ArticlePDF Available

Habitat specialization predicts demographic response and vulnerability of floodplain birds in Amazonia

Authors:

Abstract

The annual flooding cycle of Amazonian rivers sustains the largest floodplains on Earth, which harbour a unique bird community. Recent studies suggest that habitat specialization drove different patterns of population structure and gene flow in floodplain birds. However, the lack of a direct estimate of habitat affinity prevents a proper test of its effects on population histories. In this work, we used occurrence data, satellite images and genomic data (ultra‐conserved elements) from 24 bird species specialized on a variety of seasonally flooded environments to classify habitat affinities and test its influence on evolutionary histories of Amazonian floodplain birds. We demonstrate that birds with higher specialization in river islands and dynamic environments have gone through more recent demographic expansion and currently have less genetic diversity than floodplain generalist birds. Our results indicate that there is an intrinsic relationship between habitat affinity and environmental dynamics, influencing patterns of population structure, demographic history and genetic diversity. Within the floodplains, historical landscape changes have had more severe impacts on island specialists, making them more vulnerable to current and future anthropogenic changes, as those imposed by hydroelectric dams in the Amazon Basin.
Molecular Ecology. 2023;00:1–17. wileyonlinelibrary.com/journal/mec
|
1© 2023 John Wiley & Sons Ltd.
1 | INTRODUCTIO N
Amazonia is the most biodiverse region on the planet and a nexus for
discussion on how physiographic and climatic processes drive diver-
sification (Cracraf t et al., 2020; Hoorn et al., 2010). Across its basin,
several of the largest rivers on Earth along with their broad flood-
plains currently delimit the distribution of multiple upland forest
vertebrates (Cracraft, 1985; Mourthé et al., 2022). To what extent
the evolution of this continental-scale river system and past climatic
changes affecting environment distributions were responsible for
Received: 18 April 2023 
|
Revised: 17 October 2023 
|
Accepted: 14 November 2023
DOI: 10.1111/mec.17221
ORIGINAL ARTICLE
Habitat specialization predicts demographic response and
vulnerability of floodplain birds in Amazonia
Eduardo D. Schultz1,2 | Gregory Thom3| Gabriela Zuquim4,5 |
Michael J. Hickerson6| Hanna Tuomisto4| Camila C. Ribas7
1Programa de Pós-Gra duação em Biologia
(Ecologia), Instituto Nacional de Pesquisas
da Amazôn ia, Manaus, AM, Brazil
2Department of Ornithology, American
Museum of Natural History, New York,
New York, USA
3Museum of Natural Science and
Department of Biological Sciences,
Louisiana State University, Baton Rouge,
Louisiana, USA
4Department of Biology, University of
Turku, Turku, Finland
5Department of Biology, Aarhus
University, Aarhus, Denmark
6Depar tment of Biology, City College of
New York, New York, New York, USA
7Coordenação de Biodiversidade, Instituto
Nacional de Pesquisas da Amazônia,
Manaus, AM, Brazil
Correspondence
Eduardo D. Schultz, Programa de
Pós-Graduação em Biologia (Ecologia),
Instituto Nacional de Pesquisas da
Amazônia, Manaus, AM, Brazil.
Email: edsbio@gmail.com
Funding information
Conselho Nacional de Desenvolvimento
Científ ico e Tecnológico, Grant/Award
Number: 311732/2020-8; Coordenação
de Aper feiçoamento de Pessoal de Nível
Superior; Fundação de Amparo à Pesquisa
do Estado do Amazonas, Grant/Award
Number: 002/2016 - POSGRAD 2017;
United States Agency for International
Development, Grant/Award Number:
PEER— Co Ag AID-OAA-A-11-00012
Handling Editor: Yanhua qu
Abstract
The annual flooding cycle of Amazonian rivers sustains the largest floodplains on
Earth, which harbour a unique bird community. Recent studies suggest that habitat
specialization drove different patterns of population structure and gene flow in flood-
plain birds. However, the lack of a direct estimate of habitat affinity prevents a proper
test of its effects on population histories. In this work, we used occurrence data, sat-
ellite images and genomic data (ultra-conserved elements) from 24 bird species spe-
cialized on a variety of seasonally flooded environments to classify habitat affinities
and test its influence on evolutionary histories of Amazonian floodplain birds. We
demonstrate that birds with higher specialization in river islands and dynamic envi-
ronments have gone through more recent demographic expansion and currently have
less genetic diversity than floodplain generalist birds. Our results indicate that there
is an intrinsic relationship between habitat affinity and environmental dynamics, in-
fluencing patterns of population structure, demographic history and genetic diversity.
Within the floodplains, historical landscape changes have had more severe impacts on
island specialists, making them more vulnerable to current and future anthropogenic
changes, as those imposed by hydroelectric dams in the Amazon Basin.
KEY WORDS
climate change, comparative demography, dam impacts, habitat affinities, quaternary, ultra-
conserved elements
2 
|
    SCHULTZ et al.
the forma tion of such diversit y is much debated (Cr acraft et al., 2020;
Haffer, 1969; Musher et al., 2022; Ribas et al., 2012, 2022). Recent
studies on co-distributed bird species from upland forests have re-
covered discordant evolutionary histories, which might suggest that
a single community-wide pattern of diversification does not exist
(Musher et al., 2022; Naka & Brumfield, 2018; Silva et al., 2019;
Smith et al., 2014). However, intrinsic ecological aspects of the spe-
cies, such as physiological constraints and habitat affinities, might
affect how populations respond to ex trinsic environmental changes,
and are expected to result in idiosyncratic evolutionary patterns
(Papadopoulou & Knowles, 2016). For example, the preference for
distinct forest strata in upland birds predicts genetic differenti-
ation between populations, with canopy species having shallower
differentiation than species occupying the understory (Burney &
Brumfield, 2009; Smith et al., 2014).
In the Amazonian floodplains, a specialized biota occupies a myr-
iad of distinct environments created by the annual flood cycle of the
rivers. These environments, that cover approximately 750,000 km2
(~11%) of the area of the Amazon basin, are deeply influenced by
the geomorphology of the rivers and the geological provinces they
drain (Wittmann et al., 2022). While clear and black water rivers
drain old cratons with low amounts of sediments and suspended
matter, leading to slow-growing floodplain vegetation, white water
rivers originate on Andean slopes and carry high amounts of nutri-
ents and sediments that continually reshape their floodplains, which
sustain a more stratified and dynamic vegetation (Junk et al., 2011).
Furthermore, floodplain habitats away from the main river channel
tend to be more stable than the ones closer to the river edges and
on river islands, which are more impacted by varying rates of sed-
iment deposition and erosion. Higher substrate stability results in
floodplain forests with tall trees adapted to shorter periods of in-
undation, while dynamic environments are mostly dominated by
shorter, herbaceous vegetation that stays under water for many
months ever y year (Figure 1; Junk et al., 2 011; Tamura et al., 2019;
Wittmann et al., 2022). This environmental complexity admits an
avifauna with rather distinct habitat affinities within the floodplains,
including a remarkable diversity of river islands specialists (Remsen
& Parker, 1983; Rosenberg, 1990).
The current configuration of the Amazonian floodplains is the
result of the combined effect of high sea level and the intensifica-
tion of the South American Monsoons that enhanced sedimentation
across the basin since the Last Glacial Maximum (LGM; 21 thousand
years ago), accelerated through the Holocene (last 11.7 thousand
years; Goldberg et al., 2021; Sawakuchi et al., 2022). However, a
previous phase of floodplain formation was interrupted by a period
of reducing sea level, starting around 80 thousand years ago, when
increased river incision drastically reduced the extent of the flood-
plains (Pupim et al., 2019). This scenario point s to a dynamic history
of expansion and retraction of the floodplains during the quaternary
climatic cycles, with a potentially large and heterogeneous impact on
the distribution and persistence of populations of floodplain species.
Moreover, these fluctuations have likely affected the distribution of
specific habitats in different ways, with a stronger impact on the
more ephemeral and dynamic environments like river islands (Passos
et al., 2020; Sawakuchi et al., 2022). Evidence from different stud-
ies seem to corroborate this hypothesis showing a more tumultu-
ous evolutionary history in bird taxa associated with more dynamic
floodplain environments (Barbosa et al., 2021; Choueri et al., 2017;
Johnson et al., 2023; Luna et al., 2022; Sawakuchi et al., 2022; Thom
et al., 2020, 2022). Nonetheless, a comprehensive classification of
habitat specialization across the basin is still needed for a thorough
understanding of the effect of distinct habitat associations in the
responses of the endemic floodplain bird community to past land-
scape changes.
Understanding the specificities of floodplain birds’ habitat affin-
ities and their response to environmental changes involves an addi-
tional pr actical urgency considering the high anthropogenic pressure
to which Amazonian rivers are exposed. Currently, hundreds of im-
plemented, under construction and planned dams for energy gen-
eration threaten to change the sedimentation dynamics and water
flow in the entire basin (Latrubesse et al., 2017). These enterprises
affect specific environments differently, and the lack of knowledge
about differences in habitat affinities within Amazonian floodplains
prevents accurate estimates of the potential impact s of these an-
thropogenic changes in the specialized biota (Cochrane et al., 2017;
Forsberg et al., 2017; Latrubesse et al., 2021; Melo et al., 2021).
Therefore in this study, we tested whether habitat specialization
in the floodplain environments predicts population history. We ex-
pected that generalist floodplain species occupying a wide variety
of floodplain environments, or species occurring on habitats more
continuously distributed across the floodplains, to have more sta-
ble historical demographics and less spatial genetic struc ture, when
compared to species specialized on the ephemeral river edge and in-
sular environments, which are highly impacted by sedimentation dy-
namics through time. To test this hypothesis, we gathered genomic
and occurrence data for 24 bird species specialized in Amazonian
seasonally flooded environments. We used satellite images to clas-
sify habitat preferences based on the occurrence on islands and on
the dynamism of the islands with occurrence records. Then, consid-
ering our classification, we used genomic data to test if habitat use
explains patterns of demographic history and genetic diversity. We
further test the influence of population struc ture on demographic
histories and whether demographic patterns are distinct for popu-
lations occurring in different portions of the Amazon Basin. Finally,
based on the observed responses to past landscape changes we dis-
cuss the vulnerabilit y of the studied taxa to anthropogenic impacts
on the floodplain environments.
2 |MATERIALS AND METHODS
2.1  | Sampling and genomic data acquisition
We selected 24 bird species, belonging to 6 different families, that
are known specialists on Amazonian seasonally flooded environ-
ments and had tissue samples available in the largest collections of
1365294x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/mec.17221 by Pennsylvania State University, Wiley Online Library on [29/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
   
|
 3
SCHULTZ et al.
Amazonian birds in Brazil (INPA and MPEG) and USA (AMNH, ANSP,
FMNH, LSUMNS, USNM). All species are endemic of the Amazon
Basin, in some cases also including populations in the floodplains
of adjacent basins and coastal areas (Table S1, Figures S1–S5).
Together, the set of selected species occupies multiple Amazonian
seasonally flooded environments, which makes it possible to com-
pare habitat preferences and the effect these might have on the
demographic histor y of the species. The genomic approach used
to access the demographic history of the studied birds was the se-
quence capture of ultra-conserved elements (UCEs). For each spe-
cies, tissues were selected aiming to cover as broadly as possible its
distribution (Table S2, Figures S1–S5). For each sample, DNA was
extracted using Qiagen DNeasy Blood and Tissue kit (Valencia, CA)
following the manufacturer's protocol. DNA extracts were sent to
Rapid Genomics (Gainesville, FL) where UCEs were sequenced using
a probe set targeting 2321 UCEs, commonly used in avian studies
(Barbosa et al., 2021; Luna et al., 2021, 2022; Sawakuchi et al., 2022;
Thom et al., 2020, 2022).
After sequencing, data were processed from raw reads to se-
quence alignments and single-nucleotide polymorphisms (SNPs) for
downstream analyses through the Phyluce pipeline (Faircloth, 2016).
First, raw reads were cleaned from adapter contamination using
FIGURE 1 Rivers and floodplain environments in the Amazon Basin. (a) Amazon Basin map with rivers coloured by the water colour
types. (b) Early successional Várzea vegetation in the left bank of the Juruá river near Juruá – Brazil (white water). (c) Early successional
Várzea vegetation in an island in the lower Solimões river near Manacapuru – Brazil (white water). (d) Mature Várzea forest in an island in the
lower Solimões river near Coari – Brazil (white water). (e, f) Mature Igapó Forests in islands of the Mariuá archipelago in the Negro river near
the confluence with the Branco river.
1365294x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/mec.17221 by Pennsylvania State University, Wiley Online Library on [29/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
4 
|
    SCHULTZ et al.
illumiprocessor function. Then, for each species a single sample with
high amounts of reads were assembled into contigs using SPAdes
(Bankevich et al., 2012), mapped against the probes using the func-
tion phyluce_assembly_match_contigs_to_probes and recovered con-
tigs were used as reference to all other samples. After wards, each
samples' reads were mapped against the reference alignment and
phased with phyluce_snp_bwa_multiple_align and phyluce_snp_phase_
uces respec tively (Andermann et al., 2018). Finally, sequences were
aligned into separate UCEs sequences files using mafft (Katoh &
Standley, 2013) and using phyluce_align_seqcap_align and trimmed to
improve alignment quality using phyluce_align_get_gblocks_trimmed_
alignments_from_untrimmed. We excluded samples with low cover-
age, usually less than 1500 UCEs, and loci that were not present
in all samples. This approach was used to have for each species a
dataset with more loci and shared by all samples. A SNP dataset was
obtained from the UCEs dataset by randomly selecting one SNP per
UCE, without including missing data, using snps_from_uce_align-
ments.py (h t t p s : / / g i t h u b . c o m / t o b i a s h o f m a n n 8 8 / U C E - d a t a - m a n a g
e m e n t / b l o b / m a s t e r / s n p s _ f r o m _ u c e _ a l i g n m e n t s . p y ).
2.2  | Population structure and definitions of
taxonomic units
To avoid population structure mistakenly affecting demographic
inferences, the SNPs datasets were used to estimate the best-
fit number of populations and define the taxonomic units using
sNMF function from LEA package in R (Frichot et al., 2014; R Core
Tea m, 2020). For each species, we estimated the cross-entropy of K
values from 1 to 10 populations, considering four different values
for the alpha parameter (1, 10, 100, 1000) with 100 replicates each,
and selected the K value with lowest cross-entropy across all runs.
This K value was used to define the working taxa for the main de-
mographic analyses and habitat categorization (Figures S1–S 5a ). To
visually access the spatial distribution of sNMF results, the ancestral
coefficient estimated for each sample was plotted in the map at each
sample's locality, for species with lowest cross-entropy value K= 1,
the second lowest value of K= 2 was plotted (Figures S1b–S5b).
As a complement to sNMF results, the estimated effective mi-
gration surfaces (EEMS) were used to look for areas of higher or
lower gene flow than expected under isolation by distance, as these
also reflect population structuring (Petkova et al., 2015). For each
species, Euclidean genetic distances (Rogers' distance) between
samples were estimated from the SNPs data with ADEGENET'S dist.
genpop function (Jombart & Ahmed, 2011). The described distribu-
tion of the species according to IUCN (https:// www. iucnr edlist. org/ )
was used as the distribution polygon, which was further edited when
necessary to include samples outside the polygon. Depending on the
size and complexity of the species' distribution polygon, values from
500 to 700 demes were assigned. The number of demes defines the
density of the population grid within the given distribution and, from
preliminary tests, we found these values to be suitable for the stud-
ied species. For each species, we present a single run of 10 × 106
generations, excluding the first 5 × 106 as burn-in (Figures S1–S 5c ).
The convergence of the runs was visually assessed in the posterior
probability trace plots.
For working taxa, as defined in the first sNMF analysis described
earlier that still included two geographically distinct clusters in sNMF
separated by areas of higher resistance to gene flow in agreement
with EEMS results (Figures S1–S5), we performed a second round of
demographic analyses treating these two clusters as distinct units.
The taxa resulting from this second approach are identified on the
maps in Figures S1–S5, and referred hereafter as populations, to
differentiate from the taxa defined in the first approach. We used
this approach to test if our demographic results were influenced by
population structure and if more geographically restricted lineages
would reveal distinct demographic patterns that could reflect local
history. Populations with few samples (fewer than nine individuals)
were excluded, as preliminary tests showed that the demographic
analyses with few samples recovered inconsistent results.
Working taxa and populations are identified in Figures S1–S5 by
its subspecies name when available, or by geographical identifiers
representing their distributions (W – West; E – East; Mad – Madeira;
Sol – Solimões; Ama – Amazonas; Neg – Negro).
2.3  | Habitat affinities
To test for differences in how species use the Amazonian flooding
environments, we gathered meta-data from the specimens in the
collections of INPA, LSUMNS and MPEG, and screened photo, vid-
eos and sound recordings available at the Macaulay Library (www.
macau layli brary. org) and Xeno-canto (www. xeno- canto. org) sub-
mitted until December 2020. Since available occurrence data for
Amazonian birds are not sufficiently accurate to infer fine-scale
habitat association, we relied on the pat tern of island specialization
to identif y differences in habitat affinities based on how each taxon
uses fluvial islands. Because islands have finite limits and are often
identified by the obser vers, we were able to calculate how often
each taxon was registered on islands or in the floodplains at the mar-
gins. Additionally, to test if different taxa prefer islands with distinct
dynamism, we used satellite images to assess how dynamic these
islands were based on changes in the amount of exposed water sur-
face between 1984 and 2018. This interval is not informative about
changes on an evolutionar y time scale, but depicts the characteris-
tics of the environments in which the taxa occur.
To calculate the proportions of registers on islands, ever y occur-
rence record was checked and, based on locality and observation
details, the occurrence was classified into one of the following two
categories: island or rivers margin. Records that could not be classi-
fied were excluded. To reduce bias from double counting the same
individual registered by different observers, distinct registers of the
same species on the same locality and date were counted only once
(Figures S1–S 5d, Table S3).
For the classification of island dynamism preference, the oc-
currences in islands were filtered, and only records on islands that
1365294x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/mec.17221 by Pennsylvania State University, Wiley Online Library on [29/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
   
|
 5
SCHULTZ et al.
could be identified were considered (Table S4). The identification of
islands was made by combining the geolocation, locality information
and obser vation details that indicated the specific island where the
register was made. For each identified island, we quantified three
variables to represent the dynamism of the island: amount of new
land, amount of lost land and degree of stability. The values for each
variable were derived from the water occurrence change intensity
dataset w hich measured the v ariation in water occ urrence on the sur-
face over multiple years using images from Landsat 5, 7 and 8 (Pekel
et al., 2016). This dataset measures both the direction of change
(increase, decrease or no change in water surface) and it s intensity.
Data were downloaded from https:// globa l- surfa ce- water. appsp ot.
com/ download and included information from 1984 to 2018. New
land was calculated based on the number of pixels with decrease in
water occurrence, weighted by the intensity of the change; lost land
was calculated by the number of pixels with increase in water occur-
rence, weighted by the intensity of the change; stability was calcu-
lated counting the pixels with invariant land through the observed
time. Constant water pixels were excluded. For each island, the val-
ues of the three variables were added and divided by the total to
obtain a proportional estimate. To extract the values, polygons were
drawn around each studied island in Qgis 3.14 (QGIS.org, 2020) and
exported as a shapefile. In R, the values within the polygon were
extracted from the raster with extract function from raster package
(Hijmans, 2020). The water colour classific ation of the river in which
each island was situated was obtained from a previous classification
of main Amazonian rivers (Venticinque et al., 2016) using the func-
tion ‘join attribute by location’ in Qgis.
The mean stability, gained and lost land for each taxon were cal-
culated considering all islands in which that taxon was registered. For
each of the three variables, the values extracted from each island were
added an d divided by the tot al of islands the ta xon were registere d. This
way, for each taxon we ended up with a proportional estimate of sta-
bility, gained and lost land in the islands they were registered that add
up to 1 (Tables 1 and S5). Following long-described evidences of bird
specialization in different successional stages of seasonally flooded en-
vironments (Remsen & Parker, 1983) and on islands (Rosenberg, 1990),
we tested if mean stability – representing the amount of dynamism –
and proportion of registers on islands could identify ecological groups,
which could be further used to test if birds with similar habitat affini-
ties shared similar demographic histories.
2.4  | Demographic inferences and genetic diversity
Demographic inferences were estimated in DILS (Fraïsse et al., 2021).
Using an approximate Bayesian computation framework, DILS allows
the estimation of demographic models considering the potential ef-
fects of linked selection. The analysis calculates the posterior proba-
bilities of models of demographic expansion, contraction or stability,
and estimates parameters like the time of demographic expansion
and met rics of genetic divers ity. DILS runs were perfor med using sin-
gle population analysis option, without outgroups. First, the 29 taxa
identified in the population structure analysis were used as evolu-
ti ona ry u nit s. To keep mo st lo ci an d con sider in g the data wer e al rea dy
filtered in Phyluce, max_N_tolerated was set to 0.3, Lmin to 100 and
nMin to the number of sequences per file. A general mutation rate
of 2.5 × 10−9 substitutions/site/generation with a 1-year generation
time was used to calibrate the analyses. Despite the uncertainty in-
volving substitution rates, the rate used here matches the rate pro-
posed for the genome of birds by Nater et al. (2015), the mean rate
of 2.3 × 10−9 across multiple bird lineages obtained by Nadachowska-
Brzyska et al. (2015) and previous studies of Amazonian floodplain
birds, making results comparable (Barbosa et al., 2021; Sawakuchi
et al., 2022; Thom et al., 2018, 2020). The values of the priors for
minimum and maximum time of demographic change were set to 100
and 5 × 105 generations, respectively, and limits of population sizes
were set at the default values of 100 for the minimum and 106 for
the maximum, encompassing with a margin previously published es-
timates of these parameters for Amazonian floodplain birds (Barbosa
et al., 2021; Luna et al., 2021, 2022; Sawakuchi et al., 2022; Thom
et al., 2020). To check for the consistency of the analyses, for each
taxon three runs were performed with the same priors and the run
with the highest sum of goodness-of-fit p values was chosen. The
parameter values used for downstream analyses were obtained fol-
lowing the optimized posterior of the neural network method.
To further explore if population structure influenced demo-
graphic results, additional demographic inferences were performed
using the populations within taxa detected in our secondary popula-
tion structure analysis. This procedure was per formed only for taxa
with cross-entropy values of K= 2 close to K= 1 and with the two
putative populations having distinct geographic distribution accord-
ing to sNMF and EEMS results. The same approach with the same
set of priors was used in DILS as when considering the taxon as a
single unit.
For taxa with cross-entropy values of K= 2 close to K= 1 and with
the two putative populations having distinct geographic distribution
according to sNMF and EEMS results, additional demographic infer-
ences were performed for these populations separately (Table S1).
This approach was used firstly to explore if population structure in-
fluenced demographic results and was compared with the result s
considering them a single taxon (Figure S6a). Moreover, as distinct
portions of the Amazon Basin could have gone through different
geological and environmental changes, there could be geographi-
cally distinct trends in population history along the vast Amazonian
floodplains. Therefore, we also compared the demographic signal of
populations with relatively overlapping distributions (Figure S6b).
Populations with widespread or idiosyncratic distributions were not
included in this geographic comparison. Demographic analyses for
these populations were performed in DILS with the same approach
and set of priors described earlier.
To access the genetic diversity of studied taxa, we used the av-
erage Tajima's pi (π) and Waterson's theta (θ) calculated in DILS for
each taxon. Values of π and θ describe the proportion of polymor-
phic sites and are indicatives of genetic diversity, with lower values
suggesting lower genetic diversity. The diversity measures were
1365294x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/mec.17221 by Pennsylvania State University, Wiley Online Library on [29/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
6 
|
    SCHULTZ et al.
compared with the estimates of habitat affinities to test if there is a
relationship between genetic diversity and distinct habitat affinities.
2.5  | Hypotheses, statistical analyses and data
presentation
Our aim was to test if differences in habitat affinity influenced evo-
lutionary histories of co-distributed taxa. For that, we used linear
regressions and analyses of variance (ANOVAs) to test if our metric
of habitat use – represented by the propor tion of registers on islands
and the mean stabilit y of those islands – and our ecological groups
(see Results) explain the variation in demographic history (time of
demographic change) and population genetic parameters (π and θ). A
null hypothesis of no statistical correlation between habitat use and
the demographic and genetic variables means that the differences
between species are random or require an alternative explanation.
Alternatively, statistical correlation between variables indicates an
influence of habitat affinity in shaping demographic histories and/or
genetic diversity. Linear regressions between continuous variables
were used to test if more specialized birds had more pronounced
demographic responses, independent of categories. In turn, we used
the ANOVAs to test if our ecological categorization predicts similar
demographic histories for birds that occupy similar environments.
TABLE 1 Sampling, island use, genetic diversity and demographic variables for each taxon.
Tax on
Genomic
samples UCEs SNPs Habitat Islands Margins
Island
proportion
Mean
stability
Mean lost
land
Mean gained
land Average πAverage θ
Expansion
posterior
probability
HPD 2.5%
expansion time
Median
expansion time
HPD 97.5%
expansion time
Attila bolivianus 22 2242 1686 Generalists 27 106 20.3 0.6131 0.1153 0. 2717 0.0009 0.00164 0.9997 84,673.09 95, 417.19 1 07,1 26. 67
Capito aurovirens 13 2203 1177 Generalists 12 90.6 9. 4 0.6958 0 .1381 0.1660 0.0007 0.0009 0.9820 16 ,764.87 19,382. 65 22,412 .84
Conirostrum bicolor minus 14 2174 1596 Dynamic Specialists 63 12 84 0.4941 0.1074 0.3985 0.00 09 0.0013 0.9996 45,294.48 51, 65 0. 27 58,573.94
Cranioleuca gutturata 20 2256 1650 Generalists 16 80 21 .1 0.8490 0.0181 0.13 30 0.0013 0.00309 1.0000 150,903.81 161,455.87 173,331.51
Cranioleuca vulpecula 18 2192 1054 Dynamic Specialists 111 397. 37 0.2305 0.1498 0.6197 0.0004 0.00 069 0.9901 15 ,769.68 18,612. 62 21 ,93 9.0 3
Dendroplex kienerii 19 2237 1 519 Generalists 56 54 50.91 0.76 49 0.1209 0.1142 0.0008 0.00113 0.9980 44,841.92 55,043.63 66341.72
Furnarius minor 10 2194 1108 Dynamic Specialists 75 26 74.26 0.2791 0.1905 0.5305 0.0006 0.00079 0.9885 2 7,9 06. 28 32,213.99 36,746 .61
Knipolegus orenocensis sclateri 21 2239 1498 Dynamic Specialists 55 493.22 0. 2589 0.2678 0.4733 0.0009 0.00135 0.9878 24,088.68 28,454.01 33,368.78
Knipolegus poecilocercus 31 2145 1670 Generalists 32 51 38.55 0.8813 0.0331 0.0857 0.001 0.00195 0.9996 62,813.48 69,104 .33 76 ,437.4 0
Mazaria propinqua 22 2267 1450 Dynamic Specialists 82 20 80.39 0.2893 0.1668 0.5439 0.0006 0.00098 0.9990 38,286.81 43,892.87 50,207.36
Myiopagis flavivertex 21 2242 2130 Generalists 20 86 18.87 0.8192 0.0686 0.1122 0.0023 0.00432 1.0000 202,537.55 218,799.58 235,126.93
Myrmoborus lugubris lugubris 12 2256 2011 Generalists 313 18.75 0.810 0 0.1297 0.0603 0.0007 0.00088 0.9073 28 929.81 3 9,0 28 .4 8 49, 070.44
Myrmoborus lugubris berlepschi 28 2256 2011 Stable Specialists 82 297. 62 0 .610 0 0.1056 0.2844 0.00 05 0.00072 0.9823 12,642.98 15,395.36 18,661 .85
Myrmoborus lugubris femininus 13 2256 2011 Stable Specialists 9 2 81.82 0.7405 0.1435 0.1161 0.0008 0.00102 0.9958 63,114.72 80,610.59 99,674. 28
Myrmochanes hemileucus 22 2219 1063 Dynamic Specialists 129 894.16 0.3942 0.1627 0.4431 0.0004 0.00079 0.9823 18,791.46 21,009.60 23,442 .55
Myrmotherula assimilis Ama 16 2253 1985 Stable Specialists 51 11 82. 26 0.7576 0.0527 0.1898 0.001 0.00129 0.9996 72,867.78 88,796.22 107,444.59
Myrmotherula assimilis Mad 82253 1985 Generalists 214 12.5 0.8792 0.0867 0.0341 0.0011 0.0 0155 1.0000 118,448.27 138,385.30 160,544.54
Myrmotherula assimilis Sol 10 2253 1985 Generalists 18 21 4 6.15 0.6450 0.14 06 0. 2144 0.0011 0.00139 1.0000 90,453.46 111 ,468.95 136,10 9.3 9
Myrmotherula klagesi 21 2191 1314 Stable Specialists 48 10 82. 76 0.8338 0.0533 0.1129 0.0011 0.0022 1.0000 80,723.79 89, 13 0.1 1 97,986. 44
Myrmotherula multostriata 18 2237 1705 Generalists 21 105 16 .67 0.8934 0.0494 0.0572 0.0015 0.00277 1.0000 146,650.63 1 59,873 .6 0 173,603.83
Nasica longirostris 38 2227 2058 Generalists 33 207 13.75 0.7249 0.114 3 0.160 9 0.0012 0.00253 0.9992 123,446.58 135,090.56 147, 45 2. 35
Ochthornis littoralis 24 2081 14 47 Generalists 21 91 18.75 0.7318 0.1482 0.120 0 0.0007 0.00132 1.0000 56,203.12 67,703.40 80,140.40
Serpophaga hypoleuca hypoleuca 18 2233 1378 Dynamic Specialists 38 15 71.7 0.2742 0.2153 0.5105 0.0009 0.00165 1.0000 60,199.81 66 ,12 9.6 0 72,674 .39
Stigmatura napensis napensis 29 2231 1256 Dynamic Specialists 59 13 81.94 0 .3861 0.1848 0.4291 0.0007 0.00115 0.9651 22,812 .19 25,759.18 29, 099.85
Synallaxis albigularis 25 2233 1660 Generalists 60 69 46.51 0.5009 0.1242 0.3750 0.0009 0.00194 1.0000 92,530.14 9 9,3 51. 95 107,080.50
Thamnophilus cryptoleucus 23 2244 2117 Dynamic Specialists 114 39 7.4 4 0.4825 0.1512 0.3663 0.0007 0.00124 1.0000 70,125.48 81 ,141.58 92,995.31
Thamnophilus nigrocinereus cinereoniger 82249 2117 Stable Specialists 67 26 72.04 0.8382 0.0205 0.1412 0.0009 0.00095 0.9292 11 , 30 7.36 14,716.4 6 18,313.63
Thamnophilus nigrocinereus nigrocinereus 10 2249 2117 Stable Specialists 43 786 0.9248 0.0518 0.0234 0.0011 0.0014 0.9998 69,593 .2 2 91 , 697.2 0 116,710.01
Xiphorhynchus obsoletus 60 2191 1416 Generalists 80 2 61 23.46 0. 8317 0.0634 0.10 49 0.0006 0.00228 0.9990 69,9 86 .17 74,455.70 79,443.79
Note: For each taxon are depicted number of genomic samples; recovered UCEs; recovered SNPs; habitat affinity category; number of registers on
islands; number of regis ters on margins; propor tion of registers on islands; mean stability; mean lost land; mean gained land on islands it was
registered at; average π; average theta; average Tajima's D; posterior probability of population expansion model estimated in DILS; time of
demographic expansion in years, including lower and upper values of 95% confidence interval and the median.
1365294x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/mec.17221 by Pennsylvania State University, Wiley Online Library on [29/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
   
|
 7
SCHULTZ et al.
All statistical analyses on the correlation between variables
were performed in R and plotted with ggplot2 (Wickham, 2016).
The correlation coefficients between continuous variables pre-
sented in sc atter plots were estimated using stat_cor function using
Pearson method from the ggpubr package (Kassambara, 2020).
Confidence intervals shown in grey shades around the mean were
plotted with geom_smooth function from ggplot2. The ANOVA s be-
tween variables and habitat affinities categories were per formed
with stat_compare_means function from ggpubr package using
anova method.
The ternary plot with the mean values of the stability, lost land
and gained land in the islands each taxon occur (Figure 2b) was made
with ggtern (Hamilton & Ferry, 2018). The mean stability calculated
for each taxon was used in the ordination of the taxa in Figure 2c.
The ternary plots showing the proportion of the three variables
in each island the taxa were registered (Figure 3) were made with
ggtern. Islands were classified and coloured according to the rivers
where they occur. Habitat affinity categories were defined accord-
ing to the islands in which taxa were registered.
To test for phylogenetic independence of the results, we calcu-
lated the phylogenetic independent contrasts (PIC) of each variable
following the method described by Felsenstein (1985) using the fun c-
tion pic from ape package in R (Paradis & Schliep, 2019; Figure S7).
The phylogeny used as input for PIC calculation was built based on
TABLE 1 Sampling, island use, genetic diversity and demographic variables for each taxon.
Tax on
Genomic
samples UCEs SNPs Habitat Islands Margins
Island
proportion
Mean
stability
Mean lost
land
Mean gained
land Average πAverage θ
Expansion
posterior
probability
HPD 2.5%
expansion time
Median
expansion time
HPD 97.5%
expansion time
Attila bolivianus 22 2242 1686 Generalists 27 106 20.3 0.6131 0.1153 0. 2717 0.0009 0.00164 0.9997 84,673.09 95, 417.19 1 07,1 26. 67
Capito aurovirens 13 2203 1177 Generalists 12 90.6 9. 4 0.6958 0 .1381 0.1660 0.0007 0.0009 0.9820 16 ,764.87 19,382.65 22,412 .84
Conirostrum bicolor minus 14 2174 1596 Dynamic Specialists 63 12 84 0.4941 0.1074 0.3985 0.00 09 0.0013 0.9996 45,294.48 51, 65 0. 27 58,573.94
Cranioleuca gutturata 20 2256 1650 Generalists 16 80 21 .1 0.8490 0.0181 0.13 30 0.0013 0.00309 1.0000 150,903.81 161,455.87 173,331.51
Cranioleuca vulpecula 18 2192 1054 Dynamic Specialists 111 397. 37 0.2305 0.1498 0.6197 0.0004 0.00 069 0.9901 15 ,769.68 18,612. 62 21 ,93 9.0 3
Dendroplex kienerii 19 2237 1 519 Generalists 56 54 50.91 0.76 49 0.1209 0.1142 0.0008 0.00113 0.9980 44,841.92 55,043.63 66 341.72
Furnarius minor 10 2194 1108 Dynamic Specialists 75 26 74.26 0.2791 0.1905 0.5305 0.0006 0.00 079 0.9885 2 7,9 06 . 28 32,213.99 36,746.61
Knipolegus orenocensis sclateri 21 2239 1498 Dynamic Specialists 55 493.22 0. 2589 0.2678 0.4733 0.0009 0.00135 0.9878 24,088.68 28,454.01 33,36 8.78
Knipolegus poecilocercus 31 2145 1670 Generalists 32 51 38.55 0.8813 0.0331 0.0857 0.001 0.00195 0.9996 62,813.48 69,104 .33 76 ,437.4 0
Mazaria propinqua 22 2267 1450 Dynamic Specialists 82 20 80.39 0.2893 0.1668 0.5439 0.0006 0.00098 0.9990 38,286.81 43,892.87 50,207.36
Myiopagis flavivertex 21 2242 2130 Generalists 20 86 18.87 0.8192 0.0686 0.1122 0.0023 0.00432 1.0000 202,537.55 218,799.58 235,126.93
Myrmoborus lugubris lugubris 12 2256 2011 Generalists 313 18.75 0.810 0 0.1297 0.0603 0.0007 0.00088 0.9073 28 929.81 3 9,0 28 .4 8 49, 070.44
Myrmoborus lugubris berlepschi 28 2256 2011 Stable Specialists 82 297. 62 0 .610 0 0.1056 0.2844 0.00 05 0.00072 0.9823 12,642.98 15,395.36 18,661 .85
Myrmoborus lugubris femininus 13 2256 2011 Stable Specialists 9 2 81.82 0.7405 0.1435 0.1161 0.0008 0.00102 0.9958 63,114.72 80,610.59 99,674. 28
Myrmochanes hemileucus 22 2219 1063 Dynamic Specialists 129 894.16 0.3942 0.1627 0.4431 0.0004 0.00079 0.9823 18,791.46 21,009.60 23,442 .55
Myrmotherula assimilis Ama 16 2253 1985 Stable Specialists 51 11 82. 26 0.7576 0.0527 0.1898 0.001 0.00129 0.9996 72,867.78 88,796.22 107,444.59
Myrmotherula assimilis Mad 82253 1985 Generalists 214 12.5 0.8792 0.0867 0.0341 0.0011 0.0 0155 1.0000 118,448.27 138,385.30 160,544.54
Myrmotherula assimilis Sol 10 2253 1985 Generalists 18 21 4 6.15 0.6450 0.14 06 0. 2144 0.0011 0.00139 1.0000 90,453.46 111 ,468.95 136,10 9.3 9
Myrmotherula klagesi 21 2191 1314 Stable Specialists 48 10 82. 76 0.8338 0.0533 0.1129 0.0011 0.0022 1.0000 80,723.79 89, 13 0.1 1 97,986. 44
Myrmotherula multostriata 18 2237 17 05 Generalists 21 105 16.67 0.8934 0.0494 0.0572 0.0015 0.00277 1.0000 146, 650.63 1 59,873 .6 0 173,603.83
Nasica longirostris 38 2227 2058 Generalists 33 207 13.75 0.7249 0.114 3 0.160 9 0.0012 0.00253 0.9992 123,446.58 135,090.56 147, 45 2. 35
Ochthornis littoralis 24 2081 14 47 Generalists 21 91 18.75 0.7318 0.1482 0.120 0 0.0007 0.00132 1.0000 56,203.12 67,703.40 80,140.40
Serpophaga hypoleuca hypoleuca 18 2233 1378 Dynamic Specialists 38 15 71.7 0.2742 0.2153 0.5105 0.0009 0.00165 1.0000 60,199.81 66 ,12 9.6 0 72,674 .39
Stigmatura napensis napensis 29 2231 1256 Dynamic Specialists 59 13 81.94 0 .3861 0.1848 0.4291 0.0007 0.00115 0.9651 22,812 .19 25,759.18 29, 099.85
Synallaxis albigularis 25 2233 1660 Generalists 60 69 46.51 0.5009 0.1242 0.3750 0.0009 0.00194 1.0000 92,530.14 9 9,3 51. 95 107,080.50
Thamnophilus cryptoleucus 23 2244 2117 Dynamic Specialists 114 39 7.4 4 0.4825 0.1512 0.3663 0.0007 0.00124 1.0000 70,125.48 81 ,141.58 92,995.31
Thamnophilus nigrocinereus cinereoniger 82249 2117 Stable Specialists 67 26 72.04 0.8382 0.0205 0.1412 0.0009 0.00095 0.9292 11 , 30 7.36 14,716.4 6 18,313.63
Thamnophilus nigrocinereus nigrocinereus 10 2249 2117 Stable Specialists 43 786 0.9248 0.0518 0.0234 0.0011 0.0014 0.9998 69,593 .2 2 91 , 697.2 0 116,710.01
Xiphorhynchus obsoletus 60 2191 1416 Generalists 80 2 61 23.46 0. 8317 0.0634 0.10 49 0.0006 0.00228 0.9990 69,9 86 .17 74,455.70 79,443.79
Note: For each taxon are depicted number of genomic samples; recovered UCEs; recovered SNPs; habitat affinity category; number of registers on
islands; number of regis ters on margins; propor tion of registers on islands; mean stability; mean lost land; mean gained land on islands it was
registered at; average π; average theta; average Tajima's D; posterior probability of population expansion model estimated in DILS; time of
demographic expansion in years, including lower and upper values of 95% confidence interval and the median.
1365294x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/mec.17221 by Pennsylvania State University, Wiley Online Library on [29/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
8 
|
    SCHULTZ et al.
the cytochrome b (cytb) sequences of one individual per taxon. To
ensure the provenience of the samples for the tree, and to include
all taxa in the tree, we recovered the cytb sequences from the reads
from our UCE sequencing which usually contains reads of mtDNA
(Raposo do Amaral et al., 2015). For each taxon, one cytb sequence
downloaded from GenBank was used as reference to map the clean
reads from one of our samples in Geneious (Kearse et al., 2012). In
the absence of a cytb sequence deposited in GenBank for a taxon,
its closest relative was used as reference. After the alignment of the
sequences in Geneious, the best-fitting model of molecular evo-
lution was estimated based on the Bayesian information criterion
(BIC) with jModeltest (Darriba et al., 2012). The phylogenetic rela-
tionship of the taxa was recovered with MrBayes 3.2.3 (Ronquist &
Huelsenbeck, 2003), setting Capito aurovirens, the only non-Passeri-
formes of our dataset, as outgroup.
3 |RESULTS
3.1  | Genetic structure and definition of
evolutionary units
Our genomic approach recovered an average of 2224 UCE loci per
species and randomly selecting one SNP per UCE, we obtained
an average of 1661 SNPs per species (Table 1). For 19 of the 24
studied species, sNMF recovered the lowest cross-entropy value
for K= 1, suggesting a better treatment of them as a single taxon
(Figures S1–S5). For Knipolegus orenocensis (K= 2), Myrmoborus
lugubris (K= 4) and Thamnophilus nigrocinereus (K= 4), population
structure corresponded to described subspecies, and thus subspe-
cies designations were used. Two monotypic species had varying
lower cross-entropy values across different runs, Myrmotherula
FIGURE 2 Island use by Amazonian floodplain birds. In all graphics, colours depict the same ecological groups. (a) Relationship between
proportion of registers in islands (vs. river margins) and the mean stability of islands. Straight line and p and R2 values depict the correlation
between the two variables including all taxa. (b) Ternary plot of the mean values of stability, new land and lost land for the islands in which
each taxon was registered. (c) Islands and taxa sorted by the proportion of stable land and mean stability on islands taxa were registered
respectively. For each taxon, bars represent presence on the island.
1365294x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/mec.17221 by Pennsylvania State University, Wiley Online Library on [29/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
   
|
 9
SCHULTZ et al.
assimilis (K= 2 and 3) and Xiphorhynchus obsoletus (K= 1 and 2).
Myrmotherula assimilis was treated as 3 taxa (Mad, Sol and Ama, for
the populations in the Madeira, Solimões and Amazonas river ba-
sins) and X. obsoletus, conservatively, as a single population follow-
ing previous results (Luna et al., 2021; Thom et al., 2020). For a few
of these intra-specific taxa, the small number of samples hampered
accurate demographic estimates (see Demographic analyses). Thus,
K. orenocensis xinguensis, M. lugubris stictopterus, T. nigrocinereus hu-
beri and T. nigrocinereus tschudii were excluded from subsequent
analysis. The remaining 29 taxa were used as taxonomic units for
estimating island and habitat use and for the main demographic in-
ferences (Tables 1 and S1).
EEMS results show that several of the taxa defined earlier share
an area of higher resistance to gene flow in the central portion of the
basin or in different sub-basins like in the Negro and Madeira rivers
(Figures S1–S5). For many of those taxa, the second best sNMF ar-
rangement shows two corresponding geographically distinct popu-
lations (Figures S1–S5).
3.2  | Island use reveals distinct habitat affinities
Altogether, throughout the Amazon basin, we identified 107 river
islands where at least one of the 29 studied taxa had been photo-
graphed, recorded or collected (Table S5). Moreover, our data reveal
a gradient in the characteristics of these islands, ranging from stable
islands maintaining a steady floodplain forest in most of their area
during this period, to dynamic islands in which most land was created
or lost to water, resulting in the dominance of early successional en-
vironments (Figure 2). We also found high variation in the proportion
FIGURE 3 Islands characteristics. Ternary diagrams with the three variables measured: Estimated stability (S), gained land (G) and lost
land (L) of the islands in which taxa were registered. For each taxon, dots depict the islands it was recorded and colours represent the colour
of the river where the island is. On the right, all studied islands are shown at the top ternary diagram; a map of the Amazon basin with river
colours and the studied islands on the centre; and at the bottom the count of islands in each river colour by taxa in each ecological group.
1365294x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/mec.17221 by Pennsylvania State University, Wiley Online Library on [29/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
10 
|
    SCHULTZ et al.
of occurrences on islands (vs. river banks) among taxa and in the
mean stability of the islands where they occur (Figure 2, Table 1).
However, these two variables combined allowed the identification
of three habitat affinity categories (Figure 2a). A group of taxa (here-
after called ‘Generalists’) was commonly found both on riverbanks
and on islands, with typically less than 50% of their occurrences on
islands. Furthermore, the islands occupied by Generalists were rela-
tively st able, with 50% or more of their surface area having remained
land during the entire 40 year period. In contrast, two groups of ‘is-
land specialists’ had at least 70% of their occurrences on islands, but
differed in the average stability of the islands they occur. One of
the groups was associated with islands that had, on average, 60%
or more stable land (hereafter called ‘Stable Specialists’), while the
other group of taxa occurred on islands with less than 50% of stable
land, on average (hereafter called ‘Dynamic Specialists’). The ternary
diagram based on the mean values of the three estimated variables
– proportion of stable land, proportion of lost land and proportion
of gained land – for the islands where each taxon occurs shows that
dynamic specialists prefer islands with lower proportions of stable
land (Figure 2b). The ordination of the taxa based on the mean sta-
bility of the islands on which they occur also shows that Dynamic
Specialists occur in less stable islands compared to Generalists and
Stable Specialists (Figure 2c). Although Dynamic Specialists were
registered on islands with very distinct characteristics regarding the
three estimated variables, they were always on white water rivers
(Figure 3).
3.3  | Habitat affinities predict demographic
histories and genetic diversity
For all 29 taxa, three independent DILS runs recovered the same
model of demographic expansion with high posterior probability
(>0.9). All independent runs recovered very similar parameter val-
ues (e.g. time of demographic change and population size) for each
taxon. The estimated ages for the expansions among taxa spanned
from around 220 to 20 kya. The mean estimated ages for the de-
mographic expansion was negatively correlated with the proportion
of occurrences on islands ( p= .0019, R2= 0.3) and positively corre-
lated with the mean stability of the islands where the taxa occur
(p= .0041, R2= 0.27). An ANOVA between the mean estimated ex-
pansion ages and the three classes of habitat affinities additionally
endorses the relationship between habitat use and demographic his-
tory (p= .0082; Figure 4).
Genetic diversity estimates were correlated with the metrics of
habitat affinity (Figure 5). Two distinct measures of genetic diver-
sity, the average Tajima's pi (π) and Watterson's theta (θ), tend to be
lower in taxa with a higher proportion of registers in islands (p= .014,
R2= 0.2 and p= .0 03, R2= 0.28 respectively) and in taxa with lower
mean stability in the islands they occur (p= .0052, R2= 0.25 and
p= .016, R2= 0.2 respectively).
Transforming the variables into PIC resulted in even stronger
correlations (Figure S7), indicating that the reported correlations
were not driven by specific clades and reinforces the communi-
ty-wide relationship between birds' habitat affinities and responses
to environmental changes across the Amazonian floodplains.
FIGURE 4 Relationship between demographic histories and
habitat use. On y-axis, each dot depicts mean estimated times
of demographic expansion for a taxon, with black lines depicting
the confidence interval and colours the habitat af finity category
as in previous figures. (a) Relationship between demographic
histories and propor tion of registers in islands. (b) Relationship
between demographic histories and mean stability of islands in
which taxa were registered. (c) Relationship between mean times
of demographic change for each habitat af finity group. Mean
regression line, R2 and p values in (a) and (b), and ANOVA's p value
in (c) depict statistical correlation between variables. Identical
graphics identifying the taxa are presented in Figure S8.
1365294x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/mec.17221 by Pennsylvania State University, Wiley Online Library on [29/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
   
|
11
SCHULTZ et al.
3.4  | No effect of population structure and
geographic distribution in the demographic histories
Demographic analyses for separate populations within taxa were
strongly correlated with the inferences considering the 29 taxa
(p= 3e-08), suggesting no effect of population structure on the re-
sults (Figure S6a). No significant relationship was found between
similarity in distributional ranges and demographic patterns (p= .67).
Each region harbours t axa with demographic expansion happening
at different times, and, in general, island specialists had more recent
demographic expansions than Generalists (Figure S6b). This result
suggests that the demographic expansions and dates of expansions
were not geographically restricted but rather diffuse across the
Amazon Basin and are better explained by habitat affinities of the
birds than by their geographic distribution.
4 |DISCUSSION
Results presented here suggest that birds with higher specialization
in river islands have gone through more recent demographic expan-
sion and currently have less genetic diversity than floodplain generalist
birds. Am ong island specialists, birds t hat occupy less sta ble island hab-
itats showed more recent expansion dates and lower genetic diversity
than birds in more stable islands. These relationships show an intrinsic
link between habitat and biotic history mediated by habitat affinity,
revealing that the histor y of Amazonian landscapes may have different
effects on species with distinct ecological characteristics.
4.1  | Island use as a measure of habitat affinity
Our results reveal two distinct, yet complementary, facets of habitat
specialization in the Amazonian floodplains: island occupancy and
preference for dynamic versus stable environments. Combined, these
two features point to three distinct categories of habitat affinities.
Throughout the Amazon basin, Generalists are commonly found
in seasonally flooded environments along the river margins, and
when using islands, these birds tend to prefer more stable ones
(Figures 2 and 3). Mostly, these taxa are widely distributed across
the basin, occurring on rivers with dif ferent characteristics and or-
igins (Figures S1–S5). The stability estimated by our analysis refers
to environments in which no exposed water was registered since
the beginning of satellite image capture in the 1980s, even during
FIGURE 5 Relationship between genetic diversity parameters and habitat use. On y-axis, each dot depicts average values of the
estimated genetic diversity parameter for a taxon and colours the habitat affinity category as in previous figures. Each parameter is plotted
against proportion of registers in islands and mean stabilit y on registered islands. Plots at the bottom show averages for habitat affinity
groups. Mean regression line, R2 and p values depict statistical correlation between variables. Identical graphics identifying the taxa are
presented in Figure S9.
1365294x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/mec.17221 by Pennsylvania State University, Wiley Online Library on [29/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
12 
|
    SCHULTZ et al.
the high water seasons (Pekel et al., 2016). In the floodplains these
terrains represent seasonally flooded forests and seem to have ex-
isted for even longer than registered. In fact, mature floodplain for-
ests harbour trees with ages of hundreds to a thousand years and
some of the older islands date to several thousand years (Passos
et al., 2020; Sawakuchi et al., 2022; Wittmann et al., 2022). Except
for Synallaxis albigularis and Ochthornis littoralis that occupy open
environments, the Generalists in our classification are mainly flood-
plain forests dwellers, which might explain their use of these stable
islands (Remsen & Parker, 1983; Ridgely & Tudor, 2009). Moreover,
large stable islands are often formed by anabranches that trans-
form floodplains margins in islands and they tend to have environ-
ments more similar to the ones found on the margins (Wittmann
et al., 2022).
The Stable Specialists occur on generally stable islands that
harbour vegetation virtually indistinguishable in structure from the
islands on which the Generalists typically occur (Figures 2 and 3).
However, they are rarely found along river margins. Interestingly,
unlike nearly all Generalists and Dynamic Specialists, most Stable
Specialists comprise species complexes with an evident geographic,
genetic and phenotypic variation that reflects a complex history of
isolation and secondary contact in Central Amazonia over the qua-
ternary (Thom et al., 2020). We note that while a few taxa in these
complexes are classified as Generalists – as M. l. lugubris, Myrmot.
assimilis Mad and Myrmot. assimilis Sol – this may be a sampling arte-
fact, as we have a reduced number of records for these intra-specific
lineages ( Tables 1, S4 and S5).
Finally, Dynamic Specialists are taxa strongly associated with
islands and occupy a wide range of island types, although on aver-
age occurring on more dynamic islands and always on white water
rivers (Figures 1 and 2). In these rivers, even large stable islands usu-
ally have por tions where the deposition of new sediments results
in the colonization of new vegetation and the occurrence of these
early-growth specialist birds. In the islands, these birds show even
further specialization and preference for different early-growth veg-
etation environments continuously created by the sedimentation
dynamic (Rosenberg, 1990). Paradoxically, the high dynamism of is-
lands in sediment-rich white water rivers might result in the constant
availability of new environments, allowing connectivity among pop-
ulations across a large extent of the Amazon basin when historical
climatic conditions favour sediment accumulation.
4.2  | Drivers of habitat specialization in the
Amazonian floodplains
The old divergences between floodplain birds and their extant rela-
tives in other environments suggest a long history of specialization
in floodplain environments. While Amazonian upland bird line-
ages usually date to the Pleistocene (i.e. less than 2.6 Ma; Cracraft
et al., 2020; Silva et al., 2019), most floodplain lineages studied here
have stem ages that date from the Pliocene to the late Miocene
(i.e. from 2.6 to 11 Ma; Table S1; Harvey et al., 2020). It is possible
that the draining of the extensive wetlands in Western Amazonia
starting in the late Miocene, created the early stages of a basin-
wide floodplain system and allowed the development of a special-
ized avifauna (Wittmann et al., 2022). Notwithstanding, except for
a few representatives in the Orinoco basin, adjacent to the Amazon,
river island specialization is an evolutionary strategy found nowhere
else on Earth and requires additional explanations. It has been pos-
tulated that island specialists are great dispersers, allowing them
to rapidly colonize new islands and move between islands distant
from one another (Rosenberg, 1990 ). Increased competition at the
margins, either with closely related t axa or with ecologically similar
species, may help explaining these unique evolutionar y pathways
(Robinson & Terborgh, 1995). In the Stable Specialists group, all
taxa belong to the Thamnophilidae, one of the most abundant and
species-rich families in the understory of Amazonian environments.
All three Thamnophilid genera of Stable Specialists (Myrmoborus,
Thamnophilus and Myrmotherula) have multiple distinct species occu-
pying both t he floodplains al ong the margins and n on-flooded uplan d
forests (Cohn-Haft, Naka, & Fernandes, 2007; Cohn-haft, Pacheco,
et al., 20 07; Remsen & Parker, 1983; Ridgely & Tudor, 2009). In con-
trast, except for Thamnophilus cryptoleucus and Cranioleuca vulpec-
ula, that belong to genera well represented in several Amazonian
environments, Dynamic Specialists are either in monotypic genera,
like Myrmochanes and Mazaria, or belong to genera mostly associ-
ated with open environments across different Neotropical regions,
like Conirostrum, Furnarius, Knipolegus, Serpophaga and Stigmatura
(Ridgely & Tudor, 2009). Throughout the basin, habitats similar to
those on dynamic islands are also found along the margins of white
water rivers and are commonly occupied by several bird species that
are widespread across open environments in the Neotropics (Cohn-
Haft, Naka, & Fernandes, 2007; Cohn-haft, Pacheco, et al., 2007;
Remsen & Parker, 198 3). Islands, on the other hand, are marked by
lower species richness with high dominance and elevated population
densities of specialist species (Borges et al., 2019; Rosenberg, 199 0).
Therefore, by specializing in islands, both groups of island special-
ists avoid inter-specific competition and find environments of dif-
ficult access to most non-specialist species. Nevertheless, despite
the processes that drove and maintained island specialization, as
we discuss ahead, it resulted in similar demographic and genetic di-
versity patterns for Dynamic and Stable Specialists, which reflects
the recent history of formation of islands along Amazonian rivers
(Sawakuchi et al., 2022).
4.3  | Differential habitat affinities resulted in
distinct demographic histories
Within the avifauna associated with Amazonian seasonally flooded
habitat s, no evident shared demographic pattern could be recog-
nized without considering the habitat preferences of the different
taxa. Following our categorization of habitat use, we found habitat
specialization and demographic history to be correlated. Our data
show a general trend for island specialization leading to more recent
1365294x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/mec.17221 by Pennsylvania State University, Wiley Online Library on [29/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
   
|
13
SCHULTZ et al.
demographic expansion. Also, Dynamic Specialists, with an arguably
more restricted niche, had the most recent and dynamic population
history (Figure 4). Thus, our results demonstrate that detailed occur-
rence data and ecological preferences categorization, provide not
onl y a stronger hypothesis testing framework, but also sharper re so-
lution of biogeographic patterns (Papadopoulou & Knowles, 2016).
Although dynamic islands, mostly associated with white water
rivers, are more common in western Amazonia (Figure 3), our data
show a stronger effect of habitat affinity on the demographic his-
tory of the studied taxa when compared to geographic distribution
(Figure S6). Throughout the basin, co-distributed lineages with dis-
tinct habitat affinities show discordant ages of demographic expan-
sion, but island specialists generally show more recent population
expansion, regardless of their geographic distribution. Moreover,
this relationship between habitat affinity and demographic histor y
stands independent of the phylogenetic relationships of the stud-
ied taxa (Figure S7). These results agree with evidence that, rather
than localized, historical fluctuations in sedimentation dynamics
affected the distribution of suitable habitat concomitantly and ho-
mogeneously across the whole basin (Sawakuchi et al., 2022; Thom
et al., 2022).
The distinct demographic histories associated with distinc t pat-
terns of habitat use agree with available data on the formation of
Amazonian floodplains. Geological data suggest a recent origin of
the current configuration of floodplain environments. The oldest
fluvial terraces at the Solimões river date to the last 250 kya, and
the modern extensive floodplain system, including an intricate net-
work of islands, dates to the late Quaternary, when the combina-
tion of sea level rise and increased precipitation allowed sediment
accumulation throughout lowland Amazonia, especially after the
LGM (Goldberg et al., 2021; Passos et al., 2020;Pupim et al., 2019;
Sawakuchi et al., 2022). Therefore, the formation of older fluvial
terraces around 250 kya increased the availability of floodplain
environments, allowing population expansion of Generalists with
less restricted habitat affinities (Pupim et al., 2019; Sawakuchi
et al., 2022). Nowadays, several of these taxa are widely distributed
across the floodplains in rivers with heterogeneous characteristics,
both in islands and margins (Figures S1–S5) and can be found in dif-
ferent river-created habitats, including transitions to non-flooded
environments (Remsen & Parker, 198 3). Thus, the amplitude of hab-
itat use in these taxa may have allowed widespread and connected
populations even in moments of floodplains retractions due to river
incision.
On the other hand, island specialists are tightly connected to
the unique environment in which they occur, leading to stronger
demographic fluctuations in response to the availability of islands,
which can also erase older demographic signals. Therefore, the
more recent ages of demographic expansion of these birds likely
reflect the last phase of floodplain construction after the LGM
(Goldberg et al., 2021; Sawakuchi et al., 2022). In this scenario,
periodic reduction of sediment accumulation along the Amazon
basin would result in severe population reduction for Dynamic
Specialists, leading to high rates of extinction of local populations
(Thom et al., 2020). The long branches connecting these taxa to
their closest relatives and their current pattern of low structure
and genetic diversity corroborate historical instability and past ex-
tinction of local populations.
A distinct pattern is observed in Stable Specialists. In periods of
reduced sedimentation that affected the availability of islands, pop-
ulations might be fragmented in different portions of the basin, but
due to the less ephemeral environments in which they occur, they
might have sufficient available habitat to persist on more stable is-
lands. In this case, the fragmentation of populations combined with
lower extinction rates may have resulted in the observed pattern of
shared population structure and reduced gene flow in the centre of
the basin, common to several of these taxa (Figures S2, S4 and S5;
Choueri et al., 2017; Thom et al., 2020).
Therefore, our results support the hypothesis that the avifauna
associated with seasonally flooded environments in Amazonia has
a history marked by constant changes in habitat availability and
connectivity, tightly linked to climatic changes during the quater-
nary (Barbosa et al., 2021; Johnson et al., 2023; Luna et al., 2022;
Sawakuchi et al., 2022; Thom et al., 2020). However, the way spe-
cies occupy distinct floodplain environments deeply affected their
responses to landscape changes. Lower and higher environmental
specialization, as observed in Generalists and Dynamic Specialists,
respectively, resulted in a similar pattern of widespread distributions
and reduced population structure. Nonetheless, their distinct de-
mographic histories indicate that while in the Generalists, the lack
of structure is associated with having sustained higher connectiv-
ity across the basin, in the dynamic specialists it is associated with
higher local extinc tion rates and recent demographic expansion.
Stable Specialists, in contrast, show an intermediate response to
landscape changes, resulting in a spatially congruent pattern of pop-
ulation structure and demographic history (Thom et al., 2020).
4.4  | Habitat specialization and vulnerability to
environmental changes
Our findings have direct implications for conservation strategies
given that habitat dynamism is a key predictor of population per-
sistence across the landscape. Among the ever-growing threats
Amazonia faces, hundreds of dams for hydroelectric energy genera-
tion are already built, under construction or planned in the different
Amazonian countries, with no integrated planning or strategy for im-
pact evaluation and mitigation. These dams threaten the rivers, local
human populations and the biota associated with seasonally flooded
environments, both at local and regional scales, but only local im-
pacts are usually accounted for (Latrubesse et al., 2017). Throughout
the basin, different t ypes of dams impose different environmental
impact s, but all of them dramatically alter the dynamics of seasonally
flooded environments, either by permanently flooding large extents
of distinc t environments along the river upstream of the dam, or by
reducing the flux of water and sediments downstream (Latrubesse
et al., 2021). Nonetheless, the seasonality of the inundation regime
1365294x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/mec.17221 by Pennsylvania State University, Wiley Online Library on [29/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
14 
|
    SCHULTZ et al.
is crucial for the physiology, growth and phenology of trees adapted
to floodplain environments, and disrupting it results in the death
of floodplain trees, with consequences for the whole ecosystem
(Latrubesse et al., 2021; Parolin et al., 2010; Schöngart et al., 2021;
Wittmann et al., 2022).
Our demographic analyses suggest that island specialists are
more vulnerable to changes in sediment ation dynamics than taxa
that also occupy the river margins. Due to the specificit y of the
environments they occupy, changes in sediment supply and water
discharge that affect island formation may cause population frag-
mentation and local extinctio ns. Also, island spe cialists have compar-
atively lower genetic diversity, and thus increased risk of extinction
due to range fragmentation (Figure 4). The dynamic formation of
islands across multiple rivers during the recent history of Amazonia
has promoted continuous gene flow and supported the persistence
of these taxa with little population structure and low genetic diver-
sity, across great extents of the Amazon Basin (Johnson et al., 2023).
However, most of the d istribution of seve ral island specialists, par tic-
ularly Dynamic Specialists, coincides with the Amazonian sub-basins
most threatened by dams, like the Ucayali, Marañon and Madeira
(Figures S1–S5; L atrubesse et al., 2017, 2021).
Currently, IUCN considers all Stable Specialists taxa studied
here to be ‘Near Threatened or ‘Vulnerable’ due to population
decreases caused by the advances of agriculture, logging, dams
and/or mining. Several Generalist taxa, though currently consid-
ered to be ‘least concern’, are also suggested to have decreasing
populations. However, except for Conirostrum bicolor, considered
‘near threatened’, Dynamic Specialist birds studied here are mostly
considered ‘least concern’ with stable populations (IUCN, 2021).
Genomic analyses presented here capture the history of popula-
tions at a timescale of thousands to a million years ago and cannot
detect very recent demographic changes (Nadachowska-Brzyska
et al., 2022). Nonetheless, we show that in addition to currently
having a more restricted habitat, Dynamic Specialists responded
more dramatically to historical environmental changes, probably
going through strong bottlenecks that resulted in populations
with lower genetic diversity. Combined, these results suggest
that Dynamic Specialists taxa are at least as vulnerable to envi-
ronmental changes as Stable Specialist taxa and more so than
Generalist taxa. Although difficult to quantify, it is already clear
that dams reduce the availability of Dynamic Specialist habitat
(Cochrane et al., 2017 ), ultimately reducing their population sizes,
which respectively fit criteria B and A of IUCN (IUCN Standards
and Petitions Committee, 2019). The ef fective impact of damming
Amazonian rivers on the biota associated with seasonally flooded
environments will greatly depend on the future policies regarding
Amazonian river management and conservation across different
countries, and the specialized avifauna is an important indicator in
guiding decision making.
AUTHOR CONTRIBUTIONS
Conceptualization: EDS, CCR. Methodology: all authors.
Investigation: EDS. Visualization: EDS, CCR, HT, GT. Supervision:
CCR, HT, MJH. Writing – original draft: EDS. Writing – review and
editing: all authors.
ACKNO WLE DGE MENTS
We thank the curators and curatorial assistants at the American
Museum of Natural History (AMNH), New York, USA; Academy of
Natural Sciences of Drexel University, Philadelphia, USA (ANSP);
Field Museum of Natural History, Chicago, USA (FMNH); Instituto
Nacional de Pesquisas da Amazônia, Manaus, Brazil (INPA); Louisiana
State University Museum of Natural Science (LSUMZ), Baton Rouge,
USA; and Museu Paraense Emílio Goeldi, Belém, Brazil (MPEG) for
tissue loans and collection information that helped in the habitat
use classification. We thank the ornithologists and birdwatchers
who uploaded their records and field observations of floodplain
birds without which this work would never be possible. We also
thank Cassiano Gatto, James V. Remsen, Mario Cohn-Haft and Gary
Rosenberg for helping identify specific localities of past expeditions.
Finally, we thank Romina Batista, Fábio Raposo do Amaral, André O.
Sawakuchi, João M. G. Capurucho, Mario Cohn-Haft, Joel Cracraft
and two anonymous reviewers for insightful comments on drafts of
this article.
Financial support to this study was provided by the US Agency
for International Development (USAID) through a Partnerships
for Enhanced Engagement in Research grant (PEER Co Ag AID-
OAA-A-11-00012) and by FAPEAM (Fundação de Amparo à Pesquisa
do Estado do Amazonas) through the ‘Chamada Internacional
Biodiversa 2019–2020’. EDS received a doctorate fellowship from
FAPEAM (002/2016 POSGRAD 2017), PDSE fellowship from
CAPES (Edital no 41/2018 – Seleção 2019), Erasmus+ scholarship
that made a research visit to University of Turku possible and a
Chapman Postdoctoral Fellowship from the American Museum of
Natural History. CCR is supported by CNPq (311732/2020-8). Data
analyses were run on the supercomputers of CSC – IT Center for
Science in Finland.
CONFLICT OF INTEREST STATEMENT
The authors declare no competing interests.
DATA AVAIL ABILI TY STATEMENT
All data needed to evaluate the conclusions in the paper are avail-
able in the main text or the supplementary materials. Genetic data
will be available at the Sequence Read Archive (SRA) and supple-
mentar y material is available at https:// datad ryad. org/ stash/ share/
w L j 2 V 3 0 s 9 i A L A z g U U l x N _ t h Z t K i d m G - j 4 3 d x i 1 O N i T A or may be re-
quested from the authors.
ORCID
Eduardo D. Schultz https://orcid.org/0000-0003-2647-0644
Gregory Thom https://orcid.org/0000-0001-6200-0565
Gabriela Zuquim https://orcid.org/0000-0003-0932-2308
Michael J. Hickerson https://orcid.org/0000-0002-5802-406X
Hanna Tuomisto https://orcid.org/0000-0003-1640-490X
Camila C. Ribas https://orcid.org/0000-0002-9088-4828
1365294x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/mec.17221 by Pennsylvania State University, Wiley Online Library on [29/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
   
|
15
SCHULTZ et al.
REFERENCES
Andermann, T., Fernandes, A. M., Olsson, U., Töpel, M., Pfeil, B., Oxelman,
B., Aleixo, A., Faircloth, B. C., & Antonelli, A. (2018). Allele phasing
greatly improves the phylogenetic utility of ultraconser ved ele-
ments. Systematic Biology, 68(1), 32–46. https:// doi. org/ 10 . 1093/
sysbio/ syy039
Bankevich, A., Nurk, S., Antipov, D., Gurevich, A. A., Dvorkin, M.,
Kulikov, A . S., Lesin, V. M., Nikolenko, S. I., Pham, S., Prjibelski, A.
D., Pyshkin, A. V., Sirotkin, A. V., Vyahhi, N., Tesler, G., Alekseyev,
M. A., & Pevzner, P. A. (2012). SPAdes: A new genome assembly
algorithm and its applications to single-cell sequencing. Journal of
Computational Biology, 19, 455–477. https:// doi. org/ 10. 1089/ cmb.
2012. 0021
Barbosa, W. E. d. S., Ferreira, M., de Deus Schultz, E., Willians Luna, L.,
Orsi Laranjeiras, T., Aleixo, A., & Cherem Ribas, C. (2021). Habitat
association constrains population histor y in two sympatric oven-
birds along Amazonian floodplains. Journal of Biogeography, 49,
1683–1695. h t t p s : / / d o i . o r g / 1 0 . 1 1 1 1 / j b i . 1 4 2 6 6
Borges, S. H., Baccaro, F., Moreira, M., & Choueri, E. L. (2019). Bird as-
semblages on Amazonian river islands: Patterns of species diversity
and composition. Biotropica, 51(6), 903–912. https:// doi. org/ 10.
1111/ btp. 12716
Burney, C. W., & Brumfield, R. T. (20 09). Ecology predicts levels of ge-
netic differentiation in neotropical birds. The American Naturalist,
174(3), 358–368. ht tp s:// doi. or g/ 10. 10 86/ 603613
Choueri, É. L., Gubili, C., Borges, S. H., Thom, G., Sawakuchi, A. O.,
Soares, E. A . A., & Ribas, C. C. (2017). Phylogeography and pop-
ulation dynamics of Antbirds (Thamnophilidae) from Amazonian
fluvial islands. Journal of Biogeography, 44(10), 2284–2294. htt ps ://
d o i . o r g / 1 0 . 1 1 1 1 / j b i . 1 3 0 4 2
Cochrane, S. M. V., Matricardi, E. A. T., Numata, I., & Lefebvre, P. A.
(2017). Landsat-based analysis of mega dam flooding impacts in
the Amazon compared to associated environmental impact assess-
ments: Upper Madeira River example 2006–2015. Remote Sensing
Applications: Society and Environment, 7, 1–8. https:// doi. org/ 10.
1 0 1 6 / j . r s a s e . 2 0 1 7 . 0 4 . 0 0 5
Cohn-Haft, M., Naka, L. N., & Fernandes, A. M. (2007). Padrões de
distribuição da avifauna da várzea dos rios Solimões-Amazonas.
Bases Científicas Para a Conservação Da Várzea: Identificação e
Caracterização de Regiões Biogeográficas. 287–324.
Cohn-haft, M., Pacheco, A. M. F., Bechtoldt, C. L., Torres, M. D. F.,
Fernandes, A. M., Sardelli, C. H., & Macêdo, I. T. (2007). Inventário
ornitológico. Biodiversidade Do Médio Madeira: Bases Científicas Para
Propostas de Conservação (pp. 145–178). INPA.
Cracraft, J., Ribas, C. C ., D'H orta, F. M., Bate s, J., Almei da, R. P., Aleixo,
A., Boubli, J. P., Campbell, K. E., Cruz, F. W., Ferreira, M., Fritz, S.
C., Grohmann, C. H., Latrubesse, E. M., Lohmann, L. G., Musher,
L. J., Nogueira, A., Sawakuchi, A . O., & Baker, P. (2020). The origin
and evolution of Amazonian species diversity. In V. Rull & A . C .
Carnaval (Eds.), Neotropical diversification: Patterns and processes
(pp. 225–244). Springer. https:// doi. org/ 10. 1007/ 978- 3- 030-
3 1 1 6 7 - 4 _ 1 0
Cracraft, J. (1985). Historical biogeography and pat terns of differenti-
ation within the south American avifauna: Areas of endemism.
Ornithological Monographs, 336 (36), 49–84. https:// doi. org/ 10.
2307/ 40168278
Darriba, D., Taboada, G. L., Doallo, R., & Posada, D. (2012). JModelTest
2: More models, new heuristics and parallel computing. Nature
Methods, 9, 772. https:// doi. org/ 10. 1038/ nmeth. 2109
Faircloth, B. C. (2016). PHYLUCE is a software package for the analysis
of conser ved genomic loci. Bioinformatics, 32(5), 786–788. ht tp s: //
d o i . o r g / 1 0 . 1 0 9 3 / b i o i n f o r m a t i c s / b t v 6 4 6
Felsenstein, J. (1985). Phylogenies and the comparative method. The
American Naturalist, 125, 1–15. h t t p s : / / d o i . o r g / 1 0 . 1 0 8 6 / 2 8 4 3 2 5
Forsberg, B. R., Melack, J. M., Dunne, T., Barthem, R. B., Goulding, M.,
Paiva, R. C. D., Sorribas, M. V., Silva, U. L., & Weisser, S. (2017). The
potential impact of new A ndean dams on Amazon fluvial ecosys-
tems. PLoS One, 12(8), e0182254. https:// doi. org/ 10. 1371/ journ al.
pone. 0182254
Fraïsse, C., Popovic, I., Mazoyer, C., Spataro, B., Delmotte, S., Romiguier,
J., Loire, É., Simon, A., Galtier, N., Duret, L., Bierne, N., Vekemans,
X., & Roux, C. (2021). DI LS: Demographic infere nce s wit h lin ked se-
lection by using ABC. Molecular Ecology Resources, 21, 2629–2644.
https:// d oi. org/ 10. 1111/ 1755 - 09 98. 13 323
Frichot, E., Mathieu, F., Trouillon, T., Bouchard, G., & François, O. (2014).
Fast and efficient estimation of individual ancestry coefficients.
Genetics, 196(4), 973–983. h t t p s : / / d o i . o r g / 1 0 . 1 5 3 4 / g e n e t i c s . 1 1 3 .
160572
Goldberg, S. L., Schmidt, M. J., & Perron, J. T. (2021). Fast response of
Amazon Rivers to quaternary climate cycles. Journal of Geophysical
Research: Earth Surface, 126(11), e2021JF006416. https:// doi. org/
1 0 . 1 0 2 9 / 2 0 2 1 J F 0 0 6 4 1 6
Haffer, J. (1969). Speciation in Amazonian Forest birds. Science, 165(3889),
131–137. https:// doi. org/ 10. 1126/ scien ce. 165. 3889. 131
Hamilton, N. E., & Ferry, M. (2018). Ggtern: Ternary diagrams using gg-
plot2. Journal of Statistical Software, 87, 1–7. https:// doi. org/ 10.
18637/ jss. v087. c03
Harvey, M. G., Bravo, G. A., Claramunt, S., Cuervo, A. M., Derryberry,
G. E., Battilana, J., Seeholzer, G. F., McKay, J. S., O'Meara, B. C.,
Faircloth, B. C., Edwards, S. V., Pérez-Emán, J., Moyle, R. G.,
Sheldon, F. H., Aleixo, A ., Smith, B. T., Chesser, R. T., Silveira, L. F.,
Cracraft, J., … Der ryberry, E. P. (2020). The evolution of a tropical
biodiversity hotspot. Science, 370(6522), 1343–1348. ht t ps : // doi .
org/ 10. 1126/ scien ce. aaz6970
Hijmans, R. J. (2020). raster: Geographic analysis and modeling with ras-
ter data. R Package Version 3.1-5.
Hoorn, C., Wesselingh, F. P., ter Steege, H., Bermudez, M. A., Mora, A .,
Sevink, J., Sanmartin, I., Sanchez-Meseguer, A., Anderson, C. L.,
Figueiredo, J. P., Jaramillo, C., Riff, D., Negri, F. R., Hooghiemstra,
H., Lundberg, J., Stadler, T., Sarkinen, T., Antonelli, A., Sanmartín,
I., … Antonelli, A. (2010). Amazonia through time: Andean uplift,
climate change, landscape evolution, and biodiversity. Science,
330(6006), 927–931. h t t p s : / / d o i . o r g / 1 0 . 1 1 2 6 / s c i e n c e . 1 1 9 4 5 8 5
IUCN. (2021). The IUCN Red List of Threatened Species. Version 2020-3.
In https:// www. iucnr edlist. org. Downloaded on 12 Januar y 2021
IUCN Standards and Petitions Committee. (2019). IUCN standards and
petitions committee. Standards and Petitions Committee.
Johnson, O., Ribas, C . C., Aleixo, A., Naka, L. N., Harvey, M. G., &
Brumfield, R. T. (2023). Amazonian birds in more dynamic habi-
tats have less population genetic structure and higher gene flow.
Molecular Ecology, 32(9), 2186–2205. http s:// doi. org / 10. 1111/ m ec.
16886
Jombart, T., & Ahmed, I. (2011). adegenet 1.3-1: New tools for the anal-
ysis of genome-wide SNP data. Bioinformatics, 27, 3070–3071.
https:// doi. org/ 10. 1093/ bioin forma tics/ btr521
Junk, W. J., Piedade, M. T. F., Schöngart, J., Cohn-Haft, M., Adeney, J. M.,
& Wittmann, F. (2011). A classification of major naturally-occurring
Amazonian lowland wetlands. Wetlands, 31(4), 623–640. ht tp s: //
doi. org/ 10. 1007/ s1315 7- 011- 0190- 7
Kassambara, A. (2020). ggpubr: “ggplot2” Based Publication Ready Plots.
R Package Version 0.4.0.
Katoh, K., & Standley, D. M. (2013). MAFFT multiple sequence alignment
software version 7: Improvements in performance and usability.
Molecular Biology and Evolution, 30, 772–780. https:// doi. org/ 10.
1093/ molb ev/ mst010
Kearse, M., Moir, R., Wilson, A., Stones-Havas, S., Cheung, M., Sturrock,
S., Buxton, S., Cooper, A., Markowitz, S., Duran, C., Thierer, T.,
Ashton, B., Meintjes, P., & Drummond, A. (2012). Geneious basic:
An integrated and extendable desktop software platform for the
1365294x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/mec.17221 by Pennsylvania State University, Wiley Online Library on [29/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
16 
|
    SCHULTZ et al.
organization and analysis of sequence data. Bioinformatics, 28,
1647–1649. https:// doi. org/ 10. 1093/ bioin forma tics/ bts199
Latrubesse, E. M., Arima, E. Y., Dunne, T., Park, E., Baker, V. R., d'Horta,
F. M., Wight, C., Wittmann, F., Zuanon, J., Baker, P. A., Ribas, C. C.,
Norgaard, R. B., Filizola, N., Ansar, A., Flyvbjerg, B., & Stevaux, J. C.
(2017). Damm ing the rivers of th e Amazon basin. Nature, 546(76 58) ,
363–369. https:// doi. org/ 10. 1038/ natur e22333
Latrubesse, E. M., d'Ho rta, F. M., Ribas, C. C., Wittmann, F., Zuanon, J.,
Pa rk, E., Dunne, T., Arima, E. Y., & Baker, P. A. (20 21). Vu lne ra bilit y
of the biota in riverine and seasonally flooded habitats to dam-
ming of Amazonian rivers. Aquatic Conservation: Marine and
Freshwater Ecosystems, 31(5), 1136–1149. https:// doi. org/ 10.
1 0 0 2 / a q c . 3 4 2 4
Luna, L. W., Naka, L. N., Thom, G., Knowles, L. L., Sawakuchi, A. O., Aleixo,
A., & Ribas, C. C. (2022). Late Pleistocene landscape changes and
habitat specialization as promoters of population genomic diver-
gence in Amazonian floodplain birds. Molecular Ecology, 32(1), 214–
228. h t t p s : / / d o i . o r g / 1 0 . 1 1 1 1 / m e c . 1 6 7 4 1
Luna, L. W., Ribas, C. C., & Aleixo, A. (2021). Genomic differentiation
with gene flow in a widespread Amazonian floodplain-specialist
bird species. Journal of Biogeography, 49, 1670–1682. htt p s: // d oi .
o r g / 1 0 . 1 1 1 1 / j b i . 1 4 2 5 7
Melo, T. N. d., Cerqueira, M. C., D'Horta, F. M., Tuomisto, H., Van
Doninck, J., & Ribas, C. C. (2021). Impacts of a large hydroelec-
tric dam on the Madeira River (Brazil) on floodplain avifauna.
Acta Amazonica, 51(4), 298–310. https:// doi. org/ 10. 1590/ 1809-
43922 02100662
Mourthé, Í., Hilário, R. R., Carvalho, W. D., & Boubli, J. P. (2022). Filtering
effec t of large River s on primate distribution in the Brazilian
Amazonia. Frontiers in Ecology and Evolution, 10(May), 1–11. ht tp s: //
doi. org/ 10. 3389/ fevo. 2022. 857920
Musher, L . J., Giakoumis, M., Albert, J., Del-Rio, G., Rego, M., Thom, G.,
Aleixo, A., Ribas, C. C., Brumfield, R. T., Smith, B. T., & Cracraft, J.
(2022). River network rearrangements promote speciation in low-
land Amazonian birds. Science Advances, 8(14), eabn1099. h tt ps ://
doi. org/ 10. 1126/ sciadv. abn1099
Nadachowska-Brzyska, K., Konczal, M., & Babik, W. (2022). Navigating
the temporal continuum of effective population size. Methods in
Ecology and Evolution, 13(1), 22–41. h t t p s : / / d o i . o r g / 1 0 . 1 1 1 1 / 2 0 4 1 -
2 1 0 X . 1 3 7 4 0
Nadachowska-Brzyska, K., Li, C., Smeds, L., Zhang, G., & Ellegren, H.
(2015). Temporal dynamics of avian populations during pleistocene
revealed by whole-genome sequences. Current Biology, 25(10 ),
1375–1380. h t t p s : / / d o i . o r g / 1 0 . 1 0 1 6 / j . c u b . 2 0 1 5 . 0 3 . 0 4 7
Naka, L. N., & Brumfield, R. T. (2018). The dual role of Amazonian riv-
ers in the generation and maintenance of avian diversity. Science
Advances, 4(August), 8575–8576. https:// doi. org/ 10. 1126/ sciadv.
aar8575
Nater, A., Burri, R., Kawakami, T., Smeds, L., & Ellegren, H. (2015).
Resolving evolutionary relationships in closely related species with
whole-genome sequencing data. Systematic Biology, 64(6), 1000–
10 17. h t t p s : / / d o i . o r g / 1 0 . 1 0 9 3 / s y s b i o / s y v 0 4 5
Papadopoulou, A., & Knowles, L. L. (2016). Toward a paradigm shift in
comparative phylogeography driven by trait-based hypotheses.
Proceedings of the National Academy of Sciences, 113(29), 8018–
8024. http s:// doi. or g/ 10. 1073/ p nas. 16 010 69113
Paradis, E., & Schliep, K. (2019). Ape 5.0: An environment for modern
phylogenetics and evolutionary analyses in R. Bioinformatics, 35,
526–528. https:// doi. org/ 10. 1093/ bioin forma tics/ bty633
Parolin, P., Lucas, C., Piedade, M. T. F., & Wittmann, F. (2010). Drought
responses of flood-tolerant trees in Amazonian floodplains. Annals
of Botany, 105(1), 129–139. https:// doi. org/ 10. 1093/ aob/ mcp258
Passos, M. S., Soares, E. A. A., Tatumi, S. H., Yee, M., Mittani, J. C. R.,
Hayakawa, E. H., & Salazar, C. A. (2020). Pleistocene-Holocene
sedimentary deposits of the Solimões-Amazonas fluvial system,
Western Amazonia. Journal of South American Earth Sciences, 98,
102455. h t t p s : / / d o i . o r g / 1 0 . 1 0 1 6 / j . j s a m e s . 2 0 1 9 . 1 0 2 4 5 5
Pekel, J. F., Cottam, A., Gorelick, N., & Belward, A. S. (2016). High-
resolution mapping of global surface water and its long-term
changes. Nature, 540(7633), 418–422. ht tps:// doi. o rg/ 10. 10 38/
natur e20584
Petkova, D., Novembre, J., & Stephens, M. (2015). Visualizing spatial
population structure with estimated effective migration surfaces.
Nature Genetics, 48(1), 94–100. h t t p s : / / d o i . o r g / 1 0 . 1 0 3 8 / n g . 3 4 6 4
Pupim, F. N., Sawakuchi, A. O., Almeida, R. P., Ribas, C. C., Kern, A.
K., Hartmann, G. A., Chiessi, C. M., Tamura, L. N., Mineli, T. D.,
Savian, J. F., Grohmann, C. H., Bertassoli, D. J., Stern, A . G., Cruz,
F. W., & Cracraft, J. (2019). Chronology of Terra Firme formation
in Amazonian lowlands reveals a dynamic quaternary landscape.
Quaternary Science Reviews, 210, 154–163. https:// doi. org/ 10.
1 0 1 6 / j . q u a s c i r e v . 2 0 1 9 . 0 3 . 0 0 8
QGIS.org. (2020). QGIS geographic information system. QGIS Association.
h t t p : / / w w w . q g i s . o r g
R Core Team. (2020). R: A Language and Environment for Statistical
Computing. R Core Team.
Raposo do Amaral, F., Neves, L. G., Resende, M. F. R., Mobili, F., Miyaki,
C. Y., Pellegrino, K. C. M., & Biondo, C. (2015). Ultraconserved el-
ements sequencing as a low-cost source of complete mitochon-
drial genomes and microsatellite markers in non-model amniotes.
PLoS One, 10(9), e0138446. https:// doi. org/ 10. 1371/ journ al. pone.
01384 46
Remsen, J. V. J., & Parker, T. A. I. (1983). Contribution of river-created
habitats to bird species richness in Amazonia. Biotropica, 15(3),
223–231. https:// doi. org/ 10. 2307/ 2387833
Ribas, C. C., Aleixo, A., Nogueira, A. C. R., Miyaki, C. Y., & Cracraft, J.
(2012). A palaeobiogeogr aphic model for biotic diversification
within Amazonia over the past three million years. Proceedings of
the Royal Society B: Biological Sciences, 279(1729), 681–689. ht tp s: //
doi . org/ 10. 10 98/ rs pb. 2011. 11 20
Ribas, C. C., Fritz, S. C., & Baker, P. A. (2022). The challenges and po-
tential of geogenomics for biogeography and conser vation in
Amazonia. Journal of Biogeography, 49(10), 1839–1847. h ttp s:// d o i.
o r g / 1 0 . 1 1 1 1 / j b i . 1 4 4 5 2
Ridgely, R. S., & Tudor, G. (2009). Field guide to the songbirds of South
America: The passerines. University of Texas Press.
Robinson, S. K., & Terborgh, J. (1995). Aggression and habitat selection
by Amazonian birds. Journal of Animal Ecology, 64(11), 1–11. htt ps ://
d o i . o r g / 1 0 . 1 0 0 7 / s 1 3 3 9 8 - 0 1 4 - 0 1 7 3 - 7 . 2
Ronquist, F., & Huelsenbeck, J. P. (2003). MrBayes 3: Bayesian phyloge-
netic inference under mixed models. Bioinformatics, 19, 1572–1574.
https:// doi. org/ 10. 1093/ bioin forma tics/ btg180
Rosenberg, G. H. (1990). Habitat specialization and foraging behavior by
birds of Amazonian River Islands in northeastern Peru. The Condor,
92(2), 427. h t t p s : / / d o i . o r g / 1 0 . 2 3 0 7 / 1 3 6 8 2 4 0
Sawakuchi, A. O., Schultz, E. D., Pupim, F. N., Bertassoli, D. J., Souza,
D. F., Cunha, D. F., Mazoca, C. E., Ferreira, M. P., Grohmann, C.
H., Wahnfried, I. D., Chiessi, C. M., Cruz, F. W., Almeida, R. P., &
Ribas, C . C. (2022). Rainfall and sea level drove the expansion of
seasonally flooded habitats and associated bird populations across
Amazonia. Nature Communications, 13(1), 4945. https:// doi. org/ 10.
1 0 3 8 / s 4 1 4 6 7 - 0 2 2 - 3 2 5 6 1 - 0
Schöngart, J., Wittmann, F., Faria de Resende, A., Assahira, C., Sousa
Lobo, G., Rocha Duarte Neves, J., Rocha, M., Biem Mori, G., Costa
Quaresma, A., Oreste Demarchi, L., Weiss Albuquerque, B., Oliveira
Feitosa, Y., Silva Costa, G., Vieira Feitoza, G., Machado Durgante,
F., Lopes, A., Trumbore, S. E., Sanna Freire Silva, T., Steege, H., …
Piedade, M. T. F. (2021). The shadow of the Balbina dam: A synthe-
sis of over 35 years of downstream impacts on floodplain forests
in Central Amazonia. Aquatic Conservation: Marine and Freshwater
Ecosystems, 31(5), 1117–1135. h t t p s : / / d o i . o r g / 1 0 . 1 0 0 2 / a q c . 3 5 2 6
1365294x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/mec.17221 by Pennsylvania State University, Wiley Online Library on [29/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
   
|
17
SCHULTZ et al.
Silva, S. M., Townsend Peterson, A., Carneiro, L., Burlamaqui, T. C. T.,
Ribas, C. C., Sousa-Neves, T., Miranda, L. S., Fernandes, A. M.,
D'Horta, F. M., Araújo-Silva, L. E., Batista, R., Bandeira, C. H. M. M.,
Dantas, S. M., Ferreira, M., Martins, D. M., Oliveira, J., Rocha, T. C.,
Sardelli, C. H., Thom, G., … Aleixo, A . (2019). A dynamic continen-
tal moisture gradient drove Amazonian bird diversification. Science
Advances, 5(7), eaat5752. https:// doi. org/ 10 . 1126/ scia dv. aat5752
Smith, B. T., McCormack, J. E., Cuervo, A. M., Hickerson, M. J., Aleixo, A.,
Cadena, C. D., Pérez-Emán, J., Burney, C. W., Xie, X., Harvey, M. G.,
Faircloth, B. C., Glenn, T. C., Derr yberry, E. P., Prejean, J., Fields, S.,
& Brumfield, R. T. (2014). The drivers of tropical speciation. Nature,
515 (7527), 406–409. https:// d oi. org/ 10. 1038/ natu r e13687
Tamura, L. N., Almeida, R. P., Galeazzi, C. P., Freitas, B. T., Ianniruberto,
M., & Prado, A. H. (2019). Upper-bar deposits in large Amazon riv-
ers: Occurrence, morphology and internal structure. Sedimentary
Geology, 387, 1–17. h t t p s : / / d o i . o r g / 1 0 . 1 0 1 6 / j . s e d g e o . 2 0 1 9 . 0 4 . 0 0 6
Thom, G., Amaral, F. R., Hickerson, M. J., Aleixo, A ., Araujo-Silva, L. E.,
Ribas, C. C., Choueri, E., & Miyaki, C. Y. (2018). Phenotypic and
genetic structure support gene flow generating gene tree discor-
dances in an Amazonian floodplain endemic species. Systematic
Biology, 67(June), 700 718. h t t p s : / / d o i . o r g / 1 0 . 1 0 9 3 / s y s b i o / s y y 0 0 4
Thom, G., Ribas, C. C., Schultz, E., Aleixo, A., & Miyaki, C. Y. (2022).
Population dynamics of Amazonian floodplain forest species sup-
port spatial variation on genetic diversity but not range expansions
through time. Journal of Biogeography, 49(10), 1891–1901. h tt ps ://
d o i . o r g / 1 0 . 1 1 1 1 / j b i . 1 4 4 7 8
Thom, G., Xue, A. T., Sawakuchi, A. O., Ribas, C. C., Hickerson, M. J.,
Aleixo, A., & Miyaki, C. (2020). Quaternary climate changes as spe-
ciation drivers in the Amazon floodplains. Science Advances, 6(11),
eaax4718. h t t p s : / / d o i . o r g / 1 0 . 1 1 2 6 / s c i a d v . a a x 4 7 1 8
Venticinque, E., Forsberg, B., Barthem, R., Petry, P., Hess, L., Mercado, A.,
Cañas, C., Montoya, M., Durigan, C., & Goulding, M. (2016). An ex-
plicit GIS-based river basin framework for aquatic ecosystem con-
servation in the Amazon. Earth System Science Data, 8(2), 651–661.
h t t p s : / / d o i . o r g / 1 0 . 5 1 9 4 / e s s d - 8 - 6 5 1 - 2 0 1 6
Wickham, H. (2016). ggplot2 elegant graphics for data analysis (use R!).
Springer.
Wittmann, F., Householder, J. E., Piedade, M. T. F., Schöngart, J.,
Demarchi, L. O., Quaresma, A. C., & Junk, W. J. (2022). A review of
the ecological and biogeographic differences of Amazonian flood-
plain forests. Water, 14(21), 3360. h t t p s : / / d o i . o r g / 1 0 . 3 3 9 0 / w 1 4 2 1
3360
SUPPORTING INFORMATION
Additional supporting information can be found online in the
Supporting Information section at the end of this article.
How to cite this article: Schultz, E. D., Thom, G., Zuquim, G.,
Hickerson, M. J., Tuomisto, H., & Ribas, C. C. (2023). Habitat
specialization predicts demographic response and
vulnerability of floodplain birds in Amazonia. Molecular
Ecology, 00, 1–17. http s://doi.org/10.1111/mec .17221
1365294x, 0, Downloaded from https://onlinelibrary.wiley.com/doi/10.1111/mec.17221 by Pennsylvania State University, Wiley Online Library on [29/11/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Amazonian floodplain forests along large rivers consist of two distinct floras that are traced to their differentiated sediment- and nutrient-rich (várzea) or sediment- and nutrient-poor (igapó) environments. While tree species in both ecosystems have adapted to seasonal floods that may last up to 270–300 days year−1, ecosystem fertility, hydrogeomorphic disturbance regimes, water shortage and drought, fire, and even specific microclimates are distinct between both ecosystems and largely explain the differences in forest productivity and taxonomic composition and diversity. Here, we review existing knowledge about the influence of these environmental factors on the tree flora of both ecosystems, compare species composition and diversity between central Amazonian várzeas and igapós, and show that both ecosystems track distinct species life-history traits. The ecosystem-level and taxonomic differences also largely explain the biogeographic connections of várzeas and igapós to other Amazonian and extra-Amazonian ecosystems. We highlight the major evolutionary force of large-river wetlands for Amazonian tree diversity and explore the scenarios by which the large number of Amazonian floodplain specialist tree species might even contribute to the gamma diversity of the Amazon by generating new species. Finally, we call attention to the urgent need of an improved conservation of Amazonian várzea and igapó ecosystems and their tree species.
Article
Full-text available
Spatial arrangement of distinct Amazonian environments through time and its effect on specialized biota remain poorly known, fueling long-lasting debates about drivers of biotic diversification. We address the late Quaternary sediment deposition that assembled the world's largest seasonally flooded ecosystems. Genome sequencing was used to reconstruct the demographic history of bird species specialized in either early successional vegetation or mature floodplain forests. Sediment deposition that built seasonally flooded habitats accelerated throughout the Holocene (last 11,700 years) under sea level highstand and intensification of the South American Monsoon, at the same time as global increases in atmospheric methane concentration. Bird populations adapted to seasonally flooded habitats expanded due to enlargement of Amazonian river floodplains and archipelagos. Our findings suggest that the diversification of the biota specialized in seasonally flooded habitats is coupled to sedimentary budget changes of large rivers, which rely on combined effects of sea level and rainfall variations.
Article
Full-text available
Over a century after Wallace’s proposition of the riverine barrier hypothesis, the role of rivers in the diversification of species remains a matter of interest in Amazon biogeography. Amazonian rivers, in particular those large and fast flowing rivers, are widely recognized to act as barriers to the dispersal of some organisms. However, the extent to which primate species changes across interfluves (β-diversity) in response to river features remains to be explicitly tested. In this study, we examine how river characteristics affect the variation in taxonomic and phylogenetic β-diversity to elucidate the ecological processes behind the differences in primate assemblages between neighboring interfluves in the Brazilian Amazonia. We compiled International Union for Conservation of Nature (IUCN) distribution maps of 80 river-bounded primate species in 10 interfluves separated by major rivers throughout the Brazilian Amazonia. We assessed both the taxonomic (disregarding phylogenetic relationships between species) and phylogenetic β-diversities. We applied multiple linear models to evaluate whether annual discharge, sinuosity, and reflectance (as a proxy for amount of sediments) in each river or river section that separates neighboring interfluves make rivers effective barriers to primates. We found that mean discharge has a positive effect, while both sinuosity and amount of sediments have negative effects on primate β-diversity. These variables have significant effects on total taxonomic and phylogenetic β-diversity between neighboring interfluves, and their species turnover components. River features, however, have no effect on species richness differences. Genera are capable of traversing almost all interfluves, but species are replaced by others in opposite interfluves. Discharge affected both small- and large-sized primate total β-diversity, but sinuosity only affected large-sized primate assemblages in neighboring interfluves. Our results indicate that although Amazonian rivers act as barriers for many primate species, this barrier effect seems limited to the short/medium time scales, as primate lineages are able to cross them over long-time scales.
Article
Full-text available
Large Amazonian rivers impede dispersal for many species, but lowland river networks frequently rearrange, thereby altering the location and effectiveness of river barriers through time. These rearrangements may promote biotic diversification by facilitating episodic allopatry and secondary contact among populations. We sequenced genome-wide markers to evaluate the histories of divergence and introgression in six Amazonian avian species complexes. We first tested the assumption that rivers are barriers for these taxa and found that even relatively small rivers facilitate divergence. We then tested whether species diverged with gene flow and recovered reticulate histories for all species, including one potential case of hybrid speciation. Our results support the hypothesis that river rearrangements promote speciation and reveal that many rainforest taxa are micro-endemic, unrecognized, and thus threatened with imminent extinction. We propose that Amazonian hyper-diversity originates partly from fine-scale barrier displacement processes-including river dynamics-which allow small populations to differentiate and disperse into secondary contact.
Article
Full-text available
Hydroelectric dams represent an important threat to seasonally flooded environments in the Amazon basin. We aimed to evaluate how a dam in the Madeira River, one of the largest tributaries of the Amazonas River, affected floodplain avifauna. Bird occurrence was recorded through simultaneous passive acoustic monitoring in early successional vegetation and floodplain forest downstream from the dam and upstream in sites impacted by permanent flooding after dam reservoir filling. Species were identified through manual inspection and semi-automated classification of the recordings. To assess the similarity in vegetation between downstream and upstream sites, we used Landsat TM/ETM+ composite images from before (2009-2011) and after (2016-2018) reservoir filling. Downstream and upstream floodplain forest sites were similar before, but not after dam construction. Early successional vegetation sites were already different before dam construction. We recorded 195 bird species. While species richness did not differ between upstream and downstream sites, species composition differed significantly. Ten species were indicators of early successional vegetation upstream, and four downstream. Ten species were indicators of floodplain forest upstream, and 31 downstream. Seven of 24 floodplain specialist species were detected by the semi-automated classification only upstream. While we found some bird species characteristic of early successional vegetation in the upstream sites, we did not find most species characteristic of tall floodplain forest. Predominantly carnivorous, insectivorous, and nectarivorous species appear to have been replaced by generalist and widely distributed species. KEYWORDS: Amazon; ecoacoustics; indicator species; passive acoustic monitoring
Article
Full-text available
Large alluvial rivers transport water and sediment across continents and shape lowland landscapes. Repeated glacial cycles have dominated Earth's recent climate, but it is unclear whether these rivers are sensitive to such rapid changes. The Amazon River system, the largest and highest‐discharge in the world, features extensive young terraces that demonstrate geologically rapid change temporally correlated with changes in runoff from Quaternary climate cycles. To test the plausibility of a causal relationship, we use a simple model to estimate from empirical measurements how quickly a river profile responds to changes in discharge or sediment supply. Applying this model to data from 30 gauging stations along alluvial rivers throughout the Brazilian Amazon, we find that many rivers of the Amazon basin can respond faster than glacially induced changes in runoff or sediment flux. The Amazon basin is unusually responsive compared to other large river systems due to its high discharge and sediment flux, narrow floodplains, and low slopes. As a result, we predict that the Amazon basin has been highly dynamic during Quaternary glacial cycles, with cyclical aggradation and incision of lowland rivers driving repeated habitat and environmental change throughout the region. This dynamic landscape may have contributed to the exceptional biodiversity of the region and patterns of ancient human settlement.
Article
Understanding the factors that govern variation in genetic structure across species is key to the study of speciation and population genetics. Genetic structure has been linked to several aspects of life history, such as foraging strategy, habitat association, migration distance, and dispersal ability, all of which might influence dispersal and gene flow. Comparative studies of population genetic data from species with differing life histories provide opportunities to tease apart the role of dispersal in shaping gene flow and population genetic structure. Here, we examine population genetic data from sets of bird species specialized on a series of Amazonian habitat types hypothesized to filter for species with dramatically different dispersal abilities: stable upland forest, dynamic floodplain forest, and highly dynamic riverine islands. Using genome-wide markers, we show that habitat type has a significant effect on population genetic structure, with species in upland forest, floodplain forest, and riverine islands exhibiting progressively lower levels of structure. Although morphological traits used as proxies for individual-level dispersal ability did not explain this pattern, population genetic measures of gene flow are elevated in species from more dynamic riverine habitats. Our results suggest that the habitat in which a species occurs drives the degree of population genetic structuring via its impact on long-term fluctuations in levels of gene flow, with species in highly dynamic habitats having particularly elevated gene flow. These differences in genetic variation across taxa specialized in distinct habitats may lead to disparate responses to environmental change or habitat-specific diversification dynamics over evolutionary time scales.
Article
Abstract Although vicariant processes are expected to leave similar genomic signatures among co‐distributed taxa, ecological traits such as habitat and stratum can influence genetic divergence within species. Here, we combined landscape history and habitat specialization to understand the historical and ecological factors responsible for current levels of genetic divergence in three species of birds specialized in seasonally flooded habitats in muddy rivers which are widespread in the Amazon basin but have isolated populations in the Rio Branco. Populations of the White‐bellied Spinetail (Mazaria propinqua), Lesser Wagtail‐tyrant (Stigmatura napensis), and Bicolored Conebill (Conirostrum bicolor) are currently isolated in the Rio Branco by the black‐waters of the lower Rio Negro, offering a unique opportunity to test the effect of river color as a barrier to gene flow. We used ultraconserved elements (UCEs) to test alternative hypotheses of population history in a comparative phylogeographic approach by modeling genetic structure, demographic history, and testing for shared divergence time among co‐distributed taxa. Our analyses revealed that (i) all three populations from the Rio Branco floodplains are genetically distinct from other populations along the Amazon River floodplains; (ii) these divergences are the result of at least two distinct events, consistent with species habitat specialization; and (iii) the most likely model of population evolution includes lower population connectivity during the Late Pleistocene transition (~250 ka), with gene flow being completely disrupted after the Last Glacial Maximum (~21 ka). Our findings highlight how landscape evolution modulates population connectivity in habitat specialist species and how organisms can have different responses to the same historical processes of environmental change, depending on their habitat affinity.
Article
Aim The distribution and connectivity of floodplain environments along major Amazonian rivers changed considerably over time following physiographic and climatic events. However, how historical alteration in floodplain habitats affected endemic species' demography and diversification is not fully understood. We tested if historical demographic changes of populations occurring on the floodplains of a major Amazon Basin tributary could be associated with range expansions from upper and middle sections of the river, following the establishment of widespread river‐created environments during the Late Pleistocene and Holocene. Location Solimões River, Western Amazon, South America. Taxon Myrmoborus lugubris , Thamnophilus cryptoleucus and Myrmotherula assimilis . Methods We explored spatial patterns of genetic diversity and connectivity among 13 sampled localities using thousands of UltraConserved Elements. We used a total of 73 individuals of three species with populations restricted to the Solimões River. Range expansions were tested with spatial explicit methods designed to capture variations in allele frequency over space. To test whether the occupation of dynamic habitats could predict spatial patterns of genetic diversity, we quantified habitat preference for the three analysed species by creating a metric of habitat dynamicity over time. Results We found considerable variation in the spatial distribution of the genetic diversity between studied taxa rejecting shared population range expansions related to historical regionalized changes in habitat availability. Species with higher levels of specialization to dynamic environments have a more heterogeneous distribution of genetic diversity and reduced levels of gene flow across space. Main conclusions The demographic expansions along the Solimões River might be linked to spatially homogeneous oscillation in the distribution of floodplain environments, promoting effective population size changes but not range expansion. Habitat specificity might be a good predictor of population connectivity along the Amazonian floodplains.
Article
Amazonia has a very high, although still incompletely known, species diversity distributed over a mosaic of heterogeneous habitats that are also poorly characterized. As a result of this multi‐faceted complexity, Amazonia poses a great challenge to geogenomic approaches that strive to find causal links between Earth's geological history and biotic diversification, including the use of genomic data to solve geologic problems. This challenge is even greater because of the need for interdisciplinary approaches despite the difficulties of working across disciplines, where misinterpretations of the literature in disparate research fields can produce unrealistic scenarios of biotic‐geologic linkages. The exchange of information and the joint work of geologists and biologists are essential for building stronger and more realistic hypotheses about how past climate may have affected the distribution and connectivity among populations, how the evolution of drainage networks influenced biotic diversification, and how ecological traits and species interactions currently define the spatial organization of biodiversity, and thus how this organization has changed in the past and may change in the future. The heterogeneity of Amazonia and the different effects of historical processes over its distinct regions and ecosystems have to be more completely recognized in biogeography, conservation; and policymaking. In this perspective, we provide examples of geological, climatological; and ecological information relevant to studies of biotic diversification in Amazonia, where recent advances (and their limitations) may not be apparent to researchers in other fields. The three examples, which include the limitations of climate model outputs, the complicated evolution of river drainages; and the complex link between species and their habitats modulated by ecological specialization, are a small subsample intended to illustrate the urgency for more integrated interdisciplinary approaches.