequencies), based on the best partitioning scheme identified by PARTITIONFINDER (see above).
In the MEDUSA analysis, species missing from our sample, and for which no genetic
data were available, were added to terminal branches using taxonomic information or plumage similarities taken from various sources (Collar, 1997; Arndt, 2006; Forshaw, 2010). Amazona mercenaria was added to the terminal branch leading to Amazona farinosa, and Amazona xantholora was added to Amazona albifrons. Ara ambiguus, Ara militaris and Ara rubrogenys were added to the terminal branch leading to Ara severus. Aratinga frontalis had been
formerly been treated as a subspecies of Aratinga wagleri (Collar, 1997) and was hence
added to the terminal branch leading to the latter. Aratinga erythrogenys was also included in
Aratinga wagleri. Aratinga hockingi is part of the Aratinga mitrata complex (Arndt, 2006)
and was added to the terminal branch leading to the latter. Aratinga strenua was added to
Aratinga holochlora, of which it has previously been treated as a subspecies (Collar, 1997).
The three missing species in the genus Bolborhynchus (B. auriforns, B. orbignesius and
B. ferrugineifrons) were included in Bolborhynchus lineola. Enicognathus ferrugineus was
added to the terminal branch leading to Enicognathus leptorhynchus. Ognorhynchus icterotis
was added to Rhynchopsitta pachyrhyncha. Pyrrhura calliptera was included in Pyrrhura
lepida. Pyrrhura caeruleiceps, Pyrrhura subandida and Pyrrhura lucianii were formerly
treated as subspecies of Pyrrhura picta and were added to the latter’s terminal branch.
Pyrrhura devillei forms a superspecies with P. frontalis and was hence added to the latter.
Pyrrhura egregia and Pyrrhura hoematotis were added to the terminal branch leading to
Pyrrhura lepida. Pyrrhura parvifrons was included in Pyrrhura roseifrons. Pyrrhura viridicata was included in Pyrrhura rhodocephala. Rhynchopsitta terresi was added to the terminal branch leading to Rhynchopsitta pachyrhyncha. The missing species in the genus Touit
(T. costaricensis, T. dilectissima, T. melanotus, T. huetii, T. stictoptera and T. surda) were
added to Touit batavica.
Phylogenetic results
The final alignment was 5506 bp long and consisted of 622 bp of COI, 1140 bp of cytb,
1041 bp of ND2 and 2703 bp of RAG-1. It contained no indels. PARTITIONFINDER identified
two subsets as the best-fitting partitioning scheme with a HKY+I+G substitution model for
the first and second codon positions of the three mitochondrial genes as well as for the nuclear exon, and a GTR+I+G substitution model for the third codon position of the three mitochondrial genes. The comparison of the three independent runs with BEAST revealed high
convergence among the different parameters and effective sample sizes were > 250 for all
parameters. The three runs were combined, with a burn-in of 2.5 million generations for each.
The maximum clade credibility tree was then estimated from 67,503 trees. The topology of
the best-scoring tree of the ML analysis with RAXML was highly congruent with the maximum clade credibility tree of the BEAST analysis and there was no conflict between supported
nodes (Fig. S1). They only differed in the positions of Amazona agilis, Amazona albifrons,
Amazona barbadensis, Amazona farinosa, Amazona kawalli, Aratinga brevipes, Aratinga
finschii, Brotogeris sanctithomae, Brotogeris tirica, Pyrrhura rupicola and the clade consisting of Amazona xanthops and Graydidascalus brachyurus. Moreover, two basal clades
3
within Aratinga were not revealed as monophyletic in the ML analyses, in contrast to the
BEAST analyses (Fig. 1, Fig. S1).
Comparison of topology with recent molecular phylogenies of Neotropical parrots
We compared the topology of the maximum clade credibility tree from the BEAST analyses
with recently published molecular phylogenies of Arini, although comparisons, especially of
the basal splits, are hampered by the fact that no previous work used a comparably broad
taxon sampling. The phylogenetic relationships found among genera were highly congruent
with the supported nodes in Tavares et al. (2006), with the exception of the clade consisting
of Brotogeris and Myopsitta, which was there revealed as the sister group to the taxa in our
clade 5. There was also no conflict with the supported nodes in Wright et al. (2008), except
the position of the clade consisting of Deroptyus and Pionites; this was the sister group to the
remaining taxa of clade 2 in our study, whereas in Wright et al. (2008) it was nested within
the taxa in our clade 2.
Within Brotogeris, we recovered the same basal clades as Ribas et al. (2009) and
there was no conflict between supported nodes. The topology within Pyrrhura was also
broadly congruent with the results of Ribas et al. (2006) for supported nodes, with the exception of the positions of P. picta and P. emma. The relationships among the analysed species
of Aratinga were congruent with Ribas & Miyaki (2004) and Kirchman et al. (2012). Some
differences from Smith et al. (2013) were revealed in the topology within Forpus. That study,
however, included different subspecies and, in some cases, several samples per species.
Additionally, the relationships found within Pionus were congruent with the supported nodes
in Ribas et al. (2007). Amazona xanthops did not cluster with the other members of Amazona
as also observed in other studies (Russello & Amato, 2004; Tavares et al., 2006). The relationships within Amazona were found to be consistent overall with the supported nodes in
Russello & Amato (2004), with the exception of the position of A. amazonica. The relationships we found within Gypopsitta were identical to those in Ribas et al. (2005). As in that
study, Hapalopsittaca was found to be the sister group to Gypopsitta.
Testing the assignment of unsampled Pyrrhura species in the MEDUSA analysis
The clade consisting of all Pyrrhura species except Pyrrhura cruentata was found to be unexpectedly species-rich in the MEDUSA analysis. Nine unsampled Pyrrhura species were
added to terminal branches of the sampled species contained in this clade based on taxonomic
information and plumage similarities (see above), which may have influenced this result. We
therefore collapsed this clade and added the missing species either to this clade or to the terminal branch leading to Pyrrhura cruentata, although the available taxonomic evidence (see
above) makes it unlikely that they are more closely related to this species than to the remaining Pyrrhura species. When 5–9 of the unsampled species were added to Pyrrhura cruentata
(the remaining 0–4 unsampled species were included in the other Pyrrhura lineage),
MEDUSA found a rate shift on the branch leading to all Pyrrhura species. When fewer than
five of the unsampled species were added to the terminal branch of Pyrrhura cruentata, the
rate shift was identified on the branch leading to the remaining Pyrrhura species as in the
initial analyses. Hence, the allocation of the unsampled species had only a marginal effect on
the overall results.
4
Table S1 Species sampled and GenBank accession numbers for the four genes analysed for this study.
Amazona aestiva
Amazona agilis
Amazona albifrons
Amazona amazonica
Amazona arausiaca
Amazona autumnalis
Amazona barbadensis
Amazona brasiliensis
Amazona collaria
Amazona dufresniana
Amazona farinosa
Amazona festiva
Amazona finschi
Amazona guildingii
Amazona imperialis
Amazona kawalli
Amazona leucocephala
Amazona ochrocephala ochrocephala
Amazona pretrei
Amazona rhodocorytha
Amazona tucumana
Amazona ventralis
Amazona versicolor
Amazona vinacea
Amazona viridigenalis
Amazona vittata
Amazona xanthops
Anodorhynchus hyacinthinus
Anodorhynchus leari
Ara ararauna
Ara chloropterus
Ara glaucogularis
Ara macao
Ara severus
Aratinga acuticaudata
Aratinga aurea
Aratinga auricapillus
Aratinga branickii
Aratinga brevipes
Aratinga cactorum
Aratinga canicularis
Aratinga chloroptera
Aratinga euops
Aratinga finschi
Aratinga holochlora
Aratinga jandaya
Aratinga leucophthalma
Aratinga mitrata
COI
cytb
ND2
RAG1
EU340705
AY301426
AY301427
AY194399
AY301431
AY194379
AY301436
AY301437
AY301438
AY301439
AY301443
AY301444
AY301445
AY301446
AY301447
JQ235567
AY301449
AY194397
AY301457
AY301458
AY301459
AY301460
AY301461
AY301462
AY301463
AY301464
AY301465
AF370738
AF370736
AY301467
—
GU826174
EU621598
AF370748
NC_020325
AF370746
AY301466
EU621626
GU826177
AF370750
HQ629753
GU826178
GU826179
GU826180
GU826181
—
AF370749
GU826183
AY194404
AY283489
AY283506
AY283516
AY283465
AY283456
AY283462
—
AY283493
AY283454
FJ899168
—
AY283461
AY283460
AY283458
JQ235589
AY283518
AY194411
—
—
—
AY283474
AY283466
—
—
AY283514
DQ143293
AF346370
AF370763
DQ150994
AF346367
—
AF346366
AF370766
NC_020325
U70762
AY208239
—
—
AY281254
—
—
—
—
—
AY208244
AY281253
—
AY194434
—
HQ629715
AY194466
—
AY194446
—
—
—
—
AY194461
—
—
—
—
—
—
AY194460
—
—
—
—
—
—
EU327598
—
AY669485
DQ143311
AY669446
DQ143315
—
HQ270481
EU327601
—
NC_020325
DQ143321
HQ270483
EU327630
—
—
HQ629718
HQ270484
HQ270485
HQ270486
HQ270487
—
Q143298
HQ270489
JF807980
—
—
—
—
—
—
—
—
JF807981
DQ143346
—
—
—
—
—
—
—
JF807982
—
—
—
—
—
—
—
DQ143345
DQ143329
—
—
—
DQ143340
—
—
DQ143341
—
—
—
—
—
—
—
—
—
—
DQ143331
—
5
Aratinga nana
Aratinga nenday
Aratinga pertinax
Aratinga rubritorquis
Aratinga solstitialis
Aratinga wagleri
Aratinga weddelli
Bolborhynchus aymara
Bolborhynchus lineola
Brotogeris chiriri
Brotogeris chrysoptera
Brotogeris cyanoptera
Brotogeris jugularis
Brotogeris pyrrhoptera
Brotogeris sanctithomae
Brotogeris tirica
Brotogeris versicolurus
Cyanoliseus patagonus
Cyanopsitta spixii
Deroptyus accipitrinus
Diopsittaca nobilis
Enicognathus leptorhynchus
Forpus coelestis
Forpus conspicillatus
Forpus cyanopygius
Forpus passerinus
Forpus sclateris(=modestus)
Forpus xanthops
Forpus xanthopterygius
Graydidascalus brachyurus
Guaruba guarouba
Gypopsitta aurantiocephala
Gypopsitta barrabandi
Gypopsitta caica
Gypopsitta haematotis
Gypopsitta pulchra
Gypopsitta pyrilia
Gypopsitta vulturina
Hapalopsittaca amazonina
Hapalopsittaca fuertesi
Hapalopsittaca melanotis
Hapalopsittaca pyrrhops
Melopsittacus undulatus
Myiopsitta monachus
Nannopsittaca dachilleae
Nannopsittaca panychlora
Orthopsittaca manilata
Pionites leucogaster
Pionites melanocephalus
COI
cytb
ND2
RAG1
GU826184
EU621632
EU621597
—
GU826185
GU826186
—
HQ629774
GU826187
—
HQ629756
NC_015530
EU621601
—
—
—
AF370755
GQ232191
EU621610
EU621613
AF370752
EU621616
—
—
—
EU621621
—
—
—
AY301468
AF370741
—
AY661216
AY661221
AY661225
AY661231
—
AY661213
EU621625
—
GU826192
—
EU327633
EU621631
—
EU621633
EU621640
—
EU621644
—
AY219915
JX877361
—
DQ143285
—
AY669401
—
DQ143291
FJ652859
FJ652893
FJ652871
FJ652910
FJ652865
FJ652902
FJ652849
FJ652853
GQ232267
DQ823368
DQ150992
AF370769
—
JX877284
JX877337
JX877356
JX877308
JX877272
JX877314
DQ143294
AY669439
AF346371
AY669407
AY669424
AY669419
AY669433
AY669427
AY669426
AY669412
AY669443
JN393252
—
JN393246
EU621629
AF346378
—
—
DQ150991
DQ143288
—
FJ361233
AY274066
EU327600
JX477130
DQ143317
HQ270492
AY669445
HQ629738
DQ143319
FJ652922
FJ652957
FJ652938
FJ652973
FJ652928
FJ652965
FJ652911
FJ652917
EU327613
EU327614
EU327617
EU327618
EU327620
JX877369
JX877387
JX877421
EU327625
JX877411
JX877378
DQ143300
AY669484
DQ143309
AY669451
AY669468
EU327649
AY669477
AY669472
AY669470
AY669453
EU327629
JN393236
HQ270498
JN393230
DQ143295
EU327635
DQ143320
EU327637
EU327644
DQ143312
EU327648
—
DQ143326
—
—
DQ143330
—
—
—
DQ143350
DQ143348
—
—
—
—
—
—
—
DQ143334
DQ143336
DQ143338
DQ143335
DQ143332
—
—
—
—
—
—
DQ143325
DQ143344
DQ143333
—
DQ143349
—
—
—
—
—
—
—
—
—
DQ143354
DQ143328
DQ143352
—
DQ143337
DQ143342
—
6
Pionopsitta pileata
Pionus chalcopterus
Pionus fuscus
Pionus maximiliani
Pionus menstruus
Pionus senilis
Pionus seniloides
Pionus sordidus
Pionus tumultuosus
Poicephalus gulielmi
Poicephalus rufiventris
Poicephalus senegalus
Primolius auricollis
Primolius couloni
Primolius maracana
Psittacus erithacus
Pyrrhura albipectus
Pyrrhura amazonum
Pyrrhura cruentata
Pyrrhura eisenmanni
Pyrrhura emma
Pyrrhura frontalis
Pyrrhura griseipectus
Pyrrhura hoffmanni
Pyrrhura lepida
Pyrrhura leucotis
Pyrrhura melanura
Pyrrhura molinae
Pyrrhura orcesi
Pyrrhura perlata
Pyrrhura peruviana
Pyrrhura pfrimeri
Pyrrhura picta
Pyrrhura rhodocephala
Pyrrhura roseifrons
Pyrrhura rupicola
Pyrrhura snethlageae
Rhynchopsitta pachyrhyncha
Touit batavicus
Touit purpuratus
Triclaria malachitacea
COI
cytb
ND2
RAG1
AY661227
AY661232
—
—
AY301469
GU826194
—
—
—
—
—
HQ629771
AF370753
AF370762
AF370754
EU621657
HQ629779
—
—
—
—
—
—
HQ629780
—
—
—
—
—
GU826196
—
—
EU621660
—
—
—
—
EU621661
EU621666
HQ629781
HQ629783
AY669438
EF517621
EF517636
EF517622
EF517611
EF517618
EF517616
EF517634
EF517616
AY283498
—
—
DQ150995
AF370780
AF370775
AY082076
AY751640
AY751616
AY751658
AY751599
AY751626
AY751643
AY751630
AY751654
AY751645
AY751633
AY751652
AY751642
AY751636
AY751646
AY751583
AY751620
AY751602
AY751638
AY751596
AY751656
AY751612
DQ143297
—
—
AY669442
AY669479
EF517656
EF517671
EF517662
EF517637
EF517653
EF517651
EF517669
EF517649
—
—
—
DQ143302
HQ270501
—
GU816826
—
—
—
—
—
—
—
HQ629744
—
DQ143307
—
—
—
HQ270502
—
—
AY669444
—
—
—
—
DQ143303
EU327670
HQ629745
AY669486
DQ143351
—
—
DQ143347
JF807987
—
—
—
—
JF807988
GQ505226
GQ505227
DQ143343
—
—
EF517674
—
—
—
—
—
AY233360
—
—
—
DQ143353
—
—
—
—
—
—
—
—
—
—
—
DQ143339
—
—
DQ143327
7
Figure S1 Best-scoring maximum likelihood tree using RAXML for Neotropical parrots (Arini) and outgroup
species. Supported nodes (bootstrap values ≥ 70%) are marked with black circles, while moderately supported
nodes (0.5 ≤ posterior probability ≤ 0.7) are marked with grey circles.
8
Appendix S2 Testing the influence of incomplete taxon sampling on diversification rate
estimates.
To test the influence of incomplete taxon sampling and the robustness of our results, we simulated 1000 trees under a constant-rate birth–death model using the parameters as estimated
with a pure-birth model (Rabosky, 2006) based on our data. These phylogenies were simulated to 164 tips representing the current species diversity of Arini and then randomly pruned to
our taxon sampling of 133 species. To control for the potentially phylogenetically overdispersed sampling, we implemented the method of Brock et al. (2011) using different values
for the scaling parameter α. The root node of the simulated trees was then set to 25.719 Ma
(mean value recovered). We then compared the distribution of the differences in AIC scores
between the best-fitting rate-variable and the best-fitting rate-constant model of the simulated
trees and the 1000 last trees of the BEAST analyses with the root node set to 25.719 Ma (see
below). A positive difference in AIC scores indicates that the rate-variable model fits the data
better than the rate-constant model. Only the following models could be compared in the
LASER package: pure birth, constant-rate birth-death, yule2rate, DDX and DDL.
We often found positive differences in AIC scores between the best-fitting variablerate and the best-fitting constant-rate model for the 1000 trees simulated under a birth–death
model (Fig. S2). When completely random or low-levels of overdispersed sampling were
assumed, this null-distribution peaked below zero, but broadly overlapped with the distribution of the differences in AIC score for the 1000 last trees of the BEAST analyses. However,
when higher levels of overdispersed sampling were considered, the null-distribution shifted to
even more positive differences in AIC scores. Hence, the observed distribution cannot be
distinguished from a constant-rate diversification process.
Figure S2 Distribution of ΔAIC between the best-fitting constant-rate and the best-fitting variable-rate
diversification model calculated from the posterior distribution of the BEAST analysis (grey) and from a null
distribution of phylogenies simulated under a constant-rate model (white). We controlled for potentially
phylogenetically overdispersed sampling using different values of the scaling parameter α.
9
Appendix S3 Additional figures on the analyses of morphological evolution.
Figure S3 (a) Maximum clade credibility tree of the dating analysis using BEAST of Neotropical parrots (Arini)
with the colours of branches representing the reconstructed evolution of shape PC1 along lineages.
(b) Disparity-through-time plot for shape PC1 of Arini. The solid line indicates the mean subclade disparity
through time based on the maximum clade credibility tree of the BEAST analyses, while the dashed line shows
the median expected subclade disparity derived from 10,000 simulations under Brownian motion evolution. The
area shaded in grey indicates the 95% range for the simulated data.
10
Figure S4 (a) Maximum clade credibility tree of the dating analysis using BEAST of Neotropical parrots (Arini)
with the colours of branches representing the reconstructed evolution of shape PC2 along lineages.
(b) Disparity-through-time plot for shape PC2 of Arini. The solid line indicates the mean subclade disparity
through time based on the maximum clade credibility tree of the BEAST analyses, while the dashed line shows
the median expected subclade disparity derived from 10,000 simulations under Brownian motion evolution. The
area shaded in grey indicates the 95% range for the simulated data.
11
REFERENCES
Baur, H. & Leuenberger, C. (2011) Analysis of ratios
in multivariate morphometry. Systematic Biology,
60, 813–825.
Brock, C.D., Harmon, L.J. & Alfaro, M.E. (2011)
Testing for temporal variation in diversification
rates when sampling is incomplete and nonrandom. Systematic Biology, 60, 410–419.
Drummond, A.J. & Rambaut, A. (2007) BEAST:
Bayesian evolutionary analysis by sampling trees.
BMC Evolutionary Biology, 7, 214.
Forshaw, J.M. (2010) Parrots of the world: a field
guide. Christopher Helm, London.
Groth, J.G. & Barrowclough, G.F. (1999) Basal
divergences in birds and the phylogenetic utility of
the nuclear RAG-1 gene. Molecular Phylogenetics
and Evolution, 12, 115–123.
Kirchman, J.J., Schirtzinger, E.E. & Wright, T.F.
(2012) Phylogenetic relationships of the extinct
Carolina Parakeet (Conuropsis carolinensis) inferred from DNA sequence data. The Auk, 129, 197–
204.
Lanfear, R., Calcott, B., Ho, S.Y.W. & Guindon, S.
(2012) PartitionFinder: combined selection of
partitioning schemes and substitution models for
phylogenetic analyses. Molecular Biology and
Evolution, 29, 1695–1701.
Rabosky, D.L. (2006) Likelihood methods for detecting temporal shifts in diversification rates. Evolution, 60, 1152–1164.
Rambaut, A. (2008) Figtree 1.2. Available at: http://
tree.bio.ed.ac.uk/software/figtree/.
Rambaut, A. & Drummond, A.J. (2007) Tracer 1.5.
Available at: http://beast.bio.ed.ac.uk/Tracer.
Ribas, C.C., Gaban-Lima, R., Miyaki, C.Y. & Cracraft, J. (2005) Historical biogeography and
diversification within the Neotropical parrot genus
Pionopsitta (Aves: Psittacidae). Journal of Biogeography, 32, 1409–1427.
Ribas, C.C., Joseph, L. & Miyaki, C.Y. (2006)
Molecular systematics and patterns of diversification in Pyrrhura (Psittacidae), with special reference to the picta–leucotis complex. The Auk,
123, 660–680.
Ribas, C.C. & Miyaki, C.Y. (2004) Molecular systematics in Aratinga parakeets: species limits and historical biogeography in the ‘solstitialis’ group, and
the systematic position of Nandayus nenday. Molecular Phylogenetics and Evolution, 30, 663–675.
Ribas, C.C., Miyaki, C.Y. & Cracraft, J. (2009)
Phylogenetic relationships, diversification and bio-
geography in Neotropical Brotogeris parakeets.
Journal of Biogeography, 36, 1712–1729.
Ribas, C.C., Moyle, R.G., Miyaki, C.Y. & Cracraft, J.
(2007) The assembly of montane biotas: linking
Andean tectonics and climatic oscillations to independent regimes of diversification in Pionus
parrots. Proceedings of the Royal Society B: Biological Sciences, 274, 2399–2408.
Russello, M.A. & Amato, G. (2004) A molecular phylogeny of Amazona: implications for Neotropical
parrot biogeography, taxonomy, and conservation.
Molecular Phylogenetics and Evolution, 30, 421–
437.
Schweizer, M., Seehausen, O., Güntert, M. & Hertwig, S.T. (2010) The evolutionary diversification
of parrots supports a taxon pulse model with
multiple trans-oceanic dispersal events and local
radiations. Molecular Phylogenetics and Evolution, 54, 984–994.
Schweizer, M., Seehausen, O. & Hertwig, S.T. (2011)
Macroevolutionary patterns in the diversification
of parrots: effects of climate change, geological
events and key innovations. Journal of Biogeography, 38, 2176–2194.
Smith, B.T., Ribas, C.C., Whitney, B.M., HernándezBaños, B.E. & Klicka, J. (2013) Identifying biases
at different spatial and temporal scales of diversification: a case study in the Neotropical parrotlet
genus Forpus. Molecular Ecology, 22, 483–494.
Stamatakis, A. (2006) RAxML-VI-HPC: Maximum
likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics,
22, 2688–2690.
Stamatakis, A., Hoover, P. & Rougemont, J. (2008) A
rapid bootstrap algorithm for the RAxML web
servers. Systematic Biology, 57, 758–771.
Tavares, E.S., Baker, A.J., Pereira, S.L. & Miyaki,
C.Y. (2006) Phylogenetic relationships and historical biogeography of Neotropical parrots (Psittaciformes: Psittacidae: Arini) inferred from mitochondrial and nuclear DNA sequences. Systematic
Biology, 55, 454–470.
Wright, T.F., Schirtzinger, E.E., Matsumoto, T., Eberhard, J.R., Graves, G.R., Sanchez, J.J., Capelli, S.,
Müller, H., Scharpegge, J., Chambers, G.K. &
Fleischer, R.C. (2008) A multilocus molecular
phylogeny of the parrots (Psittaciformes): support
for a Gondwanan origin during the Cretaceous.
Molecular Biology and Evolution, 25, 2141–2156.
Download